**EPA
United States
Environmental Protection
Agency
      Third External Review Draft of
      Air Quality Criteria for
      Particulate Matter (April, 2002)
      Volume II

-------
                                                     EPA/600/P-99/002aC
                                                            April 2002
                                                Third External Review Draft
Air Quality Criteria for Participate Matter
                     Volume II
       National Center for Environmental Assessment-RTP Office
               Office of Research and Development
              U.S. Environmental Protection Agency
                  Research Triangle Park, NC

-------
 1                                         DISCLAIMER
 2
 3           This document is an external review draft for review purposes only and does not constitute
 4      U.S. Environmental Protection Agency policy.  Mention of trade names or commercial products
 5      does not constitute endorsement or recommendation for use.
 6
 7
 8                                           PREFACE
 9
10           National Ambient Air Quality Standards (NAAQS) are promulgated by the United States
11      Environmental Protection Agency (EPA) to meet requirements set forth in Sections 108 and 109
12      of the U.S. Clean Air Act (CAA). Sections 108 and 109 require the EPA Administrator (1) to list
13      widespread air pollutants that reasonably may be expected to endanger public health or welfare;
14      (2) to issue air quality criteria for them that assess the latest available scientific information on
15      nature and effects of ambient exposure to them; (3) to set "primary" NAAQS to protect human
16      health with adequate margin of safety and to set "secondary" NAAQS to protect against welfare
17      effects (e.g., effects on vegetation, ecosystems, visibility, climate, manmade materials, etc.); and
18      (5) to periodically (every 5 years) review and revise, as appropriate, the criteria and NAAQS for
19      a given listed pollutant or class of pollutants.
20           The original NAAQS for particulate matter (PM), issued in 1971 as "total suspended
21      parti culate" (TSP) standards, were revised in  1987 to focus on protecting against human health
22      effects associated with exposure to ambient PM less than 10 microns (< 10 //m) that are capable
23      of being deposited in thoracic (tracheobronchial and alveolar) portions of the lower respiratory
24      tract. Later periodic reevaluation of newly available scientific information, as presented in the
25      last previous version of this "Air Quality Criteria for Parti culate Matter" document published in
26      1996, provided key  scientific bases for PM NAAQS decisions published in July 1997. More
27      specifically, the PM10 NAAQS set in 1987 (150 //g/m3, 24-h; 50 //g/m3, annual average) were
28      retained in modified form and new standards  (65 //g/m3, 24-h; 15 //g/m3, annual average) for
29      particles <2.5 //m (PM25) were promulgated in July 1997.
        April 2002                                 Il-ii        DRAFT-DO NOT QUOTE OR CITE

-------
 1           This Third External Review Draft of revised Air Quality Criteria for Particulate Matter
 2      assesses new scientific information that has become available mainly between early 1996 through
 3      December 2001. The present draft is being released for public comment and review by the Clean
 4      Air Scientific Advisory Committee (CASAC) to obtain  comments on the organization and
 5      structure of the document, the issues addressed, the approaches employed in assessing and
 6      interpreting the newly available information on PM exposures and effects, and the key findings
 7      and conclusions arrived at as a consequence of this assessment.  Public comments and CASAC
 8      review recommendations will be taken into account in making any appropriate further revisions
 9      to this document for incorporation into a final draft. Evaluations contained in the present
10      document will be drawn on to provide inputs to associated PM Staff Paper analyses prepared by
11      EPA's Office of Air Quality Planning and Standards (OAQPS) to pose alternatives for
12      consideration by the EPA Administrator with regard to proposal and, ultimately, promulgation of
13      decisions on potential retention or revision of the current PM NAAQS.
14           Preparation of this document was coordinated by staff of EPA's National Center for
15      Environmental Assessment in Research Triangle Park (NCEA-RTP). NCEA-RTP scientific
16      staff, together with experts from other EPA/ORD laboratories and academia, contributed to
17      writing of document chapters; and earlier drafts of this document were reviewed by experts from
18      federal and state government agencies, academia, industry, and NGO's for use by EPA in support
19      of decision making on potential public health and environmental risks  of ambient PM.  The
20      document describes the nature, sources, distribution, measurement, and concentrations of PM in
21      outdoor (ambient) and indoor environments. It also evaluates the latest data on human exposures
22      to ambient PM and consequent health effects in exposed human populations (to support decision
23      making regarding primary, health-related PM NAAQS).  The document also evaluates ambient
24      PM environmental effects on vegetation and ecosystems, visibility, and man-made materials, as
25      well as atmospheric PM effects on climate change processes associated with alterations in
26      atmospheric transmission of solar radiation or its reflectance from the Earth's surface or
27      atmosphere (to support decision making on secondary PM NAAQS).
28           The NCEA of EPA acknowledges the contributions provided by  authors, contributors, and
29      reviewers and the diligence of its staff and contractors in the preparation of this document.
        April 2002                                II-iii        DRAFT-DO NOT QUOTE OR CITE

-------
               Air Quality Criteria for Particulate Matter


                              VOLUME I


   EXECUTIVE SUMMARY	E-l

   1.  INTRODUCTION  	   1-1

   2.  PHYSICS, CHEMISTRY, AND MEASUREMENT OF
      PARTICULATE MATTER	   2-1

   3.  CONCENTRATIONS, SOURCES, AND EMISSIGNS OF
      ATMOSPHERIC PARTICULATE MATTER  	   3-1
      Appendix 3 A:  Spatial and Temporal Variability of the Nationwide
                   AIRS PM25 and PM10.25 Data Sets	  3A-1
      Appendix 3B:  Aerosol Composition Data from the Speciation
                   Network	  3B-1
      Appendix 3C:  Organic Composition of Particulate Matter	  3C-1
      Appendix 3D:  Composition of Particulate Matter  Source Emissions ....  3D-1

   4.  ENVIRONMENTAL EFFECTS OF PARTICULATE MATTER  	   4-1
      Appendix 4A: Colloquial and Latin Names  	  4A-1

   5.  HUMAN EXPOSURE TO PARTICULATE MATTER AND ITS
      CONSTITUENTS  	   5-1
April 2002                          Il-iv        DRAFT-DO NOT QUOTE OR CITE

-------
                Air Quality Criteria for Particulate Matter
                                  (cont'd)
                                VOLUME II
   6.   DOSIMETRY OF PARTICULATE MATTER	   6-1

   7.   TOXICOLOGY OF PARTICULATE MATTER IN HUMANS AND
       LABORATORY ANIMALS 	   7-1

   8.   EPIDEMIOLOGY OF HUMAN HEALTH EFFECTS FROM
       AMBIENT PARTICULATE MATTER  	   8-1
       Appendix 8A:  Short-Term PM Exposure-Mortality  Studies:
                    Summary Table 	  8A-1
       Appendix 8B:  Particulate Matter-Morbidity Studies:
                    Summary Tables  	  8B-1

   9.   INTEGRATIVE SYNTHESIS	   9-1
       Appendix 9A:  Key Quantitative Estimates of Relative Risk for
                    Particulate Matter-Related Health Effects Based on
                    Epidemiologic Studies of U.S. and Canadian Cities
                    Assessed in the 1996 Particulate Matter Air Quality
                    Criteria Document	  9A-1
April 2002                           II-v         DRAFT-DO NOT QUOTE OR CITE

-------
                                  Table of Contents

                                                                                   Page

List of Tables	II-xiv
List of Figures  	II-xx
Authors, Contributors, and Reviewers	  D-xxvii
U.S. Environmental Protection Agency Project Team for Development of
  Air Quality Criteria for Particulate Matter	II-xxxvi
Abbreviations and Acronyms	D-xxxix

6.  DOSIMETRY OF PARTICULATE MATTER	6-1
   6.1  INTRODUCTION	6-1
        6.1.1   Size Characterization of Inhaled Particles 	6-3
        6.1.2   Structure of the Respiratory Tract	6-4
   6.2  PARTICLE DEPOSITION	6-6
        6.2.1   Mechanisms of Deposition	6-6
        6.2.2   Deposition Patterns in the Human Respiratory Tract	6-8
               6.2.2.1   Total Respiratory Tract Deposition	6-8
               6.2.2.2   Deposition in the Extrathoracic Region	6-12
               6.2.2.3   Deposition in the Tracheobronchial and Alveolar Regions	6-16
               6.2.2.4   Local Distribution of Deposition	6-16
               6.2.2.5   Deposition of Specific Size Modes of Ambient Aerosol	6-20
        6.2.3   Biological Factors Modulating Deposition	6-23
               6.2.3.1   Gender	6-23
               6.2.3.2   Age	6-25
               6.2.3.3   Respiratory Tract Disease  	6-28
               6.2.3.4   Anatomical Variability	6-30
        6.2.4   Interspecies Patterns of Deposition	6-32
   6.3  PARTICLE CLEARANCE AND TRANSLOCATION	6-39
        6.3.1   Mechanisms and Pathways of Clearance 	6-39
               6.3.1.1   Extrathoracic Region	6-41
               6.3.1.2   Tracheobronchial Region	6-42
               6.3.1.3   Alveolar Region	6-42
        6.3.2   Clearance Kinetics  	6-44
               6.3.2.1   Extrathoracic Region	6-44
               6.3.2.2   Tracheobronchial Region	6-44
               6.3.2.3   Alveolar Region	6-46
        6.3.3   Interspecies Patterns of Clearance  	6-50
        6.3.4   Factors Modulating Clearance 	6-51
               6.3.4.1   Age	6-51
               6.3.4.2   Gender	6-51
               6.3.4.3   Physical Activity 	6-51
               6.3.4.4   Respiratory Tract Disease  	6-52
   6.4  PARTICLE OVERLOAD	6-53
April 2002                                H-vi        DRAFT-DO NOT QUOTE OR CITE

-------
                              Table of Contents
                                    (cont'd)
                                                                          Page

   6.5   COMPARISON OF DEPOSITION AND CLEARANCE PATTERNS
        OF PARTICLES ADMINISTERED BY INHALATION AND
        INTRATRACHEAL INSTILLATION	6-55
   6.6   MODELING THE DISPOSITION OF PARTICLES IN THE
        RESPIRATORY TRACT	6-57
        6.6.1  Modeling Deposition, Clearance, and Retention 	6-57
        6.6.2  Models To Estimate Retained Dose	6-64
        6.6.3  Fluid Dynamics Models for Deposition Calculations	6-67
   6.7   SUMMARY AND CONCLUSIONS	6-73
   REFERENCES  	6-76

7.  TOXICOLOGY OF PARTICIPATE MATTER IN HUMANS AND LABORATORY
   ANIMALS	7-1
   7.1   INTRODUCTION	7-1
    .2  RESPIRATORY EFFECTS OF PARTICIPATE MATTER IN HEALTHY
       HUMANS AND LABORATORY ANIMALS: IN VIVO EXPOSURES  	7-2
       7.2.1  Ambient Combustion-Related and Surrogate Particulate Matter	7-7
             7.2.1.1  Ambient Particulate Matter	7-15
             7.2.1.2  Diesel Particulate Matter	7-18
             7.2.1.3  Complex Combustion-Related Particles 	7-22
       7.2.2  Acid Aerosols	7-25
       7.2.3  Metal Particles, Fumes, and Smoke 	7-27
       7.2.4  Ambient Bioaerosols	7-32
   7.3  CARDIOVASCULAR AND SYSTEMIC EFFECTS OF PARTICULATE
       MATTER IN HUMANS AND LABORATORY ANIMALS: IN VIVO
       EXPOSURES 	7-34
   7.4  SUSCEPTIBILITY TO THE EFFECTS OF PARTICULATE MATTER
       EXPOSURE 	7-45
       7.4.1  Pulmonary Effects of Particulate Matter in Compromised Hosts	7-45
       7.4.2  Genetic Susceptibility to Inhaled Particles and their Constituents	7-50
       7.4.3  Effect of Particulate Matter on Allergic Hosts	7-52
       7.4.4  Resistance to Infectious Disease	7-59
   7.5  PARTICULATE MATTER TOXICITY AND PATHOPHYSIOLOGY:
       IN VITRO EXPOSURES	7-60
       7.5.1  Introduction	7-60
       7.5.2  Experimental Exposure Data  	7-61
             7.5.2.1  Ambient Particles	7-62
             7.5.2.2  Comparison of Ambient and Combustion-Related Surrogate
                     Particles 	7-72
       7.5.3  Potential Cellular and Molecular Mechanisms	7-75
             7.5.3.1  Reactive Oxygen Species	7-75
             7.5.3.2  Intracellular Signaling Mechanisms	7-80

April 2002                            H-vii       DRAFT-DO NOT QUOTE OR CITE

-------
                                Table of Contents
                                      (cont'd)
               7.5.3.3   Other Potential Cellular and Molecular Mechanisms  	7-84
        7.5.4   Specific Particle Size and Surface Area Effects	7-86
        7.5.5   Pathophysiological Mechanisms for the Effects of Low Concentrations
               of Paniculate Air Pollution	7-90
               7.5.5.1   Direct Pulmonary Effects	7-91
               7.5.5.2   Systemic Effects Secondary to Lung Injury	7-92
               7.5.5.3   Direct Effects on the Heart	7-94
   7.6  RESPONSES TO PARTICULATE MATTER AND GASEOUS
        POLLUTANT MIXTURES 	7-95
   7.7  SUMMARY 	7-105
        7.7.1   Biological Plausibility	7-105
               7.7.1.1   Link Between Specific Particulate Matter Components
                           and Health Effects 	7-105
               7.7.1.2   Susceptibility	7-110
        7.7.2   Mechanisms of Action 	7-110
   REFERENCES  	7-111

   EPIDEMIOLOGY OF HUMAN HEALTH EFFECTS
   FROM AMBIENT PARTICULATE MATTER	8-1
   8.1  INTRODUCTION	8-1
        8.1.1   Types of Epidemiology Studies Reviewed	8-2
        8.1.2   Confounding and Effect Modification 	8-4
        8.1.3   Selection of Studies for Review and Ambient PM Increments Used
               to Report Risk Estimates	8-10
   8.2  MORTALITY EFFECTS OF PARTICULATE MATTER EXPOSURE 	8-12
        8.2.1   Introduction	8-12
        8.2.2   Mortality Effects of Short-Term Particulate Matter Exposure	8-13
               8.2.2.1   Summary of 1996 Particulate Matter Criteria Document
                       Findings and Key Issues	8-13
               8.2.2.2   Introduction to Newly Available Information on Short-Term
                       Mortality Effects  	8-17
               8.2.2.3   New Multi-City Studies	8-25
               8.2.2.4   The Role of Particulate Matter Components	8-38
               8.2.2.5   New Assessments of Cause-Specific Mortality	8-61
               8.2.2.6   Salient Points Derived from Summarization of Studies of
                       Short-Term Particulate Matter Exposure Effects on Mortality  . . . 8-66
        8.2.3   Mortality Effects of Long-Term Exposure to  Ambient Particulate
               Matter 	8-69
               8.2.3.1   Studies Published Prior to the 1996 Particulate Matter
                       Criteria Document	8-69
April 2002                               H-viii        DRAFT-DO NOT QUOTE OR CITE

-------
                                 Table of Contents
                                       (cont'd)
               8.2.3.2  Prospective Cohort Analyses of Chronic Particulate Matter
                       Exposure Mortality Effects Published Since the 1996
                       Particulate Matter Air Quality Criteria Document	8-73
               8.2.3.3  Studies by Particulate Matter Size-Fraction and Composition  ... 8-96
               8.2.3.4  Population-Based Mortality Studies in Children	8-101
               8.2.3.5  Salient Points Derived from Analyses of Chronic Particulate
                       Matter Exposure Mortality Effects 	8-105
     .3  MORBIDITY EFFECTS OF PARTICULATE MATTER EXPOSURE	8-108
        8.3.1   Cardiovascular Effects Associated with Acute Ambient Particulate
               Matter Exposure 	8-108
               8.3.1.1  Introduction 	8-108
               8.3.1.2  Summary of Key Findings on Cardiovascular Morbidity from
                       the 1996 Particulate Matter Air quality Criteria Document	8-109
               8.3.1.3  New Particulate Matter-Cardiovascular Morbidity Studies	8-110
               8.3.1.4  Issues in the Interpretation of Acute Cardiovascular Effects
                       Studies 	8-132
        8.3.2   Effects of Short-Term Particulate Matter Exposure on the Incidence of
               Respiratory Hospital Admissions and Medical Visits  	8-134
               8.3.2.1  Introduction 	8-134
               8.3.2.2  Summary of Key Respiratory Hospital Admissions
                       Findings from the 1996 Particulate Matter Air Quality
                       Criteria Document	8-135
               8.3.2.3  New Respiratory-Related Hospital Admissions Studies 	8-135
               8.3.2.4  Key New Respiratory Medical Visits Studies 	8-145
               8.3.2.5  Identification of Potential Susceptible Subpopulations	8-147
               8.3.2.6  Summary of Key Findings on Acute Particulate Matter
                       Exposure and Respiratory-Related Hospital Admissions
                       and Medical Visits	8-150
        8.3.3   Effects of Particulate Matter Exposure on Lung Function and
               Respiratory Symptoms 	8-154
               8.3.3.1  Effects of Short-Term Particulate Matter Exposure on
                       Lung Function and Respiratory Symptoms	8-155
               8.3.3.2  Long-Term Particulate Matter Exposure Effects on Lung
                       Function and Respiratory Symptoms	8-166
     .4  DISCUSSION OF EPIDEMIOLOGIC STUDIES ON HEALTH EFFECTS
        OF AMBIENT PARTICULATE MATTER	8-173
        8.4.1   Introduction	8-173
        8.4.2   Assessment of Confounding by Co-Pollutants	8-177
               8.4.2.1  Introduction 	8-177
               8.4.2.2  Issues  	8-181
April 2002                                H-ix        DRAFT-DO NOT QUOTE OR CITE

-------
                                 Table of Contents
                                       (cont'd)
               8.4.2.2  Assessments of Confounding Using Multi-Pollutant Models
                       with Observed Gases	8-197
               8.4.2.3  Assessment of Confounding in Multi-City Studies:
                       Pooling Effects	8-201
               8.4.2.4  Assessment of Confounding in Multi-City Studies:
                       Regression 	8-203
               8.4.2.5  Assessment of Confounding Based on Exposure  	8-207
               8.4.2.6  Assessment of Confounding by Factor Analysis	8-217
               8.4.2.7  Simulation Analysis of Confounding	8-218
               8.4.2.8  Discussion 	8-219
        8.4.3   Role of Particulate Matter Components 	8-220
               8.4.3.1  Fine- and Coarse-Particle Effects on Mortality	8-220
               8.4.3.2  PM10, PM2 5 (Fine), and PM10_2 5 (Coarse) Particulate Matter
                       Effects on Morbidity  	8-231
        8.4.4   The Question of Lags  	8-237
        8.4.5   New Assessments of Mortality Displacement  	8-243
        8.4.6   Concentration-Response Relationships for Ambient PM	8-245
        8.4.7   New Assessments of Consequences of Measurement Error	8-250
               8.4.7.1  Theoretical Framework for Assessment of Measurement
                       Error	8-250
               8.4.7.2  Spatial Measurement Error Issues That May Affect the
                       Interpretation of Multi-Pollutant Models with Gaseous
                       Co-Pollutants	8-256
               8.4.7.3  Measurement Error and the Assessment of Confounding by
                       Co-Pollutants in Multi-Pollutant Models	8-263
               8.4.7.5  Air Pollution Exposure Proxies in Long-Term Mortality
                       Studies 	8-264
        8.4.9   Heterogeneity of Particulate Matter Effects Estimates	8-270
               8.4.9.1  Evaluation of Heterogeneity of Particulate Matter Mortality
                       Effect Estimates	8-271
               8.4.9.2  Comparison of Spatial Relationships in the NMMAPS and
                       Cohort Reanalyses Studies	8-275
               8.4.9.3  Epidemiologic Studies of Ambient Air Pollution
                       Interventions	8-277
   8.5  KEY FINDINGS AND CONCLUSIONS DERIVED FROM
        PARTICULATE MATTER EPIDEMIOLOGY STUDIES 	8-283
   REFERENCES  	8-289

   APPENDIX 8A:  Short-Term PM Exposure-Mortality Studies: Summary Table ....  8A-1

   APPENDIX 8B: Particulate Matter-Morbidity Studies: Summary Tables  	8B-1

April 2002                                H-x         DRAFT-DO NOT QUOTE OR CITE

-------
                                Table of Contents
                                     (cont'd)
9.  INTEGRATIVE SYNTHESIS  	9-1
   9.1   INTRODUCTION	9-1
   9.2   BACKGROUND	9-3
         9.2.1   Basic Concepts  	9-3
         9.2.2   Particle Size Distributions	9-3
         9.2.3   Definitions of Particle Size Fractions	9-4
   9.3   CHARACTERIZATION OF EMISSION SOURCES  	9-10
   9.4   AMBIENT CONCENTRATIONS  	9-16
         9.4.1   Measurement of Particulate Matter	9-16
         9.4.2   Mass Concentrations	9-17
         9.4.3   Physical and Chemical Properties of Ambient PM	9-18
   9.5   AIR QUALITY MODEL DEVELOPMENT AND TESTING  	9-23
   9.6   EXPOSURE TO PARTICULATE MATTER AND COPOLLUTANTS  	9-27
         9.6.1   Central Site to Outdoor	9-28
               9.6.1.1   Exposure for Acute Epidemiology	9-28
               9.6.1.2   Exposure for Chronic Epidemiology	9-28
         9.6.2   Home Outdoor Concentrations Versus Ambient Concentrations
               Indoors and the Ambient Contribution to Total Personal Exposure	9-30
               9.6.2.1   Mass Balance Model 	9-30
               9.6.2.2   Separation of Total Personal Exposure into its Ambient
                       and Nonambient Components 	9-31
         9.6.3   Variability in the Relationship Between Concentrations and Personal
               Exposures 	9-35
         9.6.4   Exposure Relations for Co-Pollutants	9-35
         9.6.5   Summary	9-41
   9.7   EXPOSURE TO BIOLOGICALLY IMPORTANT CHARACTERISTICS
         OF PARTICULATE MATTER	9-41
         9.7.1   Exposure Relationships for Susceptible Subpopulations 	9-42
         9.7.2   Toxicologically Important Components of PM  	9-42
         9.7.3   Exposure-Measurement Techniques	9-43
         9.7.4   Comprehensive Studies to Determine Population Exposure 	9-44
         9.7.5   Air Pollutants Generated Indoors 	9-48
   9.8   DOSIMETRY: DEPOSITION AND FATE OF PARTICLES IN THE
         RESPIRATORY TRACT 	9-50
         9.8.1   Particle Deposition in the Respiratory Tract	9-50
         9.8.2   Particle Clearance and Translocation 	9-55
         9.8.3   Deposition and Clearance Patterns of Particles Administered by
               Inhalation Versus Intratracheal Instillation	9-57
         9.8.4   Inhaled Particles as Potential Carriers of Toxic Agents 	9-57
         9.8.5   Summary of Particle Dosimetry 	9-58
April 2002                              H-xi       DRAFT-DO NOT QUOTE OR CITE

-------
                                Table of Contents
                                      (cont'd)
   9.9   ASSESSMENT OF PARTICULATE-MATTER PROPERTIES LINKED
               TO HEALTH EFFECTS	9-60
         9.9.1  Introduction	9-60
         9.9.2  Specific Properties of Ambient PM Linked to Health Effects  	9-62
               9.9.2.1  Physical Properties	9-62
               9.9.2.2  Chemical Properties	9-64
               9.9.2.3  Summary  	9-67
         9.9.3  Chemical Components and Source Categories Associated with Health
               Effects in Epidemiologic Studies  	9-67
               9.9.3.1  Individual Chemical Species 	9-68
               9.9.3.2  Source Category Factors  	9-68
   9.10  SUSCEPTIBLE SUBPOPULATIONS  	9-70
         9.10.1 Introduction	9-71
         9.10.2 Preexisting Disease as a Risk Factor for Particulate Matter Health
               Effects	9-71
               9.10.2.3 Ambient PM Exacerbation of Respiratory Disease
                       Conditions 	9-73
               9.10.2.4 Ambient PM Exacerbation of Cardiovascular Disease
                       Conditions 	9-74
         9.10.3 Age-Related At-Risk Population Groups:  The Elderly and Children  	9-77
   9.11  MECHANISMS OF INJURY	9-80
   9.12  HEALTH EFFECTS OF AMBIENT PARTICULATE MATTER
         OBSERVED IN POPULATION STUDIES  	9-85
         9.12.1 Introduction	9-85
         9.12.2 Community-Health Epidemiologic Evidence for Ambient Particulate
               Matter Effects 	9-87
               9.12.2.1 Short-Term Particulate Matter Exposure Effects on
                       Mortality	9-90
               9.12.2.2 Relationships of Ambient Particulate Matter Concentrations
                       to Morbidity Outcomes  	9-116
               9.12.2.3 Methodological Issues	9-129
         9.12.3 Coherence of Reported Epidemiologic Findings  	9-136
   9.13  EVALUATION OF  STATISTICAL AND MEASUREMENT
         ERROR ISSUES  	9-138
         9.13.1 Errors Related to Concentration, Exposure, and Dose	9-138
               9.13.1.1 Opportunities for Error in the Use of Ambient PM
                       Concentration as a Surrogate for PM Dose in Epidemiologic
                       Studies 	9-139
         9.13.2 Possible Errors Related to Health and Epidemiology	9-146
         9.13.3 Apportioning Health Effects to PM (by size, chemical component,
               or source category) and Gaseous Co-Pollutants  	9-148

April 2002                               U-xii        DRAFT-DO NOT QUOTE OR CITE

-------
                               Table of Contents
                                     (cont'd)
   9.14 IMPLICATIONS OF HEALTH EFFECTS OF LONG-TERM
        EXPOSURES TO PARTICIPATE MATTER  	9-153
        9.14.1  Methodological Issues	9-153
        9.14.2  Overall Survival and Life Expectancy 	9-154
        9.14.3  Verification and Sensitivity Analyses	9-155
        9.14.4  Impact on Life-Expectancy	9-155
        9.14.5  Specific Causes of Death 	9-156
   REFERENCES  	9-157

   APPENDIX 9A:  Key Quantitative Estimates of Relative Risk for Particulate
                   Matter-Related Health Effects Based on Epidemiologic Studies
                   of U.S. and Canadian Cities Assessed in the 1996 Parti culate
                   Matter Air Quality Criteria Document	 9A-1
   REFERENCES  	 9A-6
April 2002                             H-xiii      DRAFT-DO NOT QUOTE OR CITE

-------
                                    List of Tables

Number                                                                            Page

6-1     Overview of Respiratory Tract Particle Clearance and Translocation
        Mechanisms 	6-40

7-1     Types of Particulate Matter Used in Toxicological Studies	7-4

7-2     Respiratory Effects of Ambient Parti culate Matter	7-8

7-3     Respiratory Effects of Complex Combustion-Related Parti culate Matter	7-9

7-4     Respiratory Effects of Surrogate Parti culate Matter	7-14

7-5     Respiratory Effects of Acid Aerosols in Humans and Laboratory Animals	7-26

7-6     Respiratory Effects of Metal Particles, Fumes, and Smoke in Humans and
        Laboratory Animals  	7-29

7-7     Respiratory Effects of Ambient Bioaerosols	7-33

7-8     Cardiovascular and Systemic Effects of Ambient and Combustion-Related
        Paniculate Matter	7-35

7-9     Physicochemical Properties of Parti culate Matter  	7-61

7-10    In Vitro Effects of Parti culate Matter and Parti culate Matter Constituents 	7-63

7-11    Numbers and Surface Areas of Monodisperse Particles of Unit Density of
        Different Sizes at a Mass Concentration of 10 //g/m3	7-87

7-12    Respiratory and Cardiovascular Effects of Mixtures  	7-97

8-1     Recent U.S. and Canadian Time-Series Studies of PM-Related Daily Mortality ... 8-18

8-2     Synopsis of Short-Term  Mortality Studies That Examined Relative Importance
        of PM25 and PM10.25  	8-40

8-3     Excess Total Mortality Risks Estimated to be Associated with Various
        Ambient Particle  Size-Related Indices	8-50

8-4     Summary of Parti culate Matter Chemical Components Analyzed in
        Recent Studies	8-52
April 2002                               II-xiv        DRAFT-DO NOT QUOTE OR CITE

-------
                                    List of Tables
                                        (cont'd)
Number
8-5      Summary of Source-Oriented Evaluations of Particulate Matter Components
         in Recent Studies	8-58

8-6      Comparison of Six Cities and American Cancer Society Study Findings
         from Original Investigators and Health Effects Institute Reanalysis 	8-75

8-7      Summary of Results from the Extended ACS  Study  	8-79

8-8      Relative Risk of Mortality from All Nonexternal Causes, by Sex and Air
         Pollutant, for an Alternative Covariate Model in the ASHMOG Study	8-87

8-9      Relative Risk of Mortality from Cardiopulmonary Causes, by Sex and  Air
         Pollutant, for an Alternative Covariate Model 	8-88

8-10     Relative Risk of Mortality from Lung Cancer by Air Pollutant and by Gender
         for an Alternative Covariate Model  	8-88

8-11     Comparison of Excess Relative Risks for Three Particle Metrics in the Male
         Subset of the AHSMOG Study	8-90

8-12     Comparison of Excess Relative Risks of Long-Term Mortality in the Harvard
         Six Cities, ACS, AHSMOG, and VA Studies  	8-94

8-13     Comparison of Estimated Relative Risks for All-Cause Mortality in Six
         U.S. Cities Associated with the Reported Inter-City Range of Concentrations
         of Various Parti culate Matter Metrics	8-97

8-14     Comparison of Reported SO4= and PM25 Relative Risks for Various Mortality
         Causes in the American Cancer Society Study	8-97

8-15     Comparison of Total Mortality Relative Risk Estimates and T-statistics for
         Particulate Matter Components in Three Prospective Cohort Studies	8-98

8-16     Comparison of Cardiopulmonary Mortality Relative Risk Estimates and
         T-statistics for Particulate Matter Components in Three Prospective Cohort
         Studies	8-99

8-17     Summary of Studies of PM10 or PM25 and Total CVD Hospital Visits  	8-111

8-18     Percent Increase in Hospital Admissions per lO-^g/m3 Increase in PM10 in
         14 U.S. Cities 	8-137

April 2002                                II-xv        DRAFT-DO NOT QUOTE OR CITE

-------
                                  List of Tables
                                       (cont'd)

Number

8-19    Summary of United States PM10 Respiratory Hospital Admission Studies  	8-151

8-20    Summary of United States PM25 Respiratory Hospital Admission Studies	8-152

8-21    Summary of United States PM10_25 Respiratory Hospital Admission Studies  ....  8-152

8-22    Summary of United States PM10, PM25, and PM10_25 Asthma Medical Visit
        Studies	8-153

8-23    Summary of Asthma PM10 PFT Studies	8-157

8-24    Summary of PM25 PFT Asthma Studies	8-158

8-25    Summary of Asthma PM10 Cough Studies	8-162

8-26    Summary of Asthma PM10 Phlegm  Studies  	8-163

8-27    Summary of Asthma PM10 Lower Respiratory Illness Studies	8-163

8-28    Summary of Asthma PM10 Bronchodilator Use Studies	8-164

8-29    Summary of Asthma PM25 Respiratory Symptom Studies	8-166

8-30    Summary of Non-Asthma PM10 PFT Studies	8-167

8-31    Summary of Non-Asthma PM10 Respiratory  Symptom Studies	8-168

8-32    Summary of Non-Asthma PM2 5 Respiratory Outcome Studies	8-169

8-33    Summary of Non-Asthma Coarse Fraction Studies of Respiratory Endpoints ....  8-170

8-34    Characterization of Co-Pollutant Effects on the Stability and Variance
        Inflation or Deflation of PM Effect Size Estimate (in terms of excess RR)	8-193

8-35    Some New Daily Time Series Studies for Mortality or Morbidity with
        Co-Pollutant Models and Gravimetric PM Indices	8-199

8-36    Single-Day Lags Used in Co-Pollutant Models in Lippmann et al., 2000,
        Tables 13-14	8-201



April 2002                              II-xvi       DRAFT-DO NOT QUOTE OR CITE

-------
                                   List of Tables
                                       (cont'd)

Number                                                                          Page

8-37    Number of Participants, N, in Each Block for the Exposure Study in
        Sarnat et al. (2000) 	8-209

8-38    Correlations Among Ambient Pollutants in Baltimore	8-212

8-39    Summary of Past Ecologic and Case-Control Epidemiologic Studies of
        Outdoor Air and Lung Cancer	8-227

8-40    Comparison of PM10 Effect Sizes Estimated by NMMAPS Analyses for 0, 1,
        and 2 Day Lags for the 20 Largest U.S. Cities 	8-239

8-41    Maximum, Median, and Minimum 90th Percentile of Absolute Values of
        Differences Between Fine Particle Concentrations at Pairs of Monitoring
        Sites in 27 Metropolitan Areas in Order of Decreasing Maximum Difference  . . . 8-259

8-42    Summary of Within-City Heterogeneity by Region	8-260

8-43    Summary of ACS Pollution Indices:  Units, Primary Sources, Number of
        Cities and Subjects Available for Analysis, and the Mean Levels	8-269

8A-1    Short-Term Particulate Matter Exposure Mortality Effects Studies	  8A-2

8B-1    Acute Parti culate Matter Exposure and Cardiovascular Hospital Admissions	8B-3

8B-2    Acute Particulate Matter Exposure and Respiratory Hospital Admissions
        Studies	8B-18

8B-3    Acute Particulate Matter Exposure and Respiratory Hospital Admissions
        Studies	8B-40

8B-4    Short-Term Particulate Matter Exposure Effects on Pulmonary Function
        Tests in Studies of Asthmatics 	8B-52

8B-5    Short-Term Particulate Matter Exposure Effects on Symptoms in Studies
        of Asthmatics  	8B-57

8B-6    Short-Term Particulate Matter Exposure Effects on Pulmonary Function
        Tests in Studies of Nonasthmatics 	8B-62
April 2002                               II-xvii       DRAFT-DO NOT QUOTE OR CITE

-------
                                   List of Tables
                                        (cont'd)
Number
8B-7    Short-Term Particulate Matter Exposure Effects on Symptoms in Studies
        ofNonasthmatics  	8B-69

8B-8    Long-Term Particulate Matter Exposure Respiratory Health Indicators:
        Respiratory Symptom, Lung Function 	8B-74

9-1     Constituents of Atmospheric Particles and Their Major Sources	9-11

9-2     Emissions of Primary PM25 by Various Sources in 1999	9-14

9-3     Emissions of Precursors to Secondary PM25 Formation by Various Sources
        in 1999	9-15

9-4     Comparison of Ambient Particles, Fine-Mode (nuclei mode plus accumulation
        mode) and Coarse-Mode	9-21

9-5     Concentrations of PM2 5, PM10_25 and Selected Elements in the PM2 5 and
        PM10.25 Size Range	'	9-22

9-6     Qualitative Estimates of Exposure Variables	9-38

9-7     Particulate Matter Characteristics Potentially Relevant to Health 	9-43

9-8     Volume Mean Diameter of Indoor Particle Sources	9-49

9-9     Concentration Differences Between Constituents of Nonambient
        (Indoor-Generated) and Ambient PM	9-50

9-10    Chemical Species Associated with Mortality in  Epidemiologic Studies  	9-68

9-11    Source Categories Associated with Mortality in Epidemiologic Studies	9-69

9-12    Incidence of Selected Cardiorespiratory Disorders by Age and by Geographic
        Region, 1996	9-72

9-13    Number of Acute Respiratory Conditions per 100 persons per Year, by Age:
        United States, 1996	9-78

9-14    Effect Estimates per Variable Increments in 24-Hour Concentrations of Fine
        Particle Indicators (PM25, SOJ, H+) from U.S. and Canadian Studies 	9-96


April 2002                              II-xviii       DRAFT-DO NOT QUOTE OR CITE

-------
                                    List of Tables
                                        (cont'd)
Number
9-15    Effect Estimates Per Variable Increments in 24-Hour Concentrations of
        Coarse-Fraction Particles (PM10_25) from U.S. and Canadian Studies 	9-102

9-16    Summary of Source-Oriented Evaluations of Particulate Matter Components
        in Recent Studies	9-108

9-17    Effect Estimates per Increments in Long-Term Mean Levels of Fine and
        Inhalable Particle Indicators from U.S. and Canadian Studies	9-112

9-18    Percent Increase in Hospital Admissions per 10-//g/m3 Increase in 24-Hour
        PM10 in 14 U.S. Cities	9-120

9-19    Percent Increase in Mortality per 10 //g/m3 PM10 in Seven U.S. Regions  	9-133

9-20    Percent Excess Risk (t-statistic) per 10 //g/m3 Increase in PM for the
        Relationship of Various Indicators of PM with Various Types of Mortality
        (CV = cardiovascular) in  Several Different Locations	9-143

9-21    Examples of How % Excess Risk per 10 //g/m3 Increase in PM Indicator
        Increases for Specific Chemical Components of PM	9-143

9-22    Percent Excess Risk (t-statistic) per Interquartile Increase in PM Indicator
        for the Relationship of Various Indicators of PM with Cardiovascular
        Mortality for Phoenix  	9-143

9A-1    Effect Estimates per 50-//g/m3 Increase in 24-hour PM10 Concentrations from
        U.S. and Canadian Studies 	  9A-2

9A-2    Effect Estimates per Variable Increments in 24-hour Concentrations of
        Fine Particle Indicators (PM25, SOJ, H+) from U.S. and Canadian Studies	  9A-4

9A-3    Effect Estimates per Increments in Annual Mean Levels of Fine Particle
        Indicators From U.S. and Canadian Studies	  9A-5
April 2002                               H-xix        DRAFT-DO NOT QUOTE OR CITE

-------
                                    List of Figures

Number                                                                             Page

6-1      Diagrammatic representation of respiratory tract regions in humans	6-5

6-2      Total respiratory tract deposition (as percentage deposition of amount inhaled)
         in humans as a function of particle size  	6-9

6-3      Total deposition fraction as a function of particle size in 22 healthy men and
         women under six different breathing patterns	6-11

6-4      Extrathoracic deposition (as percentage deposition of the amount inhaled)
         in humans as a function of particle size  	6-13

6-5      Tracheobronchial deposition (as percentage deposition of the amount inhaled)
         in humans as a function of particle size  	6-17

6-6      Alveolar deposition (as percentage deposition of the amount inhaled) in humans
         as a function of particle size	6-17

6-7      Lung deposition fractions in the tracheobronchial (TB) and alveolar (A) regions
         obtained by the bolus technique 	6-18

6-8      Lung deposition fractions in ten volumetric regions for particle sizes ranging
         from ultrafine particle diameter (dp) of 0.04 to 0.01 //m  (Panel A) to fine
         (dp =1.0 //m) (Panel B) and coarse (dp = 3 and 5 //m) (Panels C and D)	6-21

6-9      Regional deposition fraction in laboratory animals as a function of particle size .  . 6-34

6-10     Particle  deposition efficiency in rats and humans as a function of particle size
         for the (A) total respiratory tract, (B) thoracic region, (C) tracheobronchial
         region, and (D) alveolar region	6-36

6-11     Major clearance pathways for particles deposited in the extrathoracic region
         and tracheobronchial tree 	6-40

6-12     Diagram of known and suspected clearance pathways for poorly soluble
         particles depositing in the alveolar region  	6-41

8-la     Strong within-city association between PM and mortality, but no second-stage
         association	8-7

8-lb     Within-city association between PM and mortality ranges from negative to
         positive with mean across cities approximately zero, but with strong positive
         second-stage association	8-7

April 2002                                H-xx        DRAFT-DO NOT QUOTE OR CITE

-------
                                    List of Figures
                                         (cont'd)
Number
8-2      (a) Graphical depiction of confounding; (b) Graphical depiction of effect
         modification; (c) Graphical depiction of a causal agent with a secondary
         confounder;  (d) Graphical depiction of a causal agent and two potential
         confounders	
8-3      Estimated excess risks for PM mortality (1 day lag) for the 90 largest U.S.
         cities as shown in the original NMAPS report	8-27

8-4      Map of the United States showing the 90 cities (the 20 cities are circled) and
         the seven regions considered in the NMMAPS geographic analyses	8-28

8-5      Percent excess mortality risk (lagged 0, 1, or 2 days) estimated in the NMMAPS
         90-City Study to be associated with 10-//g/m3 increases in PM10 concentrations
         in cities aggregated within U.S. regions shown in Figure 8-2  	8-29

8-6      Percent excess risks estimated per 25 //g/m3 increase in PM2 5  or PM10_2 5 from
         new studies evaluating both PM2 5 and PM10_2 5 data for multiple years, based
         on single pollutant (PM only) models	8-43

8-7      Excess risks estimated for sulfate per 5 //g/m3 increase from the studies in
         which both PM2 5 and PM10_2 5 data were available	8-56

8-8      Natural logarithm of relative risk for total and cause-specific mortality per
         10 //g/m3 PM25 (approximately the excess relative risk as a fraction), with
         smoothed concentration-response functions	8-80

8-9      Relative risk of total and cause-specific mortality at 10 //g/m3  PM25 (mean
         of 1979-1983) of alternative statistical models	8-81

8-10     Relative risk of total and cause-specific mortality for particle metrics and
         gaseous pollutants over different averaging periods	8-82

8-11     Univariate relation between percentage of homes with central  AC and
         regression coefficients for (A) CVD, for cities nonwinter peaking PM10
         concentrations (solid line) and winter peaking PM10 concentrations (dashed
         line) and (B) univariate relation between percentage of PM10 from highway
         vehicles and regression coefficients for CVD	8-115

8-12     Acute cardiovascular hospitalizations and particulate matter exposure excess
         risk estimates derived from  selected U.S. PM10 studies	8-124
April 2002                                H-xxi        DRAFT-DO NOT QUOTE OR CITE

-------
                                    List of Figures
                                         (cont'd)
Number
8-13     Maximum excess risk of respiratory-related hospital admissions and visits per
         50-//g/m3 PM10 increment in selected studies of U.S. cities	8-154

8-14     Selected acute pulmonary function change studies of asthmatic children  	8-160

8-15     Odds ratios with 95% confidence interval for cough per 50-//g/m3 increase in
         PM10 for selected asthmatic children studies at lag 0	8-165

8-16     Graphical depiction of actual confounding of the effects of ambient A and
         ambient B  	8-182

8-17     Graphical depiction of under-fitting of A and B	8-183

8-18     Only A is causal, B is not related to the outcome, but both regressors are
         included in the model, a likely cause of variance inflation	8-184

8-19     Graphical depiction of over-fitting of A and B	8-184

8-20     Graphical depiction of mis-fitting of the effects of A and B 	8-184

8-21     Effects of PM10 on total mortality in 20 large U.S. cities, as a function of
         co-pollutant models	8-188

8-22     Effects of particles and gases on total mortality in eight Canadian cities	8-189

8-23     Effects of PM10 or PM25 on circulatory mortality in three U.S. cities  as a
         function of lag days	8-190

8-24     Total mortality from particles and gases in Santa Clara County, CA	8-191

8-25     Cause-specific fine or coarse particle mortality in Detroit, MI	8-191

8-26     Effects of fine particles on total mortality in Mexico City	8-192

8-27     Concentration of PM10 and NO2 versus distance  	8-216

8-28     Marginal posterior distribution for effects of PM10 on all cause mortality at
         lag 0,  1, and 2 for the 90 cities  	8-238
April 2002                                H-xxii       DRAFT-DO NOT QUOTE OR CITE

-------
                                    List of Figures
                                        (cont'd)
Number
8-29     Particulate matter <10 //m in aerodynamic diameter (PM10)-total mortality
         dose-response curve for the mean lag PM10 and 95% credible regions (solid
         lines), 20 largest U.S. cities, 1987-1994	8-247

8-30     The EPA-derived plot showing relationship of PM10 total mortality effects
         estimates and 95% confidence intervals for all cities in the Samet et al.
         (2000a,b) NMMAPS 90-cities analyses in relation to study size (i.e., the
         natural logarithm or numbers of deaths times days of PM observations)	8-273

8-31     The EPA-derived plots showing relationships of PM10-mortality (total,
         nonaccidental) effects estimates and 95% confidence intervals to study size
         (defined as Figure 8-10) for cities broken out by regions as per the NMMAPS
         regional analyses of Samet et al. (2000a,b)	8-274

9-1      A general framework for integrating particulate-matter research	9-2

9-2      Particle size distributions: (a)  number of particles as a function of particle
         diameter: number concentrations are shown on a logarithmic scale to display
         the wide range by site and size and (b) particle volume as a function of particle
         diameter: for the averaged urban and freeway-influenced urban number
         distributions shown in Figure 2-1 of Chapter 2 	9-5

9-3      Volume size distribution, measured in traffic, showing fine-mode and
         coarse-mode particles and the nuclei and accumulation modes within the
         fine-particle mode	9-6

9-4      Specified particle penetration (size-cut curves) through an ideal
         (no-particle-loss) inlet for five different size-selective sampling criteria	9-8

9-5      An idealized distribution of ambient particulate matter showing fine-mode
         particles and coarse-mode particles and the fractions collected by size-selective
         samplers 	9-9

9-6      Philadelphia, PA-NJ MSA 	9-19

9-7      Occurrence of differences between pairs of sites in three MS As	9-20

9-8      Major chemical components of PM2 5 as determined in the pilot study for
         EPA's national speciation network	9-23
April 2002                               H-xxiii       DRAFT-DO NOT QUOTE OR CITE

-------
                                    List of Figures
                                        (cont'd)
Number
9-9     Main components of a comprehensive atmospheric chemistry modeling
        system, such as Models 3  	9-24

9-10    Correlograms showing the variation in site-to-site correlation coefficient for
        PM25 as a function of distance between sites for several cities  	9-29

9-11    Regression analysis of daytime total personal exposures to PM10 versus
        ambient PM10 concentrations using data from the PTEAM study  	9-32

9-12    Comparison of correlation coefficients for longitudinal analyses of personal
        exposure for individual subjects versus ambient concentrations of PM25 and
        sulfate  	9-33

9-13    Regression analysis of daytime exposures to the ambient component of personal
        exposure to PM10 (ambient exposure) versus ambient PM10 concentrations  	9-34

9-14    Regression analysis of daytime exposures to the nonambient component of
        personal exposure to PM10 (nonambient exposure) versus ambient PM10
        concentrations	9-34

9-15    Distribution of individual, daily values of the infiltration factor, F^ =
        C(AI)/C and the attenuation factor, a  = A/C, estimated using data from the
        PTEAM study	9-36

9-16    Percentage of homes with air conditioning versus the regression coefficient
        for the relationship of cardiovascular-related hospital emissions to ambient
        PM10 concentrations 	9-37

9-17    Spatial variation of PM25, PM10, and PM10_25 as shown by site-to-site correlation
        coefficients as a function of distance between sites for summer 1992 and 1993 in
        Philadelphia, PA	9-45

9-18    Comparison of site-to-site correlation coefficients for PM25 and PM10_2 5 for
        several cities 	9-45

9-19    Site-to-site correlation coefficients for PM2 5 mass and some chemical
        components of PM25 in 1994 in Philadelphia, PA	9-46

9-20    Site-to-site correlation coefficients for PM2 5 mass and several source category
        factors in 1986 in the South Coast Basin (Los Angeles area)	9-46
April 2002                               II-xxiv       DRAFT-DO NOT QUOTE OR CITE

-------
                                    List of Figures
                                        (cont'd)
Number
9-21     Values of geometric mean infiltration factor, F^ = A/C, as a function of
         particle diameter for hourly nighttime data (assuming no indoor sources)
         for summer and fall seasons 	9-47

9-22     Values of penetration efficiency and deposition rate as a function of particle
         diameter estimated from model of average nighttime indoor-outdoor
         concentration data	9-48

9-23     Inhalation rates on a per body-weight basis for males (•) and females (±)
         by age (Layton, 1993)	9-79

9-24     Schematic representation of potential pathophysiological pathways and
         mechanisms by which ambient PM may increase risk of cardiovascular
         morbidity and/or mortality  	9-81

9-25     Percent excess risks estimated per 25-//g/m3 increase in PM2 5 or PM10_2 5
         from new studies evaluating both PM2 5 and PM10_25 data for multiple years	9-105

9-26     Relative risks estimated per 5-//g/m3 increase in sulfate from U.S. and Canadian
         studies in which both PM2 5 and PM10_25 data were available	9-107

9-27     Acute cardiovascular hospitalizations and PM exposure excess risk estimates
         derived from selected U.S. PM10 studies	9-118

9-28     Maximum excess risk in selected studies of U.S. cities relating PM10 estimate
         of exposure (50 //g/m3) to respiratory-related hospital admissions and visits  .... 9-124

9-29     Selected acute pulmonary function change studies of asthmatic children 	9-127

9-30     Odds ratios  for cough for a 50-//g/m3 increase in PM10 for selected asthmatic
         children studies, with lag 0 with 95% CI  	9-127

9-31     Marginal posterior distributions for effect of PM10 on total mortality at lag 1,
         with and without control for other pollutants, for the 90 cities   	9-131

9-32     An expanded version of the Risk Assessment Framework:  (a) PM sources
         to PM exposure, (b) PM exposure to PM dose	9-140

9-33     Schematic showing major nonvolatile and semivolatile components of PM2 5  ... 9-141
April 2002                               II-xxv       DRAFT-DO NOT QUOTE OR CITE

-------
                                    List of Figures
                                        (cont'd)
Number
9-34     The percent deposition of inhaled particles in the tracheobronchial and
         alveolar regions of the lung as a function of particle size	9-147

9-35     Diagram showing relationships (correlations) between A and B and between
         various concentration, exposure, and outcome measures	9-148

9-36     Diagram showing concentrations—exposure—outcome relationships
         (correlations for CO or NO2, PM2 5, and source category factors for
         vehicular traffic related PM and regional sulfate)	9-152
April 2002                               II-xxvi       DRAFT-DO NOT QUOTE OR CITE

-------
                     Authors, Contributors, and Reviewers
               CHAPTER 6. DOSIMETRY OF PARTICULATE MATTER
Principal Authors

Dr. Lawrence J. Folinsbee—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Ramesh Sarangapani—IFC Consulting, Research Triangle Park, NC 27711

Dr. Richard Schlesinger—New York University School of Medicine, Department of
Environmental Medicine, 57 Old Forge Road, Tuxedo, NY 10987

Dr. James McGrath—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Mr. James Raub—National Center for Environmental Assessment (MD-52), U.S. Environmental
Protection Agency, Research Triangle Park, NC 27711

Contributors and Reviewers

Dr. Dan Costa—National Health and Environmental Effects Research Laboratory (MD-82),
U.S. Environmental Protection Agency, Research Triangle  Park, NC 27711

Dr. Robert Devlin—National Health and Environmental Effects Research Laboratory
(MD58),U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Kevin Dreher—National Health and Environmental Effects Research Laboratory (MD-82),
U.S. Environmental Protection Agency, Research Triangle  Park, NC 27711

Dr. Andrew Ohio—National Health and Environmental Effects Research Laboratory (MD-58D),
U.S. Environmental Protection Agency, Research Triangle  Park, NC 27711

Dr. Judith Graham—National Exposure Research Laboratory (MD-75),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Chong Kim—National  Health and Environmental Effects Research Laboratory (MD-58B),
U.S. Environmental Protection Agency, Research Triangle  Park, NC 27711

Dr. Hillel Koren—National Health and Environmental Effects Research Laboratory (MD-58A),
U.S. Environmental Protection Agency, Research Triangle  Park, NC 27711
April 2002                              II-xxvii      DRAFT-DO NOT QUOTE OR CITE

-------
                     Authors, Contributors, and Reviewers
                                       (cont'd)
Contributors and Reviewers
(cont'd)

Dr. Ted Martonen—National Health and Environmental Effects Research Laboratory (MD-74),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Jim Samet—National Health and Environmental Effects Research Laboratory (MD-58),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Ravi Subramaniam—National Center for Environmental Assessment (8623D),
U.S. Environmental Protection Agency, Washington, DC 20460

Dr. William Watkinson—National Health and Environmental Effects Research Laboratory
(MD-82), U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. William Bennett—University of North Carolina at Chapel Hill, Campus Box 7310,
Chapel Hill, NC 37599

Dr. Mark Frampton—University of Rochester, 601  Elmwood Avenue, Box 692, Rochester, NY
14642

Dr. John Godleski—421 ConantRoad, Weston, MA 02493

Dr. Gunter Oberdorster—University of Rochester, Department of Environmental Medicine,
Rochester, NY 14642

Dr. Kent Pinkerton—University of California, ITEH, One Shields Avenue, Davis, CA 95616

Dr. Peter J.A. Rombout—National Institute of Public Health and Environmental Hygiene,
Department of Inhalation Toxicology, P.O. Box 1, NL-3720 BA Bilthoven, The Netherlands

Dr. Vanessa Vu—Office of Research and Development, U.S. Environmental Protection Agency
(8601), Waterside Mall, 401 M St. S.W., Washington, DC 20460
April 2002                             II-xxviii       DRAFT-DO NOT QUOTE OR CITE

-------
                     Authors, Contributors, and Reviewers
                                      (cont'd)
     CHAPTER 7.  TOXICOLOGY OF PARTICULATE MATTER IN HUMANS AND
                             LABORA TOR Y ANIMALS
Principal Authors

Dr. Lung Chi Chen—New York University School of Medicine, Nelson Institute of
Environmental Medicine, 57 Old Forge Road, Tuxedo, NY  10987

Dr. Lawrence J. Folinsbee—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Terry Gordon—New York University Medical Center, Department of Environmental
Medicine, 57 Old Forge Road, Tuxedo, NY  10987

Dr. James McGrath—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Christine Nadziejko—Department of Environmental Medicine, New York University School
of Medicine, Tuxedo, NY

Mr. James Raub—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Contributors and Reviewers

Dr. Susanne Becker—National Health and Environmental Effects Research Laboratory (MD-58),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Dan Costa—National Health and Environmental Effects Research Laboratory (MD-82),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Robert Devlin—National Health and Environmental Effects Research Laboratory (MD-58),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Kevin Dreher—National Health and Environmental Effects Research Laboratory (MD-82),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Janice Dye—National Health and Environmental Effects Research Laboratory (MD-82),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711
April 2002                              II-xxix       DRAFT-DO NOT QUOTE OR CITE

-------
                      Authors, Contributors, and Reviewers
                                       (cont'd)
Contributors and Reviewers
(cont'd)

Dr. Andrew Ohio—National Health and Environmental Effects Research Laboratory (MD-58D),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Ian Gilmour—National Health and Environmental Effects Research Laboratory (MD-82),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Tony Huang—National Health and Environmental Effects Research Laboratory (MD-58),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Judith Graham—National Exposure Research Laboratory (MD-75),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Chong Kim—National Health and Environmental Effects Research Laboratory (MD-58B),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Urmila Kodavanti—National Health and Environmental Effects Research Laboratory
(MD-82), U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Hillel Koren—National Health and Environmental Effects Research Laboratory (MD-58A),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Michael Madden—National Health and Environmental Effects Research Laboratory
(MD-58B), U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Ted Martonen—National Health and Environmental Effects Research Laboratory (MD-74),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Jim Samet—National Health and Environmental Effects Research Laboratory (MD-58),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. William Watkinson—National Health and Environmental Effects Research Laboratory
(MD-82), U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. William Bennett—University of North Carolina at Chapel Hill, Campus Box 7310,
Chapel Hill, NC 37599

Dr. MarkFrampton—University of Rochester, 601 Elmwood Avenue, Box 692, Rochester, NY
14642
April 2002                               II-xxx       DRAFT-DO NOT QUOTE OR CITE

-------
                     Authors, Contributors, and Reviewers
                                      (cont'd)
Contributors and Reviewers
(cont'd)

Dr. John Godleski—421 ConantRoad, Weston, MA 02493

Dr. Gunter Oberdorster—University of Rochester, Department of Environmental Medicine,
Rochester, NY  14642

Dr. Kent Pinkerton—University of California, ITEH, One Shields Avenue, Davis, CA 95616

Dr. Peter J.A. Rombout—National Institute of Public Health and Environmental Hygiene,
Department of Inhalation Toxicology, P.O. Box 1, NL-3720 BA Bilthoven, The Netherlands
        CHAPTER 8. EPIDEMIOLOGY OF HUMAN HEALTH EFFECTS FROM
                        AMBIENT PARTICULA TE MA TTER
Principal Authors

Dr. Lester D. Grant—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Vic Hasselblad—29 Autumn Woods Drive, Durham, NC  27713

Dr. Kazuhiko Ito—New York University Medical Center, Institute of Environmental Medicine,
Long Meadow Road, Tuxedo, NY 10987

Dr. Patrick Kinney, Columbia University, 60 Haven Avenue, B-l, Room 119,
New York, NY 10032

Dr. Dennis J. Kotchmar—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Allan Marcus—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. George Thurston—New York University Medical Center, Institute of Environmental
Medicine, Long Meadow Road,  Tuxedo, NY 10987
April 2002                             II-xxxi       DRAFT-DO NOT QUOTE OR CITE

-------
                      Authors, Contributors, and Reviewers
                                       (cont'd)
Contributors and Reviewers

Dr. Burt Brunekreef—Agricultural University, Environmental and Occupational Health,
P.O. Box 238, NL 6700 AE, Wageningen, The Netherlands

Dr. Richard Burnett—Health Canada, 200 Environmental Health Centre, Tunney's Pasture,
Ottawa, Canada KlA OL2

 Dr. Raymond Carroll—Texas A & M University, Department of Statistics, College Station, TX
77843-3143

Dr. Robert Chapman—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Steven Colome—Integrated Environmental Services, 5319 University Drive, #430,
Irvine, CA 92612

Dr. Ralph Delfmo—University of California at Irvine, Epidemiology Division, Department of
Medicine, University of California at Irvine, Irvine, CA 92717

Dr. Douglas Dockery—Harvard School of Public Health, 665 Huntington Avenue, 1-1414,
Boston, MA 02115

Dr. Peter Guttorp—University of Washington, Department of Statistics, Box 354322
Seattle, WA 98195

Dr. Scott R. Kegler—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Fred Lipfert—23 Carll Court, Northport, NY 11768

Dr. Lee-Jane Sally Liu—University of Washington, Department of Environmental Health,
Box 357234, Seattle, WA  98195

Dr. Suresh Moolgavakar—Fred Hutchinson Cancer Research Center, 1100 Fairview Avenue,
N-MP 665, Seattle, WA 98109

Dr. Robert D. Morris—Tufts University, 136 Harrison Avenue, Boston, MA 02111

Dr. Lucas Neas—National Health and Environmental Effects Research Laboratory (MD-58),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711
April 2002                              II-xxxii      DRAFT-DO NOT QUOTE OR CITE

-------
                     Authors, Contributors, and Reviewers
                                      (cont'd)
Contributors and Reviewers
(cont'd)

Dr. James Robins—Harvard School of Public Health, Department of Epidemiology,
Boston, MA  02115

Dr. Isabelle Romieu—Centers for Disease Control (CDC), 4770 Bufford Hwy, NE,
Atlanta, GA  30341

Dr. Mary Ross—Office of Air Quality Planning and Standards (MD-15),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Lianne Sheppard—University of Washington, Box 357232, Seattle, WA 98195-7232

Dr. Richard L. Smith—University of North Carolina, Department of Statistics, Box 3260
Chapel Hill, NC 27599

Dr. Leonard Stefanski—North Carolina State University, Department of Statistics, Box 8203,
Raleigh, NC  27695

Dr. Duncan Thomas—University of Southern California, Preventative Medicine Department,
1540 Alcazar Street, CH-220, Los Angeles, CA 90033-9987

Dr. Clarice Weinberg—National Institute of Environmental Health Sciences, P.O. Box 12233,
Research Triangle Park, NC 27709

Dr. William Wilson—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711
        CHAPTER 9.  INTEGRA TIVE SYNTHESIS: PARTICULA TE MATTER
          ATMOSPHERIC SCIENCE, AIR QUALITY, HUMAN EXPOSURE,
                        DOSIMETRY, AND HEALTH RISKS
Principal Authors

Dr. William E. Wilson—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Lester D. Grant—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

April 2002                             II-xxxiii      DRAFT-DO NOT QUOTE OR CITE

-------
                     Authors, Contributors, and Reviewers
                                      (cont'd)
Principal Authors
(cont'd)

Dr. Dennis J. Kotchmar—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Allan Marcus—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Joseph P. Pinto—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Mr. James Raub—National Center for Environmental Assessment (MD-52), U.S. Environmental
Protection Agency, Research Triangle Park, NC 27711

Contributors and Reviewers

Dr. Robert Chapman—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Mr. William Ewald—National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711
April 2002                             H-xxxiv      DRAFT-DO NOT QUOTE OR CITE

-------
             U.S. ENVIRONMENTAL PROTECTION AGENCY
  PROJECT TEAM FOR DEVELOPMENT OF AIR QUALITY CRITERIA
                        FOR PARTICULATE MATTER
Executive Director

Dr. Lester D. Grant—Director, National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Scientific Staff

Dr. William E. Wilson—Air Quality Coordinator, Physical Scientist, National Center for
Environmental Assessment (MD-52), U.S. Environmental Protection Agency, Research Triangle
Park,NC 27711

Dr. Lawrence J. Folinsbee—Health Coordinator, Chief, Environmental Media Assessment
Group, National Center for Environmental Assessment (MD-52), U.S. Environmental Protection
Agency, Research Triangle Park, NC 27711 (now deceased)

Dr. Dennis J. Kotchmar—Project Manager, Medical Officer, National Center for Environmental
Assessment (MD-52), U.S. Environmental Protection Agency, Research Triangle Park, NC
27711

Dr. Robert Chapman—Medical Officer, National Center for Environmental Assessment
(MD-52), U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Ms. Beverly Comfort—Health Scientist, National Center for Environmental Assessment
(MD-52), U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Mr. William Ewald—Health Scientist, National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. J.H.B. Garner—Ecological Scientist, National Center for Environmental Assessment
(MD-52), U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. David Mage—Physical Scientist, National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Allan Marcus—Statistician, National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. James McGrath—Visiting Senior Health Scientist, National Center for Environmental
Assessment (MD-52), U.S. Environmental Protection Agency, Research Triangle Park, NC
27711
April 2002                             H-xxxv      DRAFT-DO NOT QUOTE OR CITE

-------
             U.S. ENVIRONMENTAL PROTECTION AGENCY
  PROJECT TEAM FOR DEVELOPMENT OF AIR QUALITY CRITERIA
                       FOR PARTICULATE MATTER
                                      (cont'd)
Scientific Staff
(cont'd)

Dr. Joseph P. Pinto—Physical Scientist, National Center for Environmental Assessment,
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Mr. James A. Raub—Health Scientist, National Center for Environmental Assessment (MD-52),
U. S. Environmental Protection Agency, Research Triangle Park, NC 27711

Technical Support Staff

Mr. Randy Brady—Deputy Director, National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Mr. Douglas B. Fennell—Technical Information Specialist, National Center for Environmental
Assessment (MD-52), U.S. Environmental Protection Agency, Research Triangle Park, NC
27711

Ms. Emily R. Lee—Management Analyst, National Center for Environmental Assessment
(MD-52), U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Ms. Diane H. Ray—Program Specialist, National Center for Environmental Assessment
(MD-52), U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Ms. Eleanor Speh—Office Manager, Environmental Media Assessment Branch, National Center
for Environmental Assessment (MD-52), U.S. Environmental Protection Agency, Research
Triangle Park, NC 27711

Ms. Donna Wicker—Administrative Officer, National Center for Environmental Assessment
(MD-52), U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Mr. Richard Wilson—Clerk, National Center for Environmental Assessment (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711
April 2002                             U-xxxvi      DRAFT-DO NOT QUOTE OR CITE

-------
             U.S. ENVIRONMENTAL PROTECTION AGENCY
 PROJECT TEAM FOR DEVELOPMENT OF AIR QUALITY CRITERIA
                       FOR PARTICULATE MATTER
                                      (cont'd)


Document Production Staff

Dr. Carol A. Seagle—Technical Editor, Computer Sciences Corporation,
2803 Slater Road, Suite 220, Morrisville, NC 27560

Ms. Diane G. Caudill—Graphic Artist, Computer Sciences Corporation,
2803 Slater Road, Suite 220, Morrisville, NC 27560

Ms. Carolyn T. Perry—Word Processor, Computer Sciences Corporation,
2803 Slater Road, Suite 220, Morrisville, NC 27560

Ms. Kelly Quifiones—Word Processor, InfoPro, Inc., 8405 Colesville Road, 2nd Floor,
Silver Spring, MD 20910


Technical Reference Staff

Mr. John A. Bennett—Technical Information Specialist, SANAD Support Technologies, Inc.,
11820 Parklawn Drive, Suite 400, Rockville, MD 20852

Ms. Sandra L. Hughey—Technical Information Specialist, SANAD Support Technologies, Inc.,
11820 Parklawn Drive, Suite 400, Rockville, MD 20852

Ms. Beth Olen—Records Management Technician, Reference Retrieval and Database Entry
Clerk, InfoPro, Inc., 8405 Colesville Road, 2nd Floor, Silver Spring, MD 20910
April 2002                            U-xxxvii      DRAFT-DO NOT QUOTE OR CITE

-------
             Abbreviations and Acronyms
oabs light-absorption coefficient
oag light-absorption coefficient of gases
oap light-absorption coefficient of particles
°~ext
light-extinction coefficient
og geometric standard deviation
"scat
light-scattering coefficient
osg light-scattering coefficient of gases
osp light-scattering coefficient of particles
4-POBN
A
AAS
ACGffl
a-(4-pyridyl- 1 -oxide)-N-tert-butylnitrone
alveolar
atomic absorption spectrophotometry
American Conference of Governmental Industrial Hygienists
AD
ADS
AES
AIRS
AM
AQCD
AQI
AQRV
ARIES
ASOS
ATOM
ATOFMS
annular denuder system
atomic emission spectroscopy
Aerometric Information Retrieval System
alveolar macrophages
Air Quality Criteria Document
Air Quality Index
Air Quality Related Values
Aerosol Research and Inhalation Epidemiology Study
Automated Surface Observing System
aerosol and toxic deposition model
time-of-flight mass spectrometer
b
Ba
absorption coefficient
April 2002
D-xxxviii
DRAFT-DO NOT QUOTE OR CITE

-------
 BAD
brachial artery diameter
 BAL
bronchoalveolar lavage
 BALF
bronchoalveolar lavage fluid
 BAUS
brachial artery ultrasonography
 BC
black carbon (see also CB)
 BW
bronchial wash
 BYU
Bringham Young University
 C
apparent contrast
 Ca+
calcium
 CAA
Clean Air Act
 CAAM
continuous ambient mass monitor
 CAMNET
 CAPs
concentrated ambient particles
 CARS
California Air Resources Board
 CASAC
Clean Air Scientific Advisory Committee
 CASTNet
Clean Air Status and Trends Network
 CAT
computer-aided tomography
 CB
carbon black
                         base cation
 CC
carbonate carbon
 CC14
carbon tetrachloride
 CCPM
continuous coarse particle monitor
 CCSEM
computer-controlled scanning electron microscopy
 CEN
European Standardization Committee
 CF
Cystic Fibrosis
 CFA
coal fly ash
 CFCs
chlorofluorocarbons
 CFD
computational fluid dynamics
April 2002
             II-xxxix
DRAFT-DO NOT QUOTE OR CITE

-------
CFR
CH2O
GIF
CL
CMAQ
CMB
CMD
CMP
CMSA
C0
CO
CO CD
COPD
CPC
CPZ
CR
CRP
Code of Federal Regulations
formaldehyde
charcoal-impregnated cellulose fiber
chemiluminescence
Community Multi-Scale Air Quality
chemical mass balance
count mean diameter
copper smelter dust
Consolidated Metropolitan Statistical Area
initial contrast
carbon monoxide
Air Quality Criteria Document for Carbon Monoxide
chronic obstructive pulmonary disease
condensation particle counter
capsazepine
concentration-response
Coordinated Research Program
CSIRO
CSMCS
CTM
CV
Carbonaceous Species Methods Comparison Study
chemistry-transport model
coefficient of variation
D5o
Da
DAQM
DCFH
DE
DE
DEF
Denver Air Quality Model
dichlorofluorescin
deposition efficiencies
diesel exhaust
Deferoxamine
April 2002
n-xi
DRAFT-DO NOT QUOTE OR CITE

-------
 DEP
diesel exhaust particles
 DHR
dihydrorhodamine-123
 DMS
dimethyl sulfide
 DMTU
dimethylthiourea
 DOFA
domestic oil fly ash
 DPM
diesel particulate matter
 DRG
dorsal root ganglia
 dv
deciview index
 BAD
electrical aerosol detector
 EC
elemental carbon
 ECAO
Environmental Criteria and Assessment Office
 ECG
electrocardiogram
 EDXRF
energy dispersive X-ray fluorescence
 EGA
evolved gas analysis
 EGF
epidermal growth factor
 ELSIE
Elastic Light Scattering and Interactive Efficiency
 ERK
extracellular receptor kinase
 ESP
electrostatic precipitator
 ESR
electron spin resonance
 ET
extrathoacic
 ETS
environmental tobacco smoke
 EU
endotoxin units
 EXPOLIS
                          flux
 FEF
forced expiratory flow
 FEVj
forced expiratory volume in 1 second
 FID
flame ionization detection
 FMD
flow-mediated dilation
April 2002
               n-xii
DRAFT-DO NOT QUOTE OR CITE

-------
 FPD
flame photometric detector
 FRM
Federal Reference Method
 gS02
gaseous sulfur dioxide
 GC
gas chromatography
 GCMs
General Circulation Models
 GCVTC
Grand Canyon Visibility Transport Commission
 GG/MSD
gas chromatography/mass-selective detection
 GHG
greenhouse gases
 GMCSF
granulocyte macrophage colony stimulating factor
 GMPD
geometric mean particle diameter
 GSD
geometric standard deviation (see also o )
 GSH
glutathione
 H2SO4
sulfuric acid
 HAAQS
 HDM
house dust mite
 HDS
honeycomb denuder/filter pack sampler
 HEADS
Harvard-EPA Annular Denuder Sampler
 HEI
Health Effects Institute
 hivol
High blume sampler
 HNO3
nitric acid
 HR
heart rate
 HTGC-MS
high temperature gas chromotography-mass spectrometry
                         radiance
                         inhibitory kappa B alpha
                         apparent radiance of the background
                         transmitted radiance of the background
 1C
ion chromatography
 ICAM-1
intercellular adhesion molecule-1
April 2002
              H-xlii
DRAFT-DO NOT QUOTE OR CITE

-------
 ICP
inductively coupled plasma
 ICRP
International Commission on Radiological Protection
 le
equilibrium radiance or source function
 IPS
Integrated Forest Study
 IgE
immunoglobin E
 IgG
immunoglobin G
 IL
interleukin
 IMPROVE
Interagency Monitoring of Protected Visual Environments
 INAA
instrumental neutron activation analysis
 IOVPS
integrated organic vapor/particle sampler
                          intraperitoneal
                          path radiance
 IPCC
Intergovernmental Panel on Climate Change
 IPM
inhalable paniculate matter
 IPN
Inhalable Paniculate Network
 ISO
International Standards Organization
                          transmitted radiance
 INK
c-jun N-terminal kinase
 'scp
                          light scattering by coarse particles
 Jsfp
light scattering by fine particles
 Jspd
light scattering coefficient of particles under dry conditions
 'spw
                          light scattering coefficient of particles under humid conditions
 K
Koschmieder constant
 K+
potassium ion
 KOH
potassium hydroxide
 LAI
leaf area indices
 LFA-1
leukocyte function-associated antigen-1
 LN
lymph nodes
April 2002
              H-xliii
DRAFT-DO NOT QUOTE OR CITE

-------
LoS
1pm, Lpm, L/min
LPS
LWCA
MAA
MAACS
MADPro
MAPK
MAQSIP
MCM
MCT
MEK
MIP
Mm
MMAD
MMD
MMPs
MOUDI
MPL
MPO
MS
MSA
MSAs
MSH
low sulfur
liters per minute
lipopolysaccharide
liquid water content analyzer
mineral acid anion
Metropolitan Acid Aerosol Characterization Study
Mountain Acid Deposition Program
mitogen-activated protein kinase
page 3-83
mass concentrations monitor
monocrotaline
mitogen-activated protein kinase
macrophage inflammatory protein
megameters
mean median aerodynamic diameter (see og)
mass median diameter
matrix metalloproteinases
micro-orifice uniform deposit impactor
multipath lung
myeloperoxidase
mass spectroscopy
methane sulfonic acid
metropolitan statistical areas
Mount St. Helens
MSP
NAC
NAL
NAMS
N-acetylcysteine (antioxidant)
nasal lavage fluid
National Ambient Monitoring Stations
April 2002
DRAFT-DO NOT QUOTE OR CITE

-------
 NaN3
                        sodium azide
 NAPAP
                        National Acid Precipitation Assessment Program
 NAPRMN
 NARSTO
 NAST
                        National Assessment Synthesis Team
 NCRPM
                        National Council on Radiation Protection and Measurements
 ND
                        NIST diesel (also, not determined)
 NDDN
                        National Dry Deposition Network
 NDIR
                        nondispersive infrared spectrophotometry
 NESCAUM
                        Northeast States for Coordinated Air Use Management
 NF
                        nuclear factor
 NF-KB
                        nuclear factor kappa B
 NFRAQS
                        North Frontal Range Air Quality Study
 NH3
                        ammonia
 NIL
                         ammonium
 (NH4)2 S04
                        ammonium sulfate
NILH,S(X
                         ammonium acid sulfate
 NHBE
                        normal human bronchial epithelial
 NIOSH
 NIR
 NIST
                        National Institute of Standards and Technology
 NMD
                        nitroglycerine-mediated dilation
 NMD
                        number mean diameter
 NMRI
                        Naval Medical Research Institute
 NO
                        nitrogen oxide
 NO,
                        nitrogen dioxide
 NO3-
                        nitrate
 NOPL
                        naso-oro-pharyngo-laryngeal
April 2002
                                      II-xlv
DRAFT-DO NOT QUOTE OR CITE

-------
NOX
NPP
NRC
NuCM
03
OAA
OAQPS
OAR
OC
OFA
CH-
ORD
OVA
P
P so42-
PAH
PAHs
PAN
PAR
PB
PEL
nitrogen oxides
net primary production
National Research Council
nutrient cycling model
ozone
Ottowa ambient air
Office of Air Quality Planning and Standards
Office of Air and Radiation
organic carbon
oil fly ask
hydroxyl ion
Office of Research and Development
ovalbumin
partial pressure
particulate sulfate
polynuclear aromatic hydrocarbon
polycyclic aromatic hydrocarbons
peroxyacetyl nitrate
photosynthetically active radiation
polymyxin-B
planetary boundary layer
PBY
PC
PC
PC-BOSS
PCA
PCBs
pyrolitic carbon
particle concentrator
Particulate Concentrator-Brigham Young University Organic
Sampling System
principal component analysis
polychloronated biphenyls
April 2002
DRAFT-DO NOT QUOTE OR CITE

-------
 PCDD
polychlorinateddibenzo-^-dioxins
 PCDF
polychlorinated dibenzofurans
 PCM
particle composition monitor
 pdf
probability density functions
 PDGF
platelet-derived growth factor
 PEM
Personal Environmental Monitor
 PESA
proton elastic scattering analysis
 PFA
 PIXE
proton induced X-ray emission
 PM
particulate matter
 PM AQCD
PM Air Quality Criteria Document
 PM(10.25)
coarse particulate matter
 PM
    -2.5
fine particulate matter
 PMF
positive matrix factorization
 PMN
polymorphonuclear leukocytes
                          equilibrium vapor pressure
 poly I:C
polyionosinic-polycytidilic acid
 POP
persistent organic pollutant
 PROBDET
Probability of Detection Algorithm
 PTEAMS
 PTEP
PM10 Technical Enhancement Program
 PTFE
polytetrafluoroethylene
 PTFE
polytetrafluoroethylene
 PUF
polyurethane foam
 Q
respiratory flow rates
 Qabs
efficiency of absorption
 Qext
efficiency of extinction
 Qscat
efficiency of scattering
April 2002
                           DRAFT-DO NOT QUOTE OR CITE

-------
                          aerodynamic resistance
 RAAS
 RADM
Regional Acid Deposition Model
 RAMS
Real-Time Air Monitoring System
 RAMS
Regional Air Monitoring Study
 RAPS
Regional Air Pollution Study
                          boundary layer resistance
                          canopy resistance
 REMSAD
Regulatory Modeling System for Aerosols and Deposition
 RFC
residual fuels oils
 RH
relative humidity
 ROFA
residual oil fly ash
 ROFA
residual oil fly ash
 ROME
Reactive and Optics Model Emissions
 ROS
reactive oxygen species
 RPM
respirable particulate matter
 RPM
Regional Particulate Model
 RTE
rat tracheal epithelial
 RTF
Research Triangle Park
 SASS
 sec
                          saturation ratio
 SA
Sierra Anderson
 SAD
small airway disease
 SCAQS
Southern California Air Quality Study
 scos
Southern California Ozone Study
 sd
standard deviation
 SEM
scanning electron microscopy
April 2002
                           DRAFT-DO NOT QUOTE OR CITE

-------
 SES
sample equilibration system
 SEV
Sensor Equivalent Visibility
 SH
spontaneously hypertensive
 SIP
State Implementation Plans
 SIXE
synchrotron induced X-ray emission
 SL
stochastic lung
 SLAMS/NAMS
 SLAMS
State and Local Air Monitoring Stations
 SLE
St. Louis encephalitis
 SMPS
scanning mobility particle sizer
 SO,
sulfur dioxide
 scx2-
sulfate
 SOA
 SOC
semivolatile organic compounds
 SoCAB
South Coast Air Basin
 SOD
superoxide dismutase
 SOPM
secondary organic particulate matter
 SP
Staff Paper
 SPM
synthetic polymer monomers
 SRI
 SRM
standard reference method
 SSM
solid sampler module
 Stk
Stokes number
 SUVB
solar ultraviolet B radiation
 svoc
semivolatile organic compounds
 SWMMC
Southwest Metropolitan Mexico City
 T(CO)
core temperature
 TB
tracheabronchial
April 2002
              H-xlix
DRAFT-DO NOT QUOTE OR CITE

-------
 TDF
total deposition fraction
 TDMA
Tandem Differential Mobility Analyzer
 TEOM
tapered element oscillating microbalance
 TEOMs
 TIMP
tissue inhibitor of metaloproteinase
 TLN
 TNF
tumor necrosis factor
 TOFMS
aerosol time-of-flight mass spectroscopy
 TOR
thermal/optical reflectance
 TOT
thermal/optical transmission
 TPM
thoracic paniculate matter
 TRXRF
total reflection X-ray fluorescence
 TSI
 TSP
total suspended particulates
 UAM-V
Urban Airshed Model Version V
 UCM
unresolved complex mixture
 ufCB
ultrafine carbon black
 UFP
ultrafine fluorospheres
 UNEP
United Nations Environment Programme
 URG
University Research Glassware
 USGCRP
U.S. Global Change Research Program
 UVD
Utah Valley dust
 VAPS
Versatile Air Pollution Samplers
 VASM
Visibility Assessment Scoping Model
 VBE
Japanese B encephalitis
 VCAM-1
vascular cell adhesion molecule-1
                          deposition velocity
 VDI
April 2002
II-l
                           DRAFT-DO NOT QUOTE OR CITE

-------
voc
vs
v,
v,
we
WEE
WINS
WIS
WKY
WMO
Wo
WRAC
X-XRF
XAD
XRF
volatile organic compounds
sedimentation velocity
turbulent diffusion velocity
tidal volume
tungsten carbide
western equine encephalitis
Well Impactor Ninety- Six
Wistar
Wi star-Kyoto
World Meteorological Organization
single scattering albedo
Wide Range Aerosol Classifier
synchrotron induced X-ray fluorescence
polystyrene-divinyl benzene
X-ray fluorescence
V*
April 2002
Il-li
DRAFT-DO NOT QUOTE OR CITE

-------
 i              6.  DOSIMETRY OF PARTICULATE MATTER
 2
 3
 4      6.1 INTRODUCTION
 5           A basic principle in health effects evaluation is that the dose delivered to the target site of
 6      concern, rather than the external exposure, is the proximal cause of any biological response.
 7      Characterization of the exposure-dose-response continuum for particulate matter (PM), a
 8      fundamental objective of any dose-response assessment for evaluation of health effects, requires
 9      the elucidation and understanding of the mechanistic determinants of inhaled particle dose.
10      Furthermore, dosimetric information is critical to an effective extrapolation to humans of health
11      effects demonstrated by toxicological studies using experimental animals and for comparison of
12      results from  controlled clinical studies involving different types of human subjects, e.g., those
13      with preexisting respiratory disease and normals. Dosimetry provides a critical link in evaluating
14      the relevance of health effects found in animal models of susceptible humans because it allows
15      for discrimination between actual susceptibility differences from those due to differences in sites
16      of particle action.
17           Dose to target tissue is dependent initially on the deposition of particles within the
18      respiratory tract. Particle deposition refers to the removal of particles from their airborne state
19      because of their aerodynamic, thermodynamic, and/or electrostatic behavior. Once particles have
20      deposited onto the surfaces of the respiratory tract, they are subsequently subjected to either
21      absorptive or nonabsorptive particulate removal processes. This may result in their removal from
22      airway surfaces, as well as their removal, to varying degrees, from the respiratory tract itself.
23      The deposited PM thus cleared from initial deposition sites is said to have undergone
24      translocation. Clearance of deposited particles depends upon the initial site of deposition and
25      upon the physicochemical properties of the particles, both of which impact upon specific
26      translocation pathways. Retained particle burdens are determined by the dynamic relationship
27      between deposition and clearance rates.
28           This chapter is concerned with particle dosimetry, the study of the deposition, translocation,
29      clearance, and retention of particles within the respiratory tract and extrapulmonary tissues.
30      It summarizes basic concepts as presented in the 1996 EPA document, Air Quality Criteria for

        April 2002                                 6-1         DRAFT-DO NOT QUOTE OR CITE

-------
 1      Particulate Matter or "PM AQCD" (U.S. Environmental Protection Agency, 1996), specifically
 2      in Chapter 10; and it updates the state of the science based upon new literature appearing since
 3      publication of the 1996 PM AQCD.  Although our understanding of the basic mechanisms
 4      governing deposition and clearance of inhaled particles has not changed, there has been
 5      significant additional information  on the role of certain biological determinants of the
 6      deposition/clearance processes,  such as gender and age.  Also, the understanding of regional
 7      dosimetry and the particle size range over which this has been evaluated has been expanded.
 8           The dose from inhaled particles deposited and retained in the respiratory tract is governed
 9      by a number of factors. These include exposure concentration and exposure duration, respiratory
10      tract anatomy and ventilatory parameters,  and by physicochemical properties of the particles
11      themselves (e.g., particle  size, hygroscopicity, solubility).  The basic characteristics of particles
12      as they relate to deposition and retention, as well as anatomical and physiological factors
13      influencing particle  deposition and retention, were discussed in depth in the  1996 PM AQCD.
14      Thus, in this current chapter, only  an overview of basic information related to one critical factor
15      in deposition, namely particle size, is provided (Section 6.1.1), so as to allow the reader to
16      understand the different terms used in the remainder of this chapter and in subsequent ones
17      dealing with health effects.  This is followed, in Section 6.1.2, by a basic overview of respiratory
18      tract structure as it relates to deposition evaluation.  The ensuing major sections of this chapter
19      provide updated information on particle deposition, clearance, and retention in the respiratory
20      tract of humans, as well as laboratory animals, which are useful in the evaluation of PM health
21      effects.  Issues related to the phenomenon of particle overload as it may apply to human exposure
22      and the use of instillation as an exposure technique to evaluate PM health effects also are
23      discussed. The final sections of the  chapter deal with mathematical models of particle
24      disposition in the respiratory tract.
25           It must be emphasized that any dissection into discrete topics of factors that control dose
26      from inhaled particles tends to mask the dynamic and interdependent nature of the intact
27      respiratory system.  For example, although deposition is discussed separately from clearance
28      mechanisms, retention (i.e., the  actual amount of particles found in the respiratory tract at any
29      point in time) is, as noted previously, determined by the relative rates of both deposition and
30      clearance. Thus, assessment of overall  dosimetry requires integration of these various
31      components of the overall process. In summarizing the literature on particle dosimetry, when

        April 2002                                 6-2         DRAFT-DO NOT QUOTE OR CITE

-------
 1      applicable, changes from control are described if they were statistically significant at a
 2      probability (p) value less than 0.05 (i.e., p < 0.05).  When trends are described, an attempt will be
 3      made to provide the actual p values given in the published reports.
 4
 5      6.1.1  Size Characterization of Inhaled Particles
 6           Information about particle size distribution is important in the evaluation of effective
 7      inhaled dose. This section summarizes particle attributes requiring characterization and provides
 8      some general definitions important in understanding particle fate within the respiratory tract.
 9           Particles exist in the atmosphere as components of aerosols, which are airborne suspensions
10      of finely dispersed solid or liquid particles.  Because  aerosols can consist of almost any material,
11      their description in simple geometric terms  can be misleading unless important factors relating to
12      constituent particle size, shape, and density are considered.  Although the size of particles within
13      aerosols can be described based on actual physical measurements (such  as those obtained with a
14      microscope), in many cases it is better to use some equivalent diameter in place of the physical
15      diameter. The most commonly used metric is aerodynamic equivalent diameter (AED), whereby
16      particles of differing geometric size, shape,  and density are compared in terms of aerodynamic
17      behavior (i.e., terminal setting velocity) to particles that are unit density (1 gm/cm3) spheres. The
18      aerodynamic behavior of unit density spherical particles constitutes a useful standard by which
19      many types of particles can be compared in terms of certain deposition mechanisms. (See
20      Chapter 2 for a more complete discussion.)
21           It is important to note that most aerosols  present in natural and work environments are
22      polydisperse. This means that the constituent particles within an aerosol have a range of sizes
23      and are more appropriately described in terms of a size  distribution parameter. The lognormal
24      distribution (i.e.,  the situation in which the logarithms of particle diameter are distributed
25      normally) can be used for describing size distributions of most aerosols. In linear form, the
26      logarithmic mean is the median of the distribution, and  the metric of variability around this
27      central tendency is the geometric  standard deviation (og). The og, a dimensionless term, is the
28      ratio of the 84th (or 16th) % particle size to the 50th %  size.  Thus, the only two parameters
29      needed to describe a log normal distribution of particle  sizes for a specific aerosol are the median
30      diameter and the  geometric standard deviation. However, the actual size distribution may be
31      obtained in various ways. For example, when  a distribution is described by counting particles,
        April 2002                                  6-3         DRAFT-DO NOT QUOTE OR CITE

-------
 1      the median is called the count median diameter (CMD). On the other hand, the median of a
 2      distribution based on particle mass in an aerosol is the mass median diameter (MMD). When
 3      using aerodynamic diameters, a term that is encountered frequently is mass median aerodynamic
 4      diameter (MMAD), which refers to the median of the distribution of mass with respect to
 5      aerodynamic equivalent diameter. Most of the present discussion will focus on MMAD because
 6      it is the most commonly used measure of aerosol distribution. However, alternative distributions
 7      should be used for particles with actual physical sizes below about 0.5 //m because, for these,
 8      aerodynamic properties become less important.  One such metric is thermodynamic-equivalent
 9      size, which is the diameter of a spherical particle that has the same diffusion coefficient in air as
10      the particle of interest.
11
12      6.1.2 Structure of the Respiratory Tract
13           A detailed discussion of respiratory tract structure was provided in the 1996 PM AQCD
14      (U.S. Environmental Protection Agency, 1996),  and only a brief synopsis is presented here.
15      For dosimetry purposes, the respiratory tract can be divided into three regions (Figure 6-1):
16      (1) extrathoracic (ET), (2) tracheobronchial (TB), and (3) alveolar (A).  The ET region consists
17      of head airways (i.e., nasal and oral passages) through the larynx and represents the areas through
18      which inhaled air first passes.  In humans, inhalation can occur through the nose or mouth (or
19      both, known as oronasal breathing).  However, most laboratory animals commonly used in
20      respiratory toxicological studies are obligate nose breathers.
21           From the ET region, inspired air enters the TB region at the trachea. From the level of the
22      trachea, the conducting airways then undergo branching for a number of generations.  The
23      terminal bronchiole is the most peripheral of the distal conducting airways and these lead,
24      in humans, to the respiratory bronchioles, alveolar ducts, alveolar sacs, and alveoli (all of which
25      comprise the A region). All of the conducting airways, except the trachea and portions of the
26      mainstem bronchi, are surrounded by parenchymal tissue.  This is composed primarily of the
27      alveolated structures of the A region and associated blood and lymphatic vessels.  It should be
28      noted that the respiratory tract regions are comprised of various  cell types and that there are
29      distinct differences in the cells of airway surfaces in the ET, TB, and A regions. Although a
30      discussion of cellular structure of the respiratory tract is beyond the scope of this section, details
31      may be found in a number of sources (e.g., Crystal et al.,  1997).
        April 2002                                 6-4         DRAFT-DO NOT QUOTE OR CITE

-------
                                                                           S
                                                                           s
                                                                           ol
                                                                           4*
                                                                           o
                                                                          'S.
                                                                           0   Co
                                                                           -   0\

                                                                               2
                                                                           •_

                                                                           w

                                                                          *
-------
 1      6.2 PARTICLE DEPOSITION
 2           This section discusses the deposition of particles in the respiratory tract.  It begins with an
 3      overview of the basic physical mechanisms that govern deposition.  This is followed by an
 4      update on both total respiratory tract and regional deposition patterns in humans. Some critical
 5      biological factors that may modulate deposition are then presented.  The section ends with a
 6      discussion of issues related to interspecies patterns of particle deposition.
 7
 8      6.2.1 Mechanisms of Deposition
 9           Particles may deposit within the respiratory tract by five mechanisms:  (1) inertial
10      impaction, (2) sedimentation, (3) diffusion, (4) electrostatic precipitation, and (5) interception.
11           Sudden changes in airstream direction and velocity cause particles to fail to follow the
12      streamlines of airflow.  As a consequence, the particles contact, or impact, onto airway surfaces.
13      The ET and upper TB airways are characterized by high air velocities and  sharp directional
14      changes and, thus, dominate as sites of inertial impaction.  Impaction is a significant deposition
15      mechanism for particles larger than  1 //m AED.
16           All aerosol particles are continuously influenced by gravity, but particles with an
17      AED >1 //m are affected to the greatest  extent.  A particle will  acquire  a terminal settling
18      velocity when a balance is achieved between the acceleration of gravity acting on the particle and
19      the viscous resistance of the air, and it is this settling out of the airstream that takes it into contact
20      with airway surfaces. Both sedimentation and inertial  impaction can influence the deposition of
21      particles within the same size range.  These deposition processes act together in the ET and TB
22      regions, with inertial impaction dominating in the upper airways and gravitational settling
23      becoming increasingly dominant in the smaller conducting airways.
24           Particles having actual physical diameters <1 //m are subjected increasingly to diffusive
25      deposition because of random bombardment by air molecules, which results in contact with
26      airway surfaces. The root mean square displacement that a particle  experiences in a unit of time
27      along a given cartesian coordinate is a measure of its diffusivity. The density of a particle is
28      unimportant in determining particle diffusivity.  Thus,  instead of having an aerodynamic
29      equivalent size, diffusive particles of different shapes can be related to  the diffusivity of a
30      thermodynamic equivalent size based on spherical particles.

        April 2002                                6-6         DRAFT-DO NOT QUOTE OR CITE

-------
 1           The particle size region around 0.2 to 1.0 //m frequently is described as consisting of
 2      particles that are small enough to be minimally influenced by impaction or sedimentation and
 3      large enough to be minimally influenced by diffusion. Such particles are the most persistent in
 4      inhaled air and undergo the lowest extent of deposition in the respiratory tract.
 5           Interception is deposition by physical contact with airway surfaces.  The interception
 6      potential of any particle depends on its physical size, and fibers are the chief concern in relation
 7      to the interception process. Their aerodynamic size is determined predominantly by their
 8      diameter, but their length is the factor that influences probability of interception deposition.
 9           Electrostatic precipitation is deposition related to particle charge.  The minimum charge an
10      aerosol particle can have is zero. This condition rarely is achieved because of the random
11      charging of aerosol particles by air ions.  Aerosol particles will acquire charges from these ions
12      by collisions with them because of their random  thermal motion. Furthermore, many laboratory-
13      generated aerosols are charged.  Such aerosols will generally lose their charge as they attract
14      oppositely charged ions, and an equilibrium state of these competing processes eventually is
15      achieved.  This Boltzmann equilibrium represents the charge distribution of an aerosol in charge
16      equilibrium with bipolar ions. The  minimum amount of charge is very small, with a statistical
17      probability that some particles within the aerosol will have no charge and others will have one or
18      more positive and negative charges.
19           The electrical charge on some particles will result in  an enhanced  deposition over what
20      would be expected from size alone. This results from image charges induced on the surface of
21      the airway by these particles or to space-charge effects, whereby repulsion of particles containing
22      like charges results in increased migration toward the airway wall.  The effect of charge on
23      deposition is inversely proportional to particle size and airflow rate. This type of deposition is
24      often small compared to the effects of turbulence and other deposition mechanisms, and it
25      generally has been considered to be a minor contributor to overall particle deposition. However,
26      a study by Cohen et al. (1998), employing hollow airway casts of the human tracheobronchial
27      tree to assess deposition of ultrafine (0.02 //m) and fine (0.125 //m), particles found the
28      deposition of singly charged particles to be 5 to 6 times that of particles having no charge and
29      2 to 3 times that of particles at Boltzmann equilibrium. This suggests that electrostatic
30      precipitation may, in fact, be a significant deposition mechanism for ultrafine, and some fine,
31      particles within the TB region.

        April 2002                                 6-7         DRAFT-DO NOT QUOTE OR CITE

-------
 1      6.2.2  Deposition Patterns in the Human Respiratory Tract
 2           Knowledge of sites where particles of different sizes deposit in the respiratory tract and the
 3      amount of deposition therein is necessary for understanding and interpreting the health effects
 4      associated with exposure to particles. Particles deposited in the various respiratory tract regions
 5      are subjected to large differences in clearance mechanisms and pathways and, consequently,
 6      retention times. This section summarizes concepts of particle deposition in humans and
 7      laboratory animals as reported in the 1996 PM AQCD (U.S. Environmental Protection Agency,
 8      1996) and provides additional information based on studies published since that earlier
 9      document.
10           Ambient air often contains particles too massive to be inhaled. The descriptor
11      "inhalability" is used to denote the overall spectrum of particle sizes that are potentially capable
12      of entering the respiratory tract. Inhalability is defined as the ratio of the number concentration
13      of particles of a certain aerodynamic diameter that are inspired through the nose or mouth to the
14      number concentration of the same diameter particle present in ambient air (International
15      Commission on Radiological Protection, 1994).  In general, for humans, unit density particles
16      >100 (j,m diameter have a low probability of entering the mouth or nose in  still air, but there is no
17      sharp cutoff to zero probability. Also, there is no lower limit to inhalability, so long as the
18      particle exceeds a critical  size where the aggregation of atomic or molecular units is stable
19      enough to endow it with "particulate" properties, in contrast to those of free ions or gas
20      molecules.
21
22      6.2.2.1 Total Respiratory Tract Deposition
23           Total human respiratory tract deposition, as a function of particle size, is depicted in
24      Figure 6-2. These data were obtained by various investigators using different sizes of spherical
25      test particles in healthy male adults under different ventilation conditions; the large standard
26      deviations reflect interindividual and breathing pattern-related variability of deposition
27      efficiencies.  Deposition in the ET region with nose breathing is generally higher than that with
28      mouth breathing because of the superior filtration capabilities of the nasal passages, resulting in
29      somewhat higher total  deposition with mouth breathing for particles > l//m. For particles with
30      aerodynamic diameters greater than 1 //m, deposition is governed by impaction and
31      sedimentation, and it increases with increasing AED. When AED is >10 //m, almost all inhaled
        April 2002                                 6-8         DRAFT-DO NOT QUOTE OR CITE

-------
   100
    90  -
    80  -
    70  -
    60  -
    50  -
o
8"  40  -
Q
    30  -
    20  -
    10  -
            c
            o
                  0
                                           Human (oral inhalation)
                                           Human (nasal inhalation)
                           0.01
                                  0.1               1.0
                            Particle Diameter
                         10
       Figure 6-2.  Total respiratory tract deposition (as percentage deposition of amount
                   inhaled) in humans as a function of particle size. All values are means with
                   standard deviations when available. Particle diameters are aerodynamic
                   (MMAD) for those >0.5
       Source: Modified from Schlesinger (1989).
1     particles are deposited. As the particle size decreases from «0.5 //m, diffusional deposition
2     becomes dominant and total deposition depends more on the actual physical diameter of the
3     particle, with decreasing particle diameter leading to an increase in total deposition. Total
4     deposition shows a minimum for particle diameters in the range of 0.2 to 1.0 //m where, as noted
5     above, neither sedimentation, impaction, or diffusion deposition are very effective.
      April 2002
                                    6-9
DRAFT-DO NOT QUOTE OR CITE

-------
 1           Besides particle size, breathing pattern is the most important factor affecting lung
 2      deposition. Kim (2000) reported total lung deposition values in healthy adults for a wide range
 3      of breathing patterns, tidal volumes (375 to 1500 mL), flow rates (150 to 1000 mL/s), and
 4      respiratory times (2 to 12  s).  Total lung deposition increased with increasing tidal volume at a
 5      given flow rate and with increasing flow rate at a given respiratory time.  Various deposition
 6      values were correlated with a single composite parameter consisting of particle size, flow rate,
 7      and tidal volume.
 8           One of the specific size modes of the ambient aerosol that is being evaluated in terms of
 9      potential toxicity is the ultrafme mode (i.e., particles having diameters <0.1 //m). There is,
10      however, little information on total  respiratory tract deposition of such particles.  Frampton et al.
11      (2000) exposed healthy adult human males and females, via mouthpiece, to 0.0267 //m diameter
12      carbon particles (at 10 //g/m3) for 2 h at rest. The inspired and expired particle number
13      concentration and size distributions were evaluated. Total respiratory tract deposition fraction
14      was determined for six particle size fractions, ranging from 0.0075 to 0.1334 //m. They found an
15      overall total lung deposition fraction of 0.66 (by particle number) or 0.58 (by particle mass),
16      indicating that exhaled mean particle diameter was slightly larger than inhaled diameter. There
17      was no gender difference. The deposition  fraction decreased with increasing particle size within
18      the ultrafme range, from 0.76 at the smallest size to 0.47 at the largest.
19           Jaques and Kim (2000) measured total deposition fraction (TDF) of ultrafme particles
20      [number median  diameter (NMD) = 0.04-0.1 //m and og = 1.3] in 22 healthy adults (men and
21      women in equal number) under a variety of breathing conditions. The study was designed to
22      obtain a rigorous data set  for ultrafme particles that could be applied to health risk assessment.
23      TDF was measured for six different breathing patterns: tidal volume (Vt) of 500 mL at
24      respiratory flow rates (Q) of 150 and 250 mL/s; V, = 750 mL at Q of 250 and 375 mL/s; V, = 1 L
25      at Q of 250 and 500 mL/s. Aerosols were  monitored continuously by a modified condensation
26      nuclei counter during mouthpiece inhalation with the prescribed breathing patterns. For a given
27      breathing pattern, TDF increased as particle  size decreased, regardless of the breathing pattern
28      used. For example, at V, = 500 mL and Q  = 250 mL/s, TDF was 0.26, 0.30,  0.35, and 0.44 for
29      NMD = 0.10, 0.08, 0.06, and 0.04 //m, respectively (see Figure 6-3). For a given particle size,
30      TDF increased with an increase in V, and a decrease in Q, indicating an importance of breathing
31      pattern in assessing respiratory dose. The  study also found that TDF was greater for women than

        April 2002                                6-10        DRAFT-DO NOT QUOTE OR CITE

-------
\.u

o
03
^ OR
g
"H— ' n c;
w
o
°- HA

-------
 1           A property of some ambient particulate species that affects deposition is hygroscopicity, the
 2      propensity of a material for taking up and retaining moisture under certain conditions of humidity
 3      and temperature. Such particles can increase in size in the humid air within the respiratory tract
 4      and, when inhaled, will deposit according to their hydrated size rather than their initial size.  The
 5      implications of hygroscopic growth on deposition have been reviewed extensively by Morrow
 6      (1986) and Hiller (1991); whereas the complications of studying lung deposition of hygroscopic
 7      aerosols have been reviewed recently by Kim (2000). In general, compared to nonhygroscopic
 8      particles of the same initial size, the deposition of hygroscopic aerosols in different regions of the
 9      lung may be higher or lower, depending on the initial size. Thus, for particles with initial sizes
10      larger than «0.5 //m, the influence of hygroscopicity would be to increase total deposition with a
11      shift from peripheral to central or extrathoracic regions; whereas for smaller ones total deposition
12      would tend to be decreased.
13
14      6.2.2.2 Deposition in the Extrathoracic Region
15           The fraction of inhaled particles depositing in the ET region is quite variable, depending on
16      particle size, flow rate, breathing frequency and whether breathing is through the nose or the
17      mouth (Figure 6-4).  Mouth breathing bypasses much of the filtration  capabilities of the nasal
18      airways, leading to increased deposition in the lungs (TB and A regions). The ET region is
19      clearly the site of first contact with particles in the inhaled air and essentially acts as a "prefilter"
20      for the lungs.
21           Since release of the 1996 PM  AQCD, a number of studies have  explored ET deposition
22      with in vivo  studies, as well as in both physical and  mathematical model systems. In one study,
23      the relative distribution of particle deposition between the  oral and nasal passages was assessed
24      during "inhalation" by use of a physical model (silicone rubber) of the human upper respiratory
25      system, extending from the nostrils  and mouth through the main bronchi (Lennon et al.,  1998).
26      Monodisperse particles ranging in size from 0.3 to 2.5 //m were used  at various flow rates
27      ranging from 15 to 50 L/min.  Total deposition in the model, as was regional  deposition  in the
28      oral passages, lower oropharynx-trachea, nasal passages, and nasopharynx-trachea, were
29      assessed. Deposition within the nasal passages was found to agree with available data obtained
30      from a human inhalation study (Heyder and Rudolf, 1977), being proportional to particle size,
31      density, and inspiratory flow rate. It also was found that for oral inhalation, the relative

        April 2002                                6-12        DRAFT-DO NOT QUOTE OR CITE

-------
                   100
                    90  -
                    80  -
                    70  -
                ^  60  H
                c
                o
                S  50  H
                o
                §•  40  H
                Q
                    30  H
20  -
10  -
 0
                                  Human (oral inhalation)
                                  Human (nasal inhalation)
1
                                                                    'X T
                                                                    A*
                                                                        I  I  I  I III
                        0.01
                       0.1                1.0
                    Particle Diameter (pm)
                         10
       Figure 6-4.  Extrathoracic deposition (as percentage deposition of the amount inhaled)
                   in humans as a function of particle size. All values are means with standard
                   deviations, when available.  Particle diameters are aerodynamic (MMAD)
                   for those >0.5 (j,m and geometric (or diffusion equivalent) for those < 0.5 (j,m.
       Source: Modified from Schlesinger (1989).
1     distribution between the oral cavity and the oropharynx-trachea was similar; whereas for nasal
2     inhalation, the nasal passages contained most of the particles deposited in the model, with only
3     about 10% depositing in the nasopharynx-trachea region. Furthermore, the deposition efficiency
4     of the nasopharynx-trachea region was greater than that of the oropharynx-trachea region.
5     For simulated oronasal breathing, deposition in the ET region depended primarily on particle size
6     rather than flow rate. For all flows and for all breathing modes, total deposition in the ET region
      April 2002
                          6-13
   DRAFT-DO NOT QUOTE OR CITE

-------
 1      increased as particle diameter increased.  Such information on deposition patterns in the ET
 2      region is useful in refining empirical deposition models.
 3           Deposition within the nasal passages was further evaluated by Kesavanathan and Swift
 4      (1998), who examined the deposition of 1- to 10-//m particles in the nasal passages of normal
 5      adults under an inhalation regime in which the particles were drawn through the nose and out
 6      through the mouth at flow rates ranging from 15 to 35 L/min. At any particle size, deposition
 7      increased with increasing flow rate; whereas at any flow rate, deposition increased with
 8      increasing particle size. In addition, as was shown experimentally by Lennon et al. (1998) under
 9      oronasal breathing conditions, deposition of 0.3- to 2.5-//m particles within the nasal passages
10      was significantly greater than within the oral passages, and nasal inhalation resulted in greater
11      total deposition in the model than did oral inhalation. These results are consistent with other
12      studies discussed in the 1996 PM AQCD and with the known dominance of impaction deposition
13      within the ET region.
14           Rasmussen et al. (2000) measured deposition in the nasal cavity of normal adult humans of
15      0.7 //m particles consisting of sodium chloride and radioactively-labeled DTPA.  Inspiration
16      occurred under different levels of flow  rate ranging from 10-30 L/min. They found that the
17      deposition fraction in the nasal passages increased as flow rate increased and that an estimate of
18      maximum linear air velocity was the best single predictor of nasal deposition fraction.
19           For ultrafine particles (dp < 0.1 //m), deposition in the ET region is controlled by diffusion,
20      which depends only  on the particle's geometric diameter.  Prior to 1996, ET deposition for this
21      particle size range had not been studied extensively in humans, and this remains the case. In the
22      earlier  1996 PM AQCD, the only data available for ET deposition of ultrafine particles were
23      from cast studies.  More recently, deposition in the ET region was examined using mathematical
24      modeling. Three dimensional numerical simulations of flow and particle diffusion in the human
25      upper respiratory tract, which included  the nasal region, oral region, larynx, and first two
26      generations of bronchi, were performed by Yu et al. (1998). Deposition of particles of 0.001 and
27      0.1 //m in these different regions was calculated under inspiratory and expiratory flow conditions.
28      Deposition efficiencies in the total  model were lower on expiration than inspiration although
29      values for the former were quite high. Nasal deposition of ultrafine particles can also be quite
30      high. For example, nasal  deposition accounted for up to 54% of total deposition in the model
31      system for 0.001-//m particles.  The total deposition efficiency in the model was 75% (of the

        April 2002                                6-14        DRAFT-DO NOT QUOTE OR CITE

-------
 1      amount entering) for this size particle.  With oral breathing, deposition efficiency was estimated
 2      at 48% (of amount entering) (Yu et al., 1998).
 3           Swift and Strong (1996) examined the deposition of ultrafine particles, ranging in size from
 4      0.053 to 0.062 //m, in the nasal passages of normal adults during constant inspiratory flows of
 5      6 to 22 L/min. The results are consistent with results noted in studies above, namely that the
 6      nasal passages are highly efficient collectors for ultrafine particles. In this case, fractional
 7      deposition ranged from 94 to 99% (of amount inhaled). There was found to be only a weak
 8      dependence of deposition on flow rate, which contrasts with results noted above (i.e., Lennon
 9      et al., 1998) for particles >0.3 //m, but is consistent with diffusion as the main deposition
10      mechanism.
11           Cheng et al. (1997) examined oral airway deposition in a replicate cast of the human nasal
12      cavity,  oral cavity, and laryngeal-tracheal sections. Particle sizes ranged from 0.005 to 0.150 //m,
13      and constant inspiratory and expiratory flow rates of 7.5 to 30 L/min were used.  They noted that
14      the deposition fractions within the oral  cavity were essentially the same as that in the
15      laryngeal-tracheal sections for all particle sizes and flow rates. They ascribed this to the balance
16      between flow turbulence and residence time in these two regions. Svartengren et al. (1995)
17      examined the  effect of changes in external resistance  on oropharyngeal deposition of 3.6-//m
18      particles in asthmatics.  Under controlled mouthpiece breathing conditions (flow rate 0.5 L/s), the
19      median deposition as a percentage of inhaled particles in the mouth and throat was 20%
20      (mean = 33%; range 12 to 84%).  Although the mean deposition fell to 22% with added
21      resistance, the median value remained at 20% (range  13 to 47%). Fiberoptic examination of the
22      larynx revealed that there was a trend for increased mouth and throat deposition associated with
23      laryngeal narrowing. Katz et al. (1999) indicate, on the basis of mathematical model
24      calculations, that turbulence plays a key role in enhancing particle deposition in the larynx and
25      trachea.
26           The results of all of the above studies support the previously known ability of the ET
27      region, and especially the nasal passages, to act as an efficient filter for nanoparticles (<0.1 //m)
28      as well as for  larger ones (>5//m), potentially reducing the amount of particles within a wide size
29      range that are available for deposition in the TB and A regions.
30
31

        April 2002                                 6-15        DRAFT-DO NOT QUOTE OR CITE

-------
 1      6.2.2.3 Deposition in the Tracheobronchial and Alveolar Regions
 2           Particles that do not deposit in the ET region of the respiratory tract enter the lungs;
 3      however, their regional deposition within the lungs cannot be precisely measured.  Much of the
 4      available deposition data for the TB and A regions have been obtained from experiments with
 5      radioactively labeled,  poorly soluble particles (Figures 6-5 and 6-6, respectively). These have
 6      been described previously (U.S. Environmental Protection Agency, 1996). Although there are no
 7      new regional data obtained by means of the radioactive aerosol method since the publication of
 8      that document, a novel serial bolus delivery method has been introduced. Using this bolus
 9      technique, regional deposition has been measured for fine and coarse aerosols (Kim et al., 1996;
10      Kim and Hu, 1998) and for ultrafme aerosols (Kim and Jacques, 2000).  The serial bolus method
11      uses nonradioactive aerosols and can measure regional deposition in a virtually unlimited number
12      of lung compartments. Because of experimental limitations of the technique, the investigators
13      measured regional lung deposition in ten serial, 50 mL increments from the mouth to the end of a
14      typical 500 mL tidal volume. Deposition measurements in the TB and A regions were obtained
15      for both men and women for particles ranging from 0.04 to 5.0 //m in diameter.  It should be
16      noted that particle deposition in the TB and A regions was based on volumetric compartments of
17      50 to 150 mL and >150 mL, respectively. Deposition in the ET region was based on the 0 to
18      50 mL compartment.  Lung deposition fractions in the  TB and A regions obtained by the bolus
19      technique are shown in Figure 6-7.  Of total particle deposition in the lung, 23 to 32% was
20      deposited in the TB region and 68 to 77% was deposited in the A region.  Deposition in women
21      was consistently greater in the TB region by 21 to 47%, but was comparable or slightly smaller in
22      the A region when compared to men.  As a result, total lung deposition was slightly greater in
23      women than men (~5  to 15%).
24
25      6.2.2.4 Local Distribution of Deposition
26           Airway structure and its associated air flow patterns are exceedingly complex, and
27      ventilation distribution of air in different parts  of the lung is uneven.  Thus, it is expected that
28      particle deposition patterns within the ET, TB, and A regions would be highly nonuniform, with
29      some sites exhibiting deposition that is much greater than average levels within these regions.
30      This was discussed in detail previously in the 1996 PM AQCD.  Basically, using deposition data
31      from living subjects as well as from mathematical and  physical models, enhanced deposition has

        April 2002                                6-16        DRAFT-DO NOT QUOTE OR CITE

-------
                60
                50  -
            ^  40  H

            o
            :-E  so
            w
            o
            CL
            Q  20  -
                 10  -
                   0.01
                          A Human (oral inhalation)
                                               '\
                  0.1
  1.0
10
                                     Particle Diameter (|jm)
 Figure 6-5.    Tracheobronchial deposition (as percentage deposition of the amount
               inhaled) in humans as a function of particle size.  All values are means
               with standard deviations, when available. Particle diameters are
               aerodynamic (MMAD) for those >0.05 ^,m and geometric (or diffusion
               equivalent) for those < 0.5 (j,m.

 Source: Modified from Schlesinger (1989).
                      70
                      60  -

                      50  -
                   c  40  H
o  30  H
01
Q
   20  -
                        0.01
             Human (oral inhalation)
             Human (nasal inhalation)

                                                   1
                                      0.1            1.0
                                      Particle Diameter (pm)
                                                                   10
 Figure 6-6.  Alveolar deposition (as percentage deposition of the amount inhaled) in
             humans as a function of particle size. All values are means with standard
             deviations, when available.  Particle diameters are aerodynamic (MMAD)
             for those >0.05 ^m and geometric (or diffusion equivalent) for those
             < 0.5
 Source: Modified from Schlesinger (1989).
April 2002
                       6-17
DRAFT-DO NOT QUOTE OR CITE

-------
       o
      t3
       ro
       o
       Q.
       0>
      Q
                      Male       Vt = 500ml_
                      Female     Q = 250 mL/s
                                                          Male
                                                          Female
                 Vt = 500m L
                 Q = 250 mL/s
                  0.04     0.06    0.08     0.10
                     Particle Diameter (|jm)
                                                      1           3           5
                                                        Particle Diameter (|jm)
      Figure 6-7.  Lung deposition fractions in the tracheobronchial (TB) and alveolar (A)
                  regions obtained by the bolus technique. Using a breathing pattern of 500 mL
                  at 15 breaths per min, TB deposition was 1.5,10.6, and 26.1% and
                  A deposition was 7.7, 39.4, and 39.8% for particles of 1, 3, and 5 ^m in
                  diameter, respectively, for men. In comparison to men, TB deposition in
                  women was 26 to 53% greater, whereas A deposition was comparable.
                  For ultrafine particles of 0.04 to 0.1 (j,m diameter, TB and A deposition ranged
                  from 5.7 to 15.6% and 18.2 to 33.1%, respectively. Both TB and A deposition
                  decreased with increasing particle size within the ultrafine range, which is
                  consistent with deposition theory.

      Source: Kim and Hu (1998); Kim and Jaques (2000).
1

2
been shown to occur in the nasal passages and trachea and at branching points in the TB and
A regions (see Chapter 10 of U.S. Environmental Protection Agency, 1996).  Churg and Vedal
      April 2002
                                        6-18
DRAFT-DO NOT QUOTE OR CITE

-------
 1      (1996) examined retention of particles on carinal ridges and tubular sections of airways from
 2      lungs obtained at necropsy. Results indicated significant enhancement of particle retention on
 3      carinal ridges through the segmental bronchi; the ratios were similar in all airway generations
 4      examined.
 5           Kim and Fisher (1999) studied local deposition efficiencies and deposition patterns of
 6      aerosol particles (2.9 to 6.7 //m) in sequential double bifurcation tube models with two different
 7      branching geometries:  one with in-plane (A) and another with out of plane (B) bifurcation. The
 8      deposition efficiencies (DE) in each bifurcation increased with increasing Stokes number (Stk).
 9      With symmetric flow conditions, DE was somewhat smaller in the second than the first
10      bifurcation in both models. DE was greater in the second bifurcation in model B than in model
11      A. With asymmetric flows, DE was greater in the low-flow side compared to the high-flow side;
12      and this was consistent in both models. Deposition pattern analysis showed highly localized
13      deposition on and in the immediate vicinity of each bifurcation ridge, regardless  of branching and
14      flow patterns.
15           Comer et al. (2000) used a three-dimensional computer simulation technique to investigate
16      local  deposition patterns  in sequentially bifurcating airway models that were previously used in
17      experiments by Kim and Fisher (1999).  The simulation was for 3-, 5-, and 7-//m particles and
18      assumed steady, laminar, constant air flow with symmetry about the first bifurcation. The overall
19      trend of the particle deposition efficiency, i.e., an exponential increase with Stokes number, was
20      similar for all bifurcations,  and deposition efficiencies in the bifurcation regions  agreed very well
21      with experimental data. Local deposition patterns consistently showed that the majority of the
22      deposition occurred within  the carinal  region.
23           Deposition "hot spots" at airway bifurcations have undergone additional analyses using
24      mathematical modeling techniques. Using calculated deposition sites, a strong correlation has
25      been demonstrated between secondary flow patterns and deposition sites and density both for
26      large (10 //m) particles and for ultrafme particles (0.01  //m) (Heistracher and Hofmann, 1997;
27      Hofmann et al.,  1996). This supports experimental work, noted in U.S. Environmental
28      Protection Agency (1996), indicating that, like larger particles, ultrafme particles also show
29      enhanced deposition at airway branch points — even in the upper tracheobronchial tree.
30           The pattern of particle distribution on a more regional scale was evaluated by Kim et al.
31      (1996) and Kim and Hu (1998). Deposition patterns were measured in situ in nonsmoking

        April 2002                                6-19         DRAFT-DO NOT QUOTE OR CITE

-------
 1      healthy young adult males using, an aerosol bolus technique that delivered 1-, 3-, or 5-fj.m
 2      particles into specific volumetric depths within the lungs. The distribution of particle deposition
 3      was uneven; and it was noted that sites of peak deposition shifted from distal to proximal regions
 4      of the lungs with increasing particle size (Figure 6-8). Furthermore, the surface dose was found
 5      to be greater in the conducting airways than in the alveolar region for all of the particle sizes
 6      evaluated. Within the conducting airways, the largest airway regions (i.e., 50 to  100 mL volume
 7      distal to the larynx) received the greatest surface doses.
 8           Kim and Jaques (2000) used the respiratory bolus technique to measure the deposition
 9      distribution of ultrafme particles (0.04, 0.06, 0.08, and 0.1 //m) in young adults.  Under normal
10      breathing conditions (tidal volume 500 mL, respiratory flow rate 250  mL/s), bolus aerosols were
11      delivered sequentially to a lung depth ranging from 50 to 500 mL in 50-mL increments. The
12      results indicate that regional deposition varies widely along the depth of the  lung, regardless of
13      particle size  (Figure 6-8). The deposition patterns for ultrafme particles, especially for very small
14      ultrafme particles, were similar to those for coarse particles. Peak deposition occurred in the
15      lung regions situated between 150 and  200 mL from the mouth,  and sites of peak deposition
16      shifted proximally with a decrease in particle size.  Deposition dose per unit average surface area
17      was greatest in the proximal lung regions and decreased rapidly  with increased lung depth.  Peak
18      surface dose was 5 to 7 times greater than average lung dose.  These results indicate that local
19      enhancement of dose occurs in healthy lungs, which could be an important factor in eliciting
20      pathophysiological effects.
21
22      6.2.2.5 Deposition of Specific Size Modes of Ambient Aerosol
23           The studies described  in previous sections generally evaluated deposition using individual
24      particle sizes within certain  ranges without consideration of specific relevant ambient size ranges.
25      Some recent modeling studies, however, have considered the deposition profiles of particle
26      modes that exist in ambient  air, so as to provide estimates on dosimetry of these  "real  world"
27      particle size  fractions. One  such study using a lung-anatomical model (Venkataraman and Kao,
28      1999) examined the contribution of two specific size modes of the PM10 ambient aerosol, namely
29      the fine mode (defined as particles with diameters up to 2.5 //m) and the thoracic fraction of the
30      coarse mode (defined as particles with  diameters 2.5  to 10 //m),  to total lung and regional lung
31      doses (i.e., a daily dose expressed as //g/day, and a surface dose  expressed a //g/cm2/day)

        April 2002                                6-20        DRAFT-DO NOT QUOTE OR CITE

-------
                 c
                .0
                "o
                 CO
                 c
                 o
                '(7)
                 o
                 Q.
                 0
                Q
                "CD
                 o
                 o
0.1-
                      0.0
                                Dp = 3|jm
                                100
                                       200
                                              300
                                                      400
                                                             500
                            Volumetric Lung Region (ml_)
 Figure 6-8. Lung deposition fractions in ten volumetric regions for particle sizes ranging
            from ultrafine particle diameter (dp) of 0.04 to 0.01 (j,m (Panel A) to fine
            (dp = 1.0 jum) (Panel B) and coarse (dp = 3 and 5 ^m) (Panels C and D).
            Healthy young adults inhaled a small bolus of monodisperse aerosols under
            a range of normal breathing conditions (ie., tidal volume of 500 mL at
            breathing frequencies of 9,15, and 30 breaths per niin.).

 Source: Kim et al. (1996); Kim and Hu (1998); Kim and Jacques (2000).
April 2002
                 6-21
DRAFT-DO NOT QUOTE OR CITE

-------
 1      resulting from a 24-h exposure to a particle concentration of 150 //g/m3.  The study also
 2      evaluated deposition in terms of two metrics, namely mass dose and number dose. Deposition
 3      was calculated using a mathematical model for a healthy human lung under both simulated
 4      moderate exertion (1 L at 15 breaths/min) and vigorous exertion (1.5 L at 15 breaths/min), and
 5      for a compromised lung (0.5 L at 30 breaths/min). Regional deposition values were obtained for
 6      the ET, TB, and A regions. Because the exposure scenario used is quite unrealistic, only general
 7      trends should be inferred from this study rather than actual deposition values.
 8           Daily mass dose peaked in the A airways for all breathing patterns; whereas that for the
 9      coarse fractions was comparable in the TB and A regions. The mass per unit surface area of
10      various airways from the fine and coarse fractions was larger in the trachea and first few
11      generations of bronchi. It was suggested that these large surface doses may be related to
12      aggravation of upper respiratory tract illness in geographical areas where coarse particles are
13      present.
14           The daily number dose was different for fine and coarse fractions in all lung airways, with
15      the dose from the fine fraction higher by about 100 times in the ET and about 10s times in
16      internal lung airways.  The surface number dose (particles/cm2/day) was 103 to  10s times higher
17      for fine than for coarse particles in all lung airways, indicating the larger number of fine particles
18      depositing. Particle number doses did not  follow trends in mass doses and are much higher for
19      fine than coarse particles and are higher for different breathing patterns.  It also was concluded
20      that the fine fraction contributes 10,000 times greater particle number per alveolar macrophage
21      than the coarse fraction particles. As noted, these results must be viewed with caution because
22      they were obtained using a pure mathematical model that must be validated in terms of realistic
23      physiologic conditions.
24           Another evaluation of deposition that included consideration of size mode of the ambient
25      aerosol was that of Broday and  Georgopoulos (2001).  In this case, a mathematical model was
26      used to account for particle hygroscopic growth, transport, and deposition in tracking the changes
27      in the size distribution of inhaled aerosols. It was concluded that different rates of particle
28      growth in the inspired air resulted in a change in the aerosol size  distribution, such that increased
29      mass and number fractions of inspired ultrafine particles (< 0.1 //m) were found in the size range
30      between 0.1 to 1 //m and, therefore, deposited to a lesser extent due to a decrease in diffusion
31      deposition. On the other hand,  particles that were originally in the 0.1 to 1 //m  size range when

        April 2002                                6-22        DRAFT-DO NOT QUOTE OR CITE

-------
 1      inhaled will undergo enhanced deposition because of their increase in size resulting from
 2      hygroscopic growth. Hence, the initial size distribution of the inhaled polydisperse aerosol
 3      affects the evolution of size distribution once inhaled and, thus, its deposition profile in the
 4      respiratory tract.  Hygroscopicity of respirable particles must be considered for accurate
 5      predictions of deposition. Because different size fractions likely have different chemical
 6      composition, such changes in deposition patterns will affect biological responses.
 7
 8      6.2.3 Biological Factors Modulating Deposition
 9           Experimental deposition data in humans are commonly derived using healthy adult
10      Caucasian males.  Various factors can act to alter deposition patterns from those obtained in this
11      group. Evaluation of these factors is important to help understand potentially susceptible
12      subpopulations because differences  in biological response following pollutant exposure may be
13      caused by dosimetry differences as well as by differences in innate sensitivity.  The effects of
14      different biological factors on deposition  were discussed in the 1996 PM AQCD (U.S.
15      Environmental Protection Agency, 1996) and are summarized below together with additional
16      information obtained from more recent studies.
17
18      6.2.3.1 Gender
19           Males and females have different body size and ventilatory parameter distributions;
20      therefore, it is expected that there would be gender differences in deposition.  In some of the
21      controlled studies, however, men and women are breathing at the same tidal volume and
22      frequency. If the women are generally smaller than the men, the increased minute ventilation
23      compared to their normal ventilation would cause different changes in deposition patterns.
24      In these cases, it would be better for the investigators to have used size-adjusted tidal volumes.
25      This may help to explain some of the differing results discussed below.
26           Using particles in the 2.5- to 7.5-//m size range, Pritchard et al. (1986) indicated that, for
27      comparable particle sizes and inspiratory  flow rates, females had higher ET and TB deposition
28      and smaller A deposition than did males.  The ratio of A deposition to total thoracic deposition in
29      females also was found to be smaller. These differences were attributed to gender differences in
30      airway size.

        April 2002                                6-23        DRAFT-DO NOT QUOTE OR CITE

-------
 1           In another study (Bennett et al., 1996), the total respiratory tract deposition of 2-//m
 2      particles was examined in adult males and females aged 18 to 80 years who breathed with a
 3      normal resting pattern. Deposition was assessed in terms of a deposition fraction, which was the
 4      difference between the amount of particles inhaled and exhaled during oral breathing. Although
 5      there was a tendency for a greater deposition fraction in females compared to males, and because
 6      males had greater minute ventilation, the deposition rate (i.e., deposition per unit time) was
 7      greater in males than in females.
 8           Kim and Hu (1998) assessed regional deposition patterns in healthy adult males and
 9      females using particles with median aerodynamic sizes of 1, 3, and 5 //m and a bolus delivery
10      technique that involved controlled breathing.  The total deposition in the lungs was similar for
11      both genders with the smallest particle, but was greater in women for the 3- and 5-//m particles,
12      regardless of the inhalation flow rate used; this difference ranged from 9 to 31%, with higher
13      values associated with higher flow rates. The pattern of deposition was similar for both genders
14      although females showed enhanced deposition peaks for all three particle sizes. The volumetric
15      depth location of these peaks was found  to shift from peripheral (i.e., increased volumetric depth)
16      to proximal (i.e., shallow volumetric depth) regions of the lung with increasing particle size, but
17      the shift was greater in females than in males.  Thus, deposition appeared to be more localized in
18      the lungs of females compared to those of males. These differences were attributed to a smaller
19      size of the upper airways in females than in males, particularly of the laryngeal structure. Local
20      deposition of l-//m particles was somewhat flow dependent but, for larger (5-//m) particles, was
21      largely independent of flow (flows did not include those that would be typical of exercise).
22           In a related study, Kim et al. (2000) evaluated differences in deposition between males and
23      females in terms of exercise levels of ventilation and breathing patterns.  Using particles at the
24      same size noted above and a number of breathing conditions, total lung deposition was
25      comparable between men and women for l-//m particles, but was slightly greater in women than
26      men for 3- and 5-//m particles with all breathing patterns.  The gender difference was about 15%
27      at rest, and variable during exercise, depending on particle size. However, total lung deposition
28      rate (i.e., deposition per unit time) was found to be 3 to 4 times greater during moderate exercise
29      than during rest for all particle sizes.  Thus, it was concluded that exercise may increase the
30      health risk from particles because of increased large airway deposition and that women may be
31      more susceptible to this exercise-induced change.

        April 2002                                 6-24         DRAFT-DO NOT QUOTE  OR CITE

-------
 1           Jaques and Kim (2000) and Kim and Jaques (2000) expanded the evaluation of deposition
 2      in males and females to particles <1 //m.  They measured total lung deposition in healthy adults
 3      using sizes in the ultrafme mode (0.04 to 0.1 //m), in addition to those having diameters of 1 and
 4      5 //m. Total lung deposition was greater in females than in males for 0.04- and 0.06-//m
 5      particles. The difference was negligible for 0.08-and 0.1-//m particles.  Therefore, the gender
 6      effect was particle-size dependent, showing a greater deposition in females for very small
 7      ultrafme and large coarse particles, but not for particles ranging from 0.08 to 1 //m. A local
 8      deposition fraction was determined in each volumetric compartment of the lung to which
 9      particles are injected based on the inhalation procedure (Kim and Jaques, 2000).  The deposition
10      fraction was found to increase with increasing lung depth from the mouth, reach a peak value,
11      and then decrease with further increase in lung volumetric depth.  The height of the peak and its
12      depth did vary with particle size and breathing pattern.  Peak deposition for the 5-//m particles
13      was more proximal than that for the l-//m particles; whereas that for the ultrafme particles
14      occurred between these two peaks. For the ultrafme particles,  the peak deposition became more
15      proximal as particle size decreased. Although this pattern of deposition distribution was similar
16      for both men and women, the region of peak deposition was shifted closer to the mouth and peak
17      height was slightly greater for women than for men for all exposure conditions.
18
19      6.2.3.2 Age
20           Airway structure and respiratory conditions vary with age, and these variations may alter
21      the deposition pattern of inhaled particles. The limited experimental studies reported in the 1996
22      PM AQCD  (U. S. Environmental Protection Agency,  1996) indicated results ranging from no
23      clear dependence of total deposition on age to slightly higher deposition in children than adults.
24      However, children have a different resting ventilation than do adults. The experimental studies
25      must adjust for the higher minute ventilation per unit body weight in children when comparing
26      deposition results to those obtained in adults.
27           Potential regional deposition differences between children and adults have been assessed to
28      a greater extent using mathematical models.  These indicated that, if the entire respiratory tract
29      and a complete breathing cycle at normal rate are considered, then ET deposition in children
30      would be generally higher than that in adults, but TB and A regional deposition in children may
31      be either higher or lower than that in adults, depending on particle size (Xu and Yu, 1986).

        April 2002                                 6-25        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Enhanced deposition in the TB region would occur for particles <5 //m in children (Xu and Yu,
 2      1986; Hofmann et al., 1989a).
 3           An age dependent theoretical model to predict regional particle deposition in childrens'
 4      lungs that incorporates breathing parameters and morphology of the growing lung was developed
 5      by Musante and Martonen (1999).  The model was used to compare deposition of monodisperse
 6      aerosols, ranging from 0.25 to 5 //m, in the lungs of children (aged 7, 22, 48, and 98 mo) at rest
 7      to that in adults (aged 30 years) at rest. Compared to adults, A deposition was highest in the
 8      48- and 98-mo subjects for all particle sizes; TB deposition was found to be a monotonically
 9      decreasing function of age for all sizes; and total lung deposition (i.e., TB+A) was generally
10      higher in children than adults, with children of all  ages showing similar total deposition fractions.
11           This model was used by Musante and Martonen (2000a) to evaluate the deposition of a
12      polydisperse aerosol that has been extensively used in toxicological studies, namely residual oil
13      fly ash (ROFA) having an MMAD of 1.95 //m, a geometric standard deviation of 2.19, and a
14      CMD of 0.53 (assuming a particle density of 0.34  g/cm2).  Deposition was evaluated under
15      resting breathing conditions.  The mass based deposition fraction of the particles was found to
16      decrease with age from 7 mo to adulthood, but the mass deposition per unit surface area in the
17      lungs of children could be significantly greater than that in the adult.
18           Phalen and Oldham (2001) calculated the respiratory deposition of particles with sizes
19      ranging from 0.1 to 10 //m in diameter for 20 year-old adults and 2 year-old children.  Total lung
20      deposition was comparable between adults and children for all particle  sizes tested; however, TB
21      deposition was much greater in children than in  adults (from 13 to 81%, depending on particle
22      size). Particle deposition in the A region was significantly reduced in children.
23           Cheng et al. (1995) examined deposition of ultrafme particles in replica casts of the nasal
24      airways of children aged  1.5 to 4 years.  Particle sizes ranged from 0.0046 to 0.2 //m, and both
25      inspiratory and expiratory flow rates were used (3  to 16 L/min). Deposition efficiency was found
26      to decrease with increasing age for a given particle size and flow rate.
27           Oldham et al. (1997) examined the deposition of monodisperse particles having diameters
28      of 1,  5, 10, and 15 //m in hollow airway models  that were  designed to represent the trachea and
29      the first few bronchial airway generations of an adult, a 7-year-old child, and a 4-year-old child.
30      They noted that, in most cases, the total  deposition efficiency was greater in the child-size
31      models than in the adult model.

        April 2002                                 6-26        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Bennett et al. (1997a) analyzed the regional deposition of poorly soluble 4.5 //m particles
 2      inhaled via mouthpiece.  The subjects were children and adults with mild cystic fibrosis (CF), but
 3      who likely had normal upper airway anatomy such that intra- and extrathoracic deposition would
 4      be similar to that in healthy people.  The mean age of the children was 13.8 years and for the
 5      adults was 29.1 years.  Extrathoracic deposition, as a percentage of total respiratory tract
 6      deposition, was higher by about 50% in children compared to adults (30.7%, 20.1%, and 16.0%,
 7      respectively). There was an age dependence of ET deposition in the children, in that the
 8      percentage ET  deposition tended to be higher at a younger age (p = 0.08); the younger group
 9      (<14 years) (p = 0.05) had almost twice the percentage ET deposition of the older group
10      (>14 years).  Additional analyses showed an inverse correlation of extrathoracic deposition with
11      body height. There was no significant difference in lung or total respiratory tract deposition
12      between the children and adults.  Because ET deposition was age dependent, and total deposition
13      was not, this suggests that the ET region does a more effective job in children of filtering out
14      particles that would otherwise reach the TB region. However, because the lungs of children are
15      smaller than are those of adults, children may still have comparable deposition per unit surface
16      area as adults.
17           Bennett and Zeman (1998) measured the deposition of monodisperse 2 //m (MMAD)
18      particles in children (aged 7 to 14 years) and adolescents (aged 14 to 18 years) for comparison to
19      that in adults (19 to 35 years). Each subject inhaled the particles by following their previously
20      determined individual spontaneous resting breathing pattern.  Deposition was assessed by
21      measuring the amount of particles inhaled and exhaled.  There was no age-related difference in
22      deposition within the children group. There was also no significant difference in deposition
23      between the children and adolescents, between the children and adults, or between the
24      adolescents and adults. However, the investigators noted that, because the children had smaller
25      lungs and higher minute volumes relative to lung size, they likely would receive greater doses of
26      particles per lung surface area compared to adults.  Furthermore, breath-to-breath fractional
27      deposition in children did vary with tidal volume, increasing with increasing volume. The rate of
28      deposition normalized to lung surface area tended (p = 0.07) to be greater (35%) in children
29      when compared to the combined group of adolescents and adults. These additional studies still
30      do not provide  unequivocal evidence for significant differences in deposition between adults and
31      children, even when considering differences in lung surface area. However, it should be noted

        April 2002                                6-27        DRAFT-DO NOT QUOTE OR CITE

-------
 1      that differences in levels of activity between adults and children are likely to play a fairly large
 2      role in age-related differences in deposition patterns of ambient particles. Children generally
 3      have higher activity levels during the day and higher associated minute ventilation per lung size,
 4      which can contribute to a greater size-specific dose of particles. Activity levels in relationship to
 5      exposure are discussed more fully in Chapter 5.
 6           Another subpopulation of potential concern related to susceptibility to inhaled particles is
 7      the elderly. In the study of Bennett et al. (1996), in which the total respiratory tract deposition of
 8      2-//m particles was examined in people aged 18 to 80 years, the deposition fraction in the lungs
 9      of people with normal lung function was found to be independent of age, depending solely on
10      breathing pattern and airway resistance.
11
12      6.2.3.3  Respiratory Tract Disease
13           The presence of respiratory tract disease can affect airway structure and ventilatory
14      parameters, thus altering deposition compared to that occurring in healthy individuals. The effect
15      of airway diseases on deposition has been studied extensively, as described in the 1996 PM
16      AQCD (U.S. Environmental Protection Agency, 1996).  Studies described therein had shown that
17      people with chronic obstructive pulmonary disease (COPD) had very heterogeneous deposition
18      patterns, and differences in regional deposition compared to normals.  People with asthma and
19      obstructive pulmonary disease tended to have greater TB deposition than did healthy people.
20      Furthermore, there tended to be an inverse relationship between bronchoconstriction and the
21      extent of deposition in the A region; whereas total respiratory tract deposition generally increased
22      with increasing degrees of airway  obstruction. The described studies were performed during
23      controlled breathing; i.e., all subjects breathed with the same tidal volume and respiratory rate.
24      However, although resting tidal volume is similar or elevated in people with COPD compared to
25      normal,  healthy individuals the former tend to breathe at a faster rate, resulting in higher than
26      normal tidal peak flow and resting minute ventilation. Thus, some of the reported differences in
27      the deposition of particles could have been caused by increased fractional deposition with each
28      breath.  Although the extent to which lung deposition may change with respect to particle size,
29      breathing pattern, and disease status in people with COPD is still unclear, some recent studies
30      have attempted to provide additional insight into this issue.


        April 2002                                 6-28         DRAFT-DO NOT QUOTE OR CITE

-------
 1           Bennett et al. (1997b) measured the fractional deposition of insoluble 2-//m particles in
 2      people with severe to moderate COPD (mix of emphysema and chronic bronchitis, mean age
 3      62 years) and compared this to healthy older adults (mean age 67 years) under conditions where
 4      the subjects breathed using their individual resting breathing pattern, as well as a controlled
 5      breathing pattern. People with COPD tended to breathe with elevated tidal volume and at a
 6      faster rate than people with healthy lungs,  resulting in about 50% higher resting minute
 7      ventilation.  Total respiratory tract deposition was assessed in terms of deposition fraction, a
 8      measure of the amount deposited based on measures of amount of aerosol inhaled and exhaled,
 9      and deposition rate, the particles deposited per unit time. Under typical breathing conditions,
10      people with COPD had about 50% greater deposition fraction than did age-matched healthy
11      adults.  Because of the elevation in minute ventilation, people with COPD  had average
12      deposition rates about 2.5 times that of healthy adults.  Similar to previously reviewed studies
13      (U.S. Environmental Protection Agency, 1996), these investigators observed an increase in
14      deposition with an increase in airway resistance, suggesting that, at rest,  COPD resulted in
15      increased deposition of fine particles in proportion to the severity of airway disease.  The
16      investigators also reported a decrease in deposition with increasing mean effective airspace
17      diameter; this suggested that the enhanced deposition was associated more with the chronic
18      bronchitic component of COPD than with the emphysematous component. Greater deposition
19      was noted with natural breathing compared to the  fixed pattern.
20           Kim and Kang (1997) measured lung deposition of l-//m particles inhaled via the mouth by
21      healthy adults (mean age 27 years) and by those with various degrees of airway obstruction,
22      namely smokers (mean age 27 years), smokers with small airway disease (SAD; mean age
23      37 years), asthmatics (mean age 48 years), and patients with COPD (mean  age 61 years)
24      breathing under the same controlled pattern. Deposition fraction was obtained by measuring the
25      number of particles inhaled and exhaled, breath by breath.  There was  a marked increase in
26      deposition in people with COPD. Deposition was 16%, 49%, 59%, and  103% greater in
27      smokers, smokers with SAD, asthmatics and people with COPD, respectively, than in healthy
28      adults.  Deposition in COPD patients was  significantly greater than that associated with either
29      SAD or asthma; there was no significant difference in deposition between people with SAD  and
30      asthma.  Deposition fraction was found to be correlated with percent predicted forced expiratory
31      volume (FEVj) and forced expiratory flow (FEF25-75%). Airway  resistance was not correlated

        April 2002                                6-29        DRAFT-DO NOT QUOTE OR CITE

-------
 1      strongly with total lung deposition. Kohlhaufl et al. (1999) showed increased deposition of fine
 2      particles (0.9 //m) in women with bronchial hyperresponsiveness.
 3           Segal et al. (2000a) developed a mathematical model for airflow and particle motion in the
 4      lung that was used to evaluate how lung cancer affects deposition patterns in the lungs of
 5      children. It was noted that the presence of airway tumors could affect deposition by increasing
 6      probability of inertial deposition and diffusion.  The former would occur on upstream surfaces of
 7      tumors and the latter on downstream surfaces. It was concluded that particle deposition is
 8      affected by the presence of airway disease, that effects may be systematic and could be predicted,
 9      and that, therefore, they could be incorporated into dosimetry models.
10           Brown et al. (2001) examined the relationship between regional lung deposition for coarse
11      particles (5 //m) and ventilation patterns in healthy adults and in patients with cystic fibrosis
12      (CF).  They found that deposition in the TB region was positively associated with regional
13      ventilation in healthy subjects, but negatively associated in CF patients. The relationships were
14      reversed for deposition in the A region. These data suggest that significant coarse particle
15      deposition may occur in the TB region of poorly ventilated lungs, as occurs in CF; whereas TB
16      deposition follows ventilation in healthy subjects.
17           Thus, the database related to particle deposition and lung disease suggests that total lung
18      deposition generally is increased with obstructed airways, regardless of deposition distribution
19      between the TB and A regions. Airflow distribution is very uneven in diseased lungs because of
20      the irregular pattern of obstruction, and there can be closure of small airways. In this situation, a
21      part of the lung is inaccessible, and particles can penetrate deeper into other, better ventilated
22      regions. Thus, deposition can be enhanced locally in regions  of active ventilation, particularly in
23      the A region. The relationships between lung deposition and  airway obstruction or ventilation
24      distribution were previously studied in vivo in animal  models (Kim, 1989; Kim et al., 1989).
25
26      6.2.3.4  Anatomical Variability
27           As indicated above, variations in anatomical parameters between genders, and between
28      healthy people and those with obstructive lung disease, can affect deposition patterns. However,
29      previous analyses generally have overlooked the effect on deposition of normal interindividual
30      variability in airway structure in healthy individuals.  This is an important consideration in
31      dosimetry modeling, which often is based on a single idealized structure.  Studies that have

        April 2002                                6-30        DRAFT-DO NOT QUOTE  OR CITE

-------
 1      become available since the 1996 PM AQCD have attempted to assess the influence of such
 2      variation in respiratory tract structure on deposition patterns.
 3           The ET region is the first to contact inhaled particles and, therefore, deposition within this
 4      region would reduce the amount of particles available for deposition in the lungs.  Variations in
 5      relative deposition within the ET region will, therefore, propagate through the rest of the
 6      respiratory tract, creating differences in calculated doses from individual to individual.
 7      A number of studies have examined the influence of variations in airway geometry on deposition
 8      in the ET region.
 9           Cheng et al. (1996) examined nasal airway deposition in healthy adults using particles
10      ranging in size from 0.004 to 0.15 //m and at two constant inspiratory flow rates, 167 and
11      33 mL/s.  Deposition was evaluated in relation to measures of nasal geometry as determined by
12      magnetic resonance imaging and acoustic rhinometry.  They noted that interindividual variability
13      in deposition was correlated with the wide variation of nasal dimensions, in that greater surface
14      area, smaller cross-sectional area, and increasing complexity of airway shape were all associated
15      with enhanced deposition.
16           Using a regression analysis of data on nasal airway deposition derived from Cheng et al.
17      (1996), Guilmette et al. (1997) noted that the deposition efficiency within this region was highly
18      correlated with both nasal airway surface area and volume; this indicated that airway size and
19      shape factors were important in explaining intraindividual variability noted in experimental
20      studies of human nasal airway aerosol deposition.  Thus, much of the variability in measured
21      deposition among people resulted from differences in the size and shape of specific airway
22      regions.
23           Kesavanathan and Swift (1998) also evaluated the influence of geometry in affecting
24      deposition in the nasal passages of normal adults from two ethnic groups. Mathematical
25      modeling of the results indicated that the shape of the nostril affected particle deposition in the
26      nasal passages, but that there still remained large intersubject variations in deposition when this
27      was accounted for,  and which was likely caused by geometric variability in  the mid and posterior
28      regions of the nasal passages.
29           Bennett et al.  (1998) studied the role of anatomic dead space (ADS) in particle deposition
30      and retention in bronchial airways, using an aerosol bolus technique. They  found that the
31      fractional deposition was dependant on the subject's ADS and that a significant  number of

        April 2002                                 6-31         DRAFT-DO NOT QUOTE OR CITE

-------
 1      particles was retained beyond 24 h.  This finding of prolonged retention of insoluble particles in
 2      the airways is consistent with the findings of Scheuch et al. (1995) and Stahlhofen et al. (1986a)
 3      and with the predictions of asymmetric stochastic human lung models (Asgharian et al., 2001).
 4      Bennett et al. (1999) also found a lung volume-dependent asymmetric distribution of particles
 5      between the left and right lung; the leftright ratio was increased at increased percentage of total
 6      lung capacity (e.g., at 70% TLC, L:R was 1.60).
 7           From the analysis of detailed deposition patterns measured by a serial bolus mouth delivery
 8      method, Kim and Hu (1998) and Kim and Jaques (2000) found a marked enhancement in
 9      deposition in the very shallow region (lung penetration depth <150 mL) of the lungs in females.
10      The enhanced local deposition for both ultrafme and coarse particles was attributed to a smaller
11      size of the upper airways, particularly of the laryngeal structure.
12           Hofmann et al. (2000) examined the role of heterogeneity of airway structure in the rat
13      acinar region in affecting deposition patterns within this area of the lungs. By the use of different
14      morphometric models, they showed that substantial variability in predicted particle deposition
15      would result.
16
17      6.2.4  Interspecies Patterns of Deposition
18           The primary purpose of this document is to assess the health effects of particles in humans.
19      As such, human dosimetry studies have been stressed.  Such studies avoid uncertainties
20      associated with extrapolation of dosimetry from laboratory animals to humans. Nevertheless,
21      animal models have been and are currently being used in evaluations of health effects from
22      particulate matter because there are ethical limits to the types of studies that  can be performed on
23      human subjects. Because of this, there is considerable need to understand dosimetry in animals
24      and to understand dosimetric differences between animals and humans. In this regard, there are a
25      number of newly published studies that were designed to assess particle dosimetry in commonly
26      used animals and to relate this to dosimetry in humans.
27           The various species used in inhalation toxicology studies that serve as  the basis for
28      dose-response assessment may not receive identical doses in a comparable respiratory tract
29      region (i.e., ET, TB, or A) when exposed to the same aerosol at the same inhaled concentration.
30      Such interspecies differences are important because any toxic effect is often  related to the
31      quantitative pattern of deposition within the respiratory tract as well as to the exposure
        April 2002                               6-32        DRAFT-DO NOT QUOTE OR CITE

-------
 1      concentration; this pattern determines not only the initial respiratory tract tissue dose, but also the
 2      specific pathways by which deposited material is cleared and redistributed (Schlesinger, 1985).
 3      Differences in patterns of deposition between humans and animals were summarized previously
 4      in the 1996 PM AQCD (U.S. Environmental Protection Agency, 1996) and by others
 5      (Schlesinger et al., 1997). Such differences in initial deposition must be considered when
 6      relating biological responses obtained in laboratory animal studies to effects in humans.
 7           It is difficult to compare systematically interspecies deposition patterns obtained from
 8      various reported studies because of variations in experimental protocols, measurement
 9      techniques, definitions of specific respiratory tract regions, and so on. For example, tests with
10      humans are generally conducted under protocols that standardize the breathing pattern; whereas
11      those using laboratory animals involve a wider variation in respiratory exposure conditions (e.g.,
12      spontaneous breathing versus ventilated breathing or varying degrees of sedation).  Much of the
13      variability in the reported data for individual species may be due to the lack of normalization for
14      specific respiratory parameters during exposure. In addition, the various studies have used
15      different exposure techniques, such as nasal mask, oral mask, oral tube, or tracheal intubation.
16      Regional deposition is affected by the exposure route and delivery technique employed.
17           Figure  6-9 shows the regional deposition data versus particle diameter in commonly used
18      laboratory animals obtained by various investigators, as compiled by Schlesinger (1988; 1989).
19      The results are described in detail in the 1996 PM AQCD (U.S. Environmental Protection
20      Agency, 1996).  In general, there is much variability in the data; however, it is possible to make
21      some generalizations concerning comparative deposition patterns. The relationship between total
22      respiratory tract deposition and particle size is approximately the same in humans and most of
23      these animals; deposition increases on both sides of a minimum that occurs for particles of 0.2 to
24      1 //m. Interspecies differences in regional deposition occur due to anatomical  and physiological
25      factors. In most laboratory animal species, deposition in the ET region is near 100  percent for a
26      particle diameter (dp) greater than 5 //m (Raabe et al., 1988), indicating greater efficiency than
27      that seen in humans. In the TB region, there is a relatively constant, but lower, deposition
28      fraction for dp greater than  1 //m in all species compared to humans. Finally, in the A region,
29      deposition fraction peaks at a lower particle size (dp about 1 //m) in laboratory animals than in
30      humans.
31

        April 2002                                6-33         DRAFT-DO NOT QUOTE OR CITE

-------
IUU


80
60

40

20
n
I I I IIW^
O Rat II F
n Hamster  T
'k ^' $
fjf 1 V
u ™, . :
I | I T ^BAAsjari

0.01 0.1 1.0 1C
                                  Particle Diameter (|jm)
Figure 6-9.  Regional deposition fraction in laboratory animals as a function of particle
            size. Particle diameters are aerodynamic (MMAD) for those > 0.5 ^m and
            geometric (or diffusion equivalent) for those < 0.5 (j,m.

Source:  Schlesinger (1988).
April 2002
   6-34
     DRAFT-DO NOT QUOTE OR CITE

-------
 1           One of the issues that must be considered in interspecies comparisons of hazards from
 2      inhaled particles is inhalability of the aerosol in the atmosphere of concern. Although this may
 3      not be an issue for humans per se as far as exposure to ambient particles is concerned, it can be
 4      an important issue when attempting to extrapolate to humans the results of studies using animal
 5      species commonly employed in inhalation toxicological studies (Miller et al., 1995).
 6      For example, differences between rat and human become very pronounced for particles >5 //m,
 7      and some differences are also evident for particles as small as 1 //m (Figure 6-10).
 8           A number of studies have addressed various aspects of interspecies differences in
 9      respiratory tract deposition using mathematical modeling approaches. Hofmann et al. (1996)
10      compared deposition between rat and human lungs, using three-dimensional asymmetric
11      bifurcation models and mathematical procedures for obtaining air flow and particle trajectories.
12      Deposition in segmental bronchi and terminal bronchioles was evaluated under both inspiration
13      and expiration at particle sizes of 0.01, 1.0, and 10 //m, which covers the range of deposition
14      mechanisms from diffusion to impaction. Total deposition efficiencies of all particles in the
15      upper and lower airway bifurcations were comparable in magnitude for both rat and human.
16      However, the investigators noted that penetration probabilities from preceding airways must be
17      considered. When considering the higher penetration probability in the human lung, the resulting
18      bronchial deposition fractions were generally higher in human than in rat.  For all particle sizes,
19      deposition at rat bronchial bifurcations was less enhanced on the carinas compared to that found
20      in human airways.
21           Hofmann et al. (1996) attempted to account for interspecies differences in branching
22      patterns in deposition analyses. Numerical simulations of three-dimensional particle deposition
23      patterns within selected (species-specific) bronchial bifurcations indicated that morphologic
24      asymmetry was a major determinant of the heterogeneity of local deposition patterns. They noted
25      that many interspecies deposition calculations used morphometry that was described by
26      deterministic lung models (i.e., the number of airways in each airway generation is constant, and
27      all airways in a given generation have identical lengths and diameters). Such models cannot
28      account for variability and branching asymmetry of airways in the lungs.  Thus, their study
29      employed computations that used stochastic morphometric models of human and rat lungs
30      (Koblinger and Hofmann, 1985, 1988; Hofmann et al., 1989b) and evaluated regional and local
31      particle deposition.  Stochastic models of lung structure describe, in mathematical terms, the

        April 2002                                6-3 5        DRAFT-DO NOT QUOTE OR CITE

-------
                 100
Total Respiratory Tract
  Human
-  Oral Breathing
                   0
                 100
  Human
-  Nasal Breathing
                 100 r  Rat
                   0.01      0.1       1.0
                        Particle Diameter (|jm)
                    Tracheobronchial Region
                      Human
                      Oral Breathing
                                            10
               01
               Q
                 100 r
                     Rat
                   0.01      0.1       1.0
                        Particle Diameter (|jm)
                                            10
              B
   Extrathoracic Region
100r Human
   . Oral Breathing
                                                   .-.  100
                                                   CD
                                                   Q
                                                          Human
                                                         - Nasal Breathing
                                                      100 r Rat
                                    .01       0.1      1.0
                                         Particle Diameter (|jm)


                                                D

                                     Alveolar Region
                                      Human
                                     . Oral Breathing
                                                       0
                                                      100
                                CD
                                Q
                                                          Human
                                                         - Nasal Breathing
                                                      100 r Rat
                                   0.01      0.1       1.0
                                        Particle Diameter (^m)
 Figure 6-10.  Particle deposition efficiency in rats and humans as a function of particle
                size for the (A) total respiratory tract, (B) thoracic region,
                (C) tracheobronchial region, and (D)  alveolar region. Each curve
                represents an eye fit through mean values (or centers of ranges) for the
                data compiled by Schlesinger (1985).  Particle diameters are aerodynamic
                (MMAD) for those >0.5 ^,m and geometric (or diffusion equivalent) for
                those < 0.5 (j,m.

 Source: Modified from Schlesinger (1989).
April 2002
                          6-36
      DRAFT-DO NOT QUOTE OR CITE

-------
 1      inherent asymmetry and variability of the airway system, including diameter, length, and angle.
 2      They are based on statistical analyses of actual morphometric analyses of lungs. The model also
 3      incorporated breathing patterns for humans and rats.  In a later analysis (Hoffmann and
 4      Bergmann, 1998), the dependence of deposition on particle size was found to be qualitatively
 5      similar in both rats and humans, with deposition minima in the size range of 0.1 to 1 //m for total
 6      deposition as well as deposition within the TB and A regions. In addition, a deposition
 7      maximum occurred at about 0.02 to 0.03 //m and between 3 and 5 //m in both species. The
 8      deposition decrease in the A region at the smallest and largest sizes resulted from the filtering
 9      efficiency of upstream airways. Although deposition patterns were qualitatively similar in rat
10      and human, deposition in the human lung appeared to be consistently  higher than in the rat in all
11      regions of the lung (TB and A) over the entire size range. Both species showed a similar pattern
12      of dependence of deposition on flow rate.
13           The above model also assessed local deposition. In both human and rat, deposition of
14      0.001-//m particles was highest in the upper bronchial airways; whereas 0.1- and l-//m particles
15      showed higher deposition in more peripheral airways, namely the bronchiolar airways  in rat and
16      the respiratory bronchioles in humans. Deposition was variable within any branching generation
17      because of differences in airway dimensions, and regional and total deposition also exhibited
18      intrasubject variations. Airway geometric differences between rats and humans were reflected in
19      deposition.  Because of the greater branching asymmetry in rats, prior to about generation 12,
20      each generation showed deposition maxima at two particle sizes, reflecting deposition in major
21      and minor daughters.  These geometric differences became reduced with depth into the lung;
22      beyond generation 12, these two maxima were no longer seen.
23           Another comparison of deposition in lungs of humans and rats was performed by Musante
24      and Martonen (2000b).  An interspecies mathematical dosimetry model was used to determine
25      the deposition of ROFA in the lungs under sedentary and light activity breathing patterns. This
26      latter condition was mimicked in the rat by increasing the CO2 level in the exposure system. The
27      MMAD of the particle size distribution was 1.95 //m with a geometric standard deviation of 2.19.
28      They noted that physiologically comparable respiratory intensity levels did not necessarily
29      correspond to comparable dose distribution in the lungs. Because of this, the investigators
30      speculate that the resting rat may not be a good model for the resting human. The ratio of aerosol
31      mass deposited in the TB region to that in the A region for the human at rest was 0.961,

        April 2002                                6-37        DRAFT-DO NOT QUOTE OR CITE

-------
 1      indicating fairly uniform deposition throughout the lungs.  On the other hand, in the resting rat,
 2      the ratio was 2.24, indicating greater deposition in the TB region than in the A region. However,
 3      by mimicking light activity in the rat, the ratio was reduced to 0.97, similar to the human. These
 4      data suggest that ventilatory characteristics in animal models may have to be adjusted to provide
 5      for comparable regional deposition to that in humans.
 6           The relative distribution of particles deposited within the bronchial and alveolar regions of
 7      the airways may differ in the lungs of animals and humans for the same total amount  of deposited
 8      matter because of structural differences.  The effect of such structural differences between rat and
 9      human airways on particle deposition patterns was examined by Hofmann et al. (1999; 2000) in
10      an attempt to find the most appropriate morphometric parameter to characterize local particle
11      deposition for extrapolation modeling purposes. Particle deposition patterns were evaluated as
12      functions of three morphometric parameters, namely (1) airway generation, (2) airway diameter,
13      and (3) cumulative path length.  It was noted that airway diameter was a more appropriate
14      morphometric parameter for comparison of particle deposition patterns in human and rat lungs
15      than was airway generation.
16           The manner in which particle dose is expressed, that is, the specific dose metric, may affect
17      relative differences in deposition between humans and other animal species. For example,
18      although deposition when expressed on a mass per unit alveolar surface area basis may not be
19      different between rats and humans, dose metrics based on particle number per various anatomical
20      parameters (e.g., per alveolus or alveolar macrophage) can differ between rats and humans,
21      especially for particles around 0.1 to 0.3 //m (Miller et al., 1995). Furthermore, in humans with
22      lung disease (such as asthma or COPD), differences between rat and human can be even more
23      pronounced.
24           The probability of any biological effect occurring in humans or animals depends on
25      deposition and retention of particles, as well as the underlying tissue sensitivity.  Interspecies
26      dosimetric extrapolation must consider these differences in evaluating dose-response
27      relationships.  Thus, even similar deposition patterns may not result in similar effects in different
28      species,  because dose also is affected by clearance mechanisms. In addition, the  total number of
29      particles deposited in the lung may not be the most relevant dose metric for interspecies
30      comparisons.  For example, it may be the number of deposited particles per unit surface area or
31      dose to a specific cell (e.g., alveolar macrophage) that determines response for specific regions.

        April 2002                                6-3 8        DRAFT-DO NOT QUOTE OR CITE

-------
 1      More specifically, even if deposition is similar in rat and human, there would be a higher
 2      deposition density in the rat because of the smaller surface area of rat lung. Thus, species-
 3      specific differences in deposition density should be considered when health effects observed in
 4      laboratory animals are being evaluated for potential effects occurring in humans.
 5
 6
 7      6.3 PARTICLE CLEARANCE AND TRANSLOCATION
 8           This section discusses the clearance and translocation of particles that have deposited in the
 9      respiratory tract.  First, a basic overview of biological mechanisms and pathways of clearance in
10      the various region of the respiratory tract is presented. This is then followed by an update on
11      regional kinetics of particle clearance. Interspecies patterns of clearance are then addressed,
12      followed by new information on biological factors that may modulate clearance.
13
14      6.3.1  Mechanisms and Pathways of Clearance
15           Particles that deposit on airway surfaces may be cleared from the respiratory tract
16      completely or may be translocated to other sites within this system by various regionally distinct
17      processes. These clearance mechanisms, which are outlined in Table 6-1, can be categorized as
18      either absorptive (i.e., dissolution) or nonabsorptive (i.e., transport of intact particles) and may
19      occur simultaneously or with temporal variations.  It should be mentioned that particle solubility
20      in terms of clearance refers to solubility within the respiratory tract fluids and cells. Thus, a
21      poorly soluble particle is considered to be one whose rate of clearance by dissolution is
22      insignificant compared to its rate of clearance as an intact particle. All deposited particles,
23      therefore, are subject to clearance by the same basic mechanisms, with their ultimate fate a
24      function of deposition site, physicochemical properties (including solubility and any toxicity),
25      and sometimes deposited mass or number concentration. Clearance routes from the various
26      regions of the respiratory tract have been discussed previously in detail (U.S. Environmental
27      Protection Agency, 1996; Schlesinger et al., 1997).  They are  schematically shown in Figure 6-11
28      (for extrathoracic and tracheobronchial regions) and in Figure 6-12 (for poorly soluble particle
29      clearance from the alveolar region) and are reviewed only briefly below.
30

        April 2002                                6-39        DRAFT-DO NOT QUOTE OR CITE

-------
   TABLE 6-1. OVERVIEW OF RESPIRATORY TRACT PARTICLE CLEARANCE
                         AND TRANSLOCATION MECHANISMS

 Extrathoracic region (ET)
    Mucociliary transport
    Sneezing
    Nose wiping and blowing
    Dissolution and absorption into blood

 Tracheobronchial region (TB)
    Mucociliary transport
    Endocytosis by macrophages/epithelial cells
    Coughing
    Dissolution and absorption into blood/lymph

 Alveolar region (A)
    Macrophages, epithelial cells
    Interstitial
    Dissolution and absorption into blood/lymph	

 Source:  Schlesinger (1995).
                                      (  Nasal Passages  j^
)                             Dissolution /*
                            *—c
                                                                  Mucociliary
                                                                  Transport
) Dissolution
-*	(   Tracheobronchial Tree
Figure 6-11.  Major clearance pathways for particles deposited in the extrathoracic region
              and tracheobronchial tree.

Source: Adapted from Schlesinger et al. (1997).
April 2002
                 6-40
DRAFT-DO NOT QUOTE OR CITE

-------
uefjusn&u i-cHiiut;
P
i
Phagocytosis by "\
Alveolar Macrophages J
1 ^
Movement within . 	 niaSS,a~
Alveolar Lumen " 	 Alveola
|
Bronchiolar/ Bronchial < 	 Inte
Endocyt
Epithelij
je Through
r Epithelium
i
1
rstitium "^
osis I
Iveol
alCe

1 ^ ^_
~ Lymphatic Channels "^
Mucociliary Blanket i
-
^y
ar
^ t
Passage through
Pulmonary Capillary
Endothelium
— /" Phagocytosis byA
- 1 Interstitial 1
V^ Macrophages J


                        Gl Tract
           Figure 6-12.  Diagram of known and suspected clearance pathways for poorly
                        soluble particles depositing in the alveolar region.  (The magnitude of
                        various pathways may depend upon size of deposited particle.)
           Source:  Modified from Schlesinger et al. (1997).
 1      6.3.1.1  Extrathoracic Region
 2           The clearance of poorly soluble particles deposited in the posterior portions of the nasal
 3      passages occurs via mucociliary transport, with the general flow of mucus being towards the
 4      nasopharynx.  Mucus flow in the most anterior portion of the nasal passages is forward, clearing
 5      deposited particles to the vestibular region, where removal occurs by sneezing, wiping, or
 6      blowing.  Soluble material deposited on the nasal epithelium is accessible to underlying cells via
 7      diffusion through the mucus.  Dissolved substances may be translocated subsequently into the
 8      bloodstream.  The nasal passages have a rich vasculature, and uptake into the blood from this
 9      region may occur rapidly.
10           Clearance of poorly soluble particles deposited in the oral passages is by coughing and
11      expectoration or by swallowing into the gastrointestinal tract. Soluble particles are likely to be
12      rapidly absorbed after deposition, but it depends on the rate of dissolution of the particle and the
13      molecular size of the solute.
       April 2002
6-41
DRAFT-DO NOT QUOTE OR CITE

-------
 1      6.3.1.2 Tracheobronchial Region
 2           Poorly soluble particles deposited within the TB region are cleared by mucociliary transport
 3      towards the oropharynx, followed by swallowing. Poorly soluble particles also may traverse the
 4      epithelium by endocytotic processes, entering the peribronchial region, be engulfed via
 5      phagocytosis by airway macrophages (which can then move cephalad on the mucociliary
 6      blanket), or enter the airway lumen from the bronchial or bronchiolar mucosa.  Soluble particles
 7      may be absorbed through the epithelium into the blood.  It has been shown that blood flow
 8      affects translocation from the TB region, in that decreased bronchial blood flow is associated
 9      with increased airway retention of soluble particles (Wagner and Foster, 2001). There is,
10      however, evidence that even soluble particles may be cleared by mucociliary transport (Bennett
11      and Howite, 1989; Matsui et al., 1998; Wagner and Foster, 2001).
12
13      6.3.1.3 Alveolar Region
14           Clearance from the A region occurs via a number of mechanisms and pathways. Particle
15      removal by macrophages comprises the main nonabsorptive clearance process  in this region.
16      These cells, which reside on the epithelium, phagocytize and transport deposited material that
17      they contact by random motion or via directed migration under the influence of chemotactic
18      factors.
19           Although alveolar macrophages normally comprise up to about 5% of the total alveolar
20      cells in healthy, nonsmoking humans and other mammals, the actual cell count may be altered by
21      particle loading.  The magnitude of any increase in cell number is related to the number of
22      deposited particles rather than to total deposition by weight. Thus, equivalent masses of an
23      identically deposited substance would not produce the same response if particle sizes differed,
24      and the deposition of smaller particles would tend to result in a greater elevation in macrophage
25      number than would deposition of larger particles.
26           Particle-laden macrophages may be cleared  from the A region along a number of pathways.
27      As noted in Figure 6-11, this includes cephalad transport via the mucociliary system after the
28      cells reach the distal terminus of the mucus blanket; movement within the interstitium to a
29      lymphatic channel; or perhaps traversing of the alveolar-capillary endothelium, directly entering
30      the bloodstream.  Particles within the lymphatic system may be translocated to tracheobronchial
31      lymph nodes, which can become reservoirs of retained material.  Particles subsequently reaching

        April 2002                                6-42        DRAFT-DO NOT QUOTE OR CITE

-------
 1      the postnodal lymphatic circulation will enter the blood.  Once in the systemic circulation, these
 2      particles, or transmigrated macrophages, can travel to extrapulmonary organs. Deposited
 3      particles that are not ingested by alveolar macrophages may enter the interstitium, where they are
 4      subject to phagocytosis by resident interstitial macrophages, and may travel to perivenous,
 5      peribronchiolar or subpleural sites, where they become trapped, increasing particle burden. The
 6      migration and grouping of particles and macrophages within the lungs can lead to the
 7      redistribution of initially diffuse deposits into focal aggregates.  Some particles or components
 8      can bind to epithelial cell membranes or macromolecules, or to other cell components, delaying
 9      clearance from the lungs.
10           Churg and Brauer (1997) examined lung autopsy tissue from 10 never-smokers from
11      Vancouver, Canada. They noted that the geometric mean particle diameter (GMPD) in lung
12      parenchymal tissue was 0.38 //m (og = 2.4).  Ultrafme particles accounted for less than 5% of the
13      total retained particulate  matter.  Metal particles had a GMPD of 0.17 //m, and  silicates 0.49 //m.
14      Ninety-six percent of retained PM was less than 2.5 //m.  A subsequent study considered
15      retention of actual ambient particles in the lungs, which is related to deposition. Brauer et al.
16      (2001) showed that small particles could undergo significant steady-state retention within the
17      lungs. Using lungs obtained at autopsy from long-term, nonsmoking residents of an area having
18      high levels of ambient PM (Mexico City, Mexico) and those from an area with relatively low PM
19      levels (Vancouver, Canada), the investigators measured the particle concentration per gram of
20      lung within the parenchyma. They found that living in the high PM region resulted in
21      significantly greater retention of both fine and ultrafine particles within the lungs; levels in the
22      lungs from Mexico City contained over 7.4 times the concentration of these particles as did the
23      lungs from residents of Vancouver. These results indicate a clear relationship between ambient
24      exposure concentration and  retention in the A region.
25           Clearance by the absorptive mechanism involves dissolution in the alveolar surface fluid,
26      followed by transport through the epithelium and into the interstitium, and then diffusion into the
27      lymph or blood. Solubility is influenced by the particle's surface to volume ratio and other
28      properties, such as hydrophilicity and lipophilicity (Mercer, 1967; Morrow, 1973; Patton, 1996).
29
30
31

        April 2002                                6-43        DRAFT-DO NOT QUOTE OR CITE

-------
 1      6.3.2 Clearance Kinetics
 2           The kinetics of clearance have been reviewed in U.S. Environmental Protection Agency
 3      (1996) and in a number of monographs (e.g., Schlesinger et al., 1997) and are discussed only
 4      briefly here.  The actual time frame over which clearance occurs affects the cumulative dose
 5      delivered to the respiratory tract, as well as the dose delivered to extrapulmonary organs.
 6
 7      6.3.2.1  Extrathoracic Region
 8           Mucus flow rates in the posterior nasal passages are highly nonuniform, but the median rate
 9      in a healthy adult human is about 5 mm/min, resulting in a mean anterior to posterior transport
10      time of about 10 to 20 min for poorly soluble particles (Rutland and Cole, 1981; Stanley et al.,
11      1985). Particles deposited in the anterior portion of the nasal passages are cleared more slowly
12      by mucus transport and are usually more effectively removed by sneezing, wiping, or nose
13      blowing (Fry and Black, 1973; Morrow, 1977).
14
15      6.3.2.2  Tracheobronchial Region
16           Mucus transport in the tracheobronchial tree occurs at different rates in different local
17      regions; the velocity of movement is fastest in the trachea, and it becomes progressively slower
18      in more distal airways. In healthy nonsmoking humans, using noninvasive procedures and no
19      anesthesia, average tracheal mucus transport rates have been measured at 4.3 to 5.7 mm/min
20      (Yeates et al., 1975, 1981; Foster et al., 1980; Leikauf et al., 1981, 1984); whereas that in the
21      main bronchi has been measured at -2.4 mm/min (Foster et al., 1980). Estimates for human
22      medium bronchi range between  0.2 to 1.3 mm/min; whereas those in the most distal ciliated
23      airways range down to 0.001 mm/min (Morrow et al., 1967; Cuddihy and Yeh, 1988; Yeates and
24      Aspin, 1978).
25           The total duration of bronchial clearance or some other time parameter often is used as an
26      index of mucociliary kinetics. Although clearance from the TB region is generally rapid, there is
27      experimental evidence, discussed in U.S. Environmental Protection Agency (1996), that a
28      fraction of material deposited in the TB region is retained much longer than the 24 h commonly
29      used as the outer range of clearance time for particles within this region (Stahlhofen et al.,
30      1986a,b; Scheuch and Stahlhofen, 1988; Smaldone et al., 1988). A study by  Asgharian et al.
31      (2001) showed that it is not necessary to invoke a slow- and fast-phase for TB clearance to have
        April 2002                                6-44        DRAFT-DO NOT QUOTE OR CITE

-------
 1      particles retained longer than 24 h. Based upon asymmetric stochastic human lung modeling
 2      data, intersubject variability in retained mass arising from the periphery of the TB can explain the
 3      experimental observations while still fitting a single compartment clearance model. Other
 4      studies described below, however, do support the concept that TB regional clearance consists of
 5      both a fast and a slow component.
 6           Falk et al. (1997) studied clearance in healthy adults using monodisperse Teflon particles
 7      (6.2 //m) inhaled at two flow rates. A considerable fraction (about 50%) of particles deposited in
 8      small airways had not cleared within 24 h following exposure.  These particles cleared with a
 9      half time of 50 days. Although the deposition sites of the particles were not confirmed
10      experimentally, calculations suggested these to be in the smaller ciliated airways. Camner et al.
11      (1997) also noted that clearance from the TB region was incomplete by 24 h postexposure and
12      suggested that this may be caused by incomplete clearance  from bronchioles.  Healthy adults
13      inhaled teflon particles (6, 8, and 10 //m) under low flow rates to maximize deposition in the
14      small ciliated  airways.  The investigators noted a decrease in 24-h retention with increasing
15      particle size, indicating a shift toward either a smaller retained fraction, deposition more
16      proximally in  the respiratory tract, or both.  They calculated that a large fraction, perhaps as high
17      as 75% of particles depositing in generations 12 through 16, was still retained at 24 h
18      postexposure.
19           In a study to examine retention kinetics in the tracheobronchial tree  (Falk et al., 1999),
20      nonsmoking healthy adults inhaled radioactively tagged 6. l-//m particles at both a normal flow
21      rate and a slow flow rate designed to deposit particles preferentially in small ciliated airways.
22      Lung retention was measured from 24 h to 6 mo after exposure. Following normal flow rate
23      inhalation, 14% of the particles retained at 24 h cleared with a half time of 3.7 days and 86%
24      with a half time of 217 days. Following slow flow rate inhalation, 35% of the particles retained
25      at 24 h cleared with a half time of 3.6 days and 65% with a half time  of 170 days. Estimates
26      using a number of mathematical models indicated higher deposition in the bronchiolar region
27      (generations 9 through 15) with the slow rate inhalation compared to the normal rate.  The
28      experimental data and predictions of the deposition modeling indicated that 40% of the particles
29      deposited in the conducting  airways during the slow inhalation were retained after 24 h. The
30      particles that cleared with the shorter half time were mainly deposited in the bronchi olar region,


        April 2002                                6-45        DRAFT-DO NOT QUOTE OR CITE

-------
 1      but only about 25% of the particles deposited in this region cleared in this phase. This study
 2      provided additional confirmation for a phase of slow clearance from the bronchial tree.
 3           The underlying sites and mechanisms of long-term TB retention in the smaller airways are
 4      not known. Some proposals were presented in the earlier 1996 PM AQCD (U.S. Environmental
 5      Protection Agency, 1996).  This slow clearing tracheobronchial compartment likely is associated
 6      with bronchioles <1  mm in diameter (Lay et al., 1995; Kreyling et al., 1999; Falk et al., 1999).
 7      Based on a study in which an adrenergic agonist was used to stimulate mucus flow, so as to
 8      examine the role of mucociliary transport in the bronchioles, it was found that clearance from the
 9      smaller airways was not influenced by the drug, suggesting to the investigators that mucociliary
10      transport was not as  an effective clearance mechanism from this region as it is in larger airways
11      (Svartengren et  al., 1998, 1999). Although slower or less effective mucus transport may result in
12      longer retention times in small airways, other factors may account for long-term TB retention.
13      One such proposal is the movement of particles into the gel phase because of surface tension
14      forces in the liquid lining of the small airways (Gehr et al., 1990,  1991).  The issue of particle
15      retention in the tracheobronchial tree certainly is not resolved.
16           Long-term TB  retention patterns are not uniform.  There is an enhancement at bifurcation
17      regions (Radford and Martell, 1977; Henshaw and Fews, 1984; Cohen et al., 1988), the likely
18      result of both greater deposition and less effective mucus clearance within these areas.  Thus,
19      doses calculated based on uniform surface retention density may be misleading, especially if the
20      material is lexicologically slow acting.
21
22      6.3.2.3 Alveolar Region
23           Particles deposited in the A region generally are retained longer than are those deposited in
24      airways cleared by mucociliary transport. There are limited data on alveolar clearance rates in
25      humans. Within any species, reported clearance rates vary widely because, in part, of different
26      properties of the particles used in the various studies. Furthermore, some chronic experimental
27      studies have employed high concentrations of poorly soluble particles that may have interfered
28      with normal clearance mechanisms, resulting in clearance rates different from those that would
29      typically occur at lower exposure levels.  Prolonged exposure to high particle concentrations is
30      associated with  what is termed particle "overload."  This is discussed in greater detail in
31      Section 6.4.

        April 2002                                6-46        DRAFT-DO NOT QUOTE OR CITE

-------
 1           There are numerous pathways of A region clearance, and the utilization of these may
 2      depend on the nature of the particles being cleared. Little is known concerning relative rates
 3      along specific pathways.  Thus, generalizations about clearance kinetics are difficult to make.
 4      Nevertheless, A region clearance is usually described as a multiphasic process, each phase
 5      considered to represent removal by a different mechanism or pathway and often characterized by
 6      increased retention half times following toxicant exposure.
 7           The initial uptake of deposited particles by alveolar macrophages is very rapid and
 8      generally occurs within 24 h of deposition (Lehnert and Morrow, 1985; Naumann and
 9      Schlesinger, 1986; Lay et al., 1998).  The time for clearance of particle-laden alveolar
10      macrophages via the mucociliary system depends on the site of uptake relative to the distal
11      terminus of the mucus blanket at the bronchiolar level. Furthermore, clearance pathways and
12      subsequent kinetics may  depend to some extent on particle size. For example, some smaller
13      ultrafine particles (< 0.02 //m) may be less effectively phagocytosed than larger ones
14      (Oberdorster, 1993).
15           Uningested particles may penetrate into the interstitium within a few hours following
16      deposition. This transepithelial passage seems to increase as particle loading increases,
17      especially to that level above which macrophage numbers increase (Ferin, 1977; Ferin et al.,
18      1992; Adamson and Bowden, 1981). It also may be particle size dependent, because insoluble
19      ultrafine particles (<0.1 //m diameter) of low intrinsic toxicity show increased access to the
20      interstitum and greater lymphatic uptake than do larger particles of the same material
21      (Oberdorster et al., 1992; Ferin et al., 1992). However, ultrafine particles of different materials
22      may not enter the interstitium to the same extent.  Similarly, a depression of phagocytic activity,
23      a reduction in macrophage ability to migrate to sites of deposition (Madl et al., 1998), or the
24      deposition of large numbers of ultrafine particles may increase the number of free particles in the
25      alveoli, perhaps enhancing removal by other routes.  In any case, free particles may reach the
26      lymph nodes perhaps within a few days after deposition (Lehnert et al., 1988; Harmsen et al.,
27      1985) although this route is not definitive and may be species dependent.
28           The extent of lymphatic uptake of particles may depend on the effectiveness of other
29      clearance pathways, in that lymphatic translocation likely increases when phagocytic activity of
30      alveolar macrophages decreases.  This may be a factor in lung overload. However,  it seems that
31      the deposited mass or number of particles must exceed some threshold below which increases in

        April 2002                                6-47        DRAFT-DO NOT QUOTE OR CITE

-------
 1      loading do not affect translocation rate to the lymph nodes (Ferin and Feldstein, 1978; LaBelle
 2      and Brieger, 1961).  In addition, the rate of translocation to the lymphatic system may be
 3      somewhat particle size dependent. Although no human data are available, translocation of latex
 4      particles to the lymph nodes of rats was greater for 0.5- to 2-//m particles than for 5- and 9-//m
 5      particles (Takahashi et al., 1992), and particles within the 3- to 15-//m size range were found to
 6      be translocated at faster rates than were larger sizes (Snipes and Clem, 1981). On the other hand,
 7      translocation to the lymph nodes was similar for both 0.4-//m barium sulfate or 0.02-//m gold
 8      colloid particles (Takahashi et al., 1987). It seems that particles <2 //m clear to the lymphatic
 9      system at a rate independent of size; and it is particles of this size, rather than those >5 //m, that
10      would have significant deposition within the A region following inhalation.  In any case, the
11      normal rate of translocation to the lymphatic system is quite slow; and elimination from the
12      lymph nodes is even slower, with half times estimated in tens of years (Roy, 1989).
13           Soluble particles depositing in the A region may be cleared rapidly via absorption through
14      the epithelial surface into the blood. Actual rates depend on the size of the particle (i.e.,  solute
15      size), with smaller molecular weight solutes clearing faster than larger ones. Absorption may be
16      considered as a two stage process, with the first stage being dissociation of the deposited
17      particles into material that can be absorbed into the circulation (i.e., dissolution) and the  second
18      stage being uptake of this material. Each of these stages may be time dependent. The rate of
19      dissolution depends on a number  of factors, including particle surface area and chemical
20      structure. A portion of the dissolved material may be absorbed more slowly because  of binding
21      to respiratory tract components. Accordingly, there is a very wide range for absorption rates,
22      depending on the physicochemical properties of the material deposited.
23           As indicated in both the toxicology and epidemiology chapters of this document (Chapters
24      7 and 8), one of the health outcome of concern relates to ambient PM effects on the
25      cardiovascular system.  Thus, an important dosimetric issue involves the pathways by which
26      inhaled and deposited particles in the lungs could impact upon extrapulmonary systems.
27      Clearance and translocation pathways by which this may occur have  been recently described.
28      Nemmar et al. (2001) instilled hamsters with radioactively-labeled colloidal albumin  particles
29      (diameter < 0.080 //m) as a model for ambient ultrafine particles and measured the label
30      appearing in systemic blood and various extrapulmonary organs up to 1 h postexposure.  They
31      found label in blood within 5 minutes after instillation.  In their subsequent studies in which

        April 2002                                 6-48        DRAFT-DO NOT QUOTE OR CITE

-------
 1      healthy volunteers were challenged with inhalation of 99mTechnitum-labeled ultrafine carbon
 2      particles (Nemmar et al., 2002), the radioactivity was detected in blood as early as 1  min,
 3      reaching a maximum between 10 and 20 min after inhalation of the aerosol. While label was
 4      also noted in the other extrapulmonary organs examined (namely liver, heart, spleen, kidneys,
 5      and brain), the liver had the highest levels and these increased with increasing time postexposure,
 6      while the second highest levels were noted in the heart or kidney, depending upon the instilled
 7      concentration. This suggests that ultrafine particles can rapidly diffuse from the lungs into the
 8      systemic circulation, thus providing a pathway by which ambient PM may rapidly affect
 9      extrapulmonary organs.
10           In another study, Takenaka et al. (2001) exposed rats by inhalation to 0.015 //m particles of
11      elemental  silver and found evaluated levels of these particles in various extrapulmonary organs
12      up to 7 days postexposure. They found that the amount of particles in the lungs decreased
13      rapidly with time and, by day 7, only about 4% of the initial lung burden remained. At day 0,
14      particles were already found in the blood.  The particles were found to be distributed in the liver,
15      kidney, heart, and brain by 1 day postexposure.  The particle concentration was highest in the
16      kidney, followed by the liver, and then the heart. This study also indicates that inhaled ultrafine
17      particles were rapidly cleared from the lungs into the systemic circulation.  However, a similar
18      cleance pattern was found after intratracheal instillation of AgNO3 solution. Therefore,  the
19      investigators postulated that the rapid clearance of elemental silver particles was due to a fast
20      dissolution of ultrafine silver particles into the lung fluid and subsequent diffusion into the blood
21      stream, although a possibility of direct translocation of solid particles into the blood  stream was
22      not excluded.  The investigators also instilled an aqueous suspension of elemental silver into
23      some animals; in this case, there was more retention in the lungs, which was ascribed to
24      phagocytic accumulation of agglomerated particles in alveolar macrophages and slow dissolution
25      of particles in cells.  Thus, this study also  suggested that particle size and the tendency of
26      particles to aggregate can affect the translocation pathway from the lungs. Earlier studies
27      (Huchon et al., 1987; Peterson et al., 1989; Morrison et al., 1998) investigated lung clearance of
28      labeled macromolecule solutes with widely varying molecular weight and labeled albumin, as
29      well as albumin ultrafine aggregates.  Clearance rates found from these earlier studies were much
30      slower than recent studies described above, suggesting that the possibility of a fast clearing
31      pathway of solid ultrafine particles may need further study.

        April 2002                                6-49        DRAFT-DO NOT QUOTE OR CITE

-------
 l      6.3.3 Interspecies Patterns of Clearance
 2           The inability to study the retention of certain materials in humans for direct risk assessment
 3      requires use of laboratory animals. Because dosimetry depends on clearance rates and routes,
 4      adequate toxicologic assessment necessitates that clearance kinetics in such animals be related to
 5      those in humans. The basic mechanisms and overall patterns of clearance from the  respiratory
 6      tract are similar in humans and most other mammals. However, regional clearance  rates can
 7      show substantial variation between species, even for similar particles deposited under
 8      comparable exposure conditions, as extensively reviewed elsewhere (U.S. Environmental
 9      Protection Agency, 1996; Schlesinger et al., 1997; Snipes et al., 1989).
10           In general, there are species-dependent rate constants for various clearance pathways.
11      Differences in regional and total clearance rates between some species are a reflection of
12      differences in mechanical clearance processes. For example, the relative proportion of particles
13      cleared from the A region in the short- and longer-term phases differs between laboratory rodents
14      and larger mammals, with a greater percentage cleared in the faster phase in rodents.  A recent
15      study (Oberdorster et al., 1997) showed interstrain differences in mice and rats in the handling of
16      particles by alveolar macrophages. Macrophages of B6C3F1 mice could not phagocytize 10-//m
17      particles, but those of C57 black/61 mice did.  In addition, the nonphagocytized 10-//m particles
18      were efficiently eliminated from the alveolar region; whereas previous work in rats  found that
19      these large particles, after uptake by macrophages, were retained persistently (Snipes and Clem,
20      1981; Oberdorster et al.,  1992). The ultimate  implication of interspecies differences in clearance
21      needing to be considered in assessing particle  dosimetry is that the retention of deposited
22      particles can differ between species and may result in differences in response to similar PM
23      exposure atmospheres.
24           Hsieh and Yu (1998) summarized the existing data on pulmonary clearance of inhaled,
25      poorly soluble particles in the rat, mouse, guinea pig, dog, monkey, and human.  Clearance at
26      different initial lung burdens, ranging from 0.001 to 10 mg particles/g lung, was analyzed using a
27      two-phase exponential decay function. Two clearance phases in the alveolar region, namely fast
28      and slow, were associated with mechanical clearance along two pathways, the former with the
29      mucociliary system and the latter with the lymph nodes.  Rats and mice were noted  to be fast
30      clearers in comparison to the other species. Increasing the initial lung burden resulted in an
31      increasing mass fraction of particles cleared by the slower phase.  As lung burden increased
        April 2002                                6-50        DRAFT-DO NOT QUOTE OR CITE

-------
 1      beyond 1 mg particles/g lung, the fraction cleared by the slow phase increased to almost 100%
 2      for all species. However, the rate for the fast phase was similar in all species and did not change
 3      with increasing lung burden of particles; whereas the rate for the slow phase decreased with
 4      increasing lung burden. At elevated burdens, the effect on clearance rate was greater in rats than
 5      in humans, an observation consistent with previous findings (Snipes, 1989).
 6
 7      6.3.4 Factors Modulating Clearance
 8           A number of factors have previously been assessed in terms of modulation of normal
 9      clearance patterns, including:  age, gender, workload, disease, and irritant inhalation. Such
10      factors have been discussed in detail previously (U.S. Environmental Protection Agency, 1996).
11
12      6.3.4.1 Age
13           Studies previously described in the 1996 PM AQCD (U.S. Environmental Protection
14      Agency, 1996) indicated that there appeared to be no clear evidence for any age-related
15      differences in clearance from the lung or total respiratory tract, either from  child to adult, or
16      young adult to elderly.  Studies of mucociliary function have shown either no changes or some
17      slowing in mucous clearance function with age after maturity, but at a rate that would be unlikely
18      to significantly affect overall clearance kinetics.
19
20      6.3.4.2 Gender
21           Previously reviewed studies (U.S. Environmental Protection Agency, 1996) indicated no
22      gender-related differences in nasal mucociliary clearance rates in children (Passali and Bianchini
23      Ciampoli, 1985) nor in tracheal transport rates in adults (Yeates et al., 1975).
24
25      6.3.4.3 Physical Activity
26           The effect of increased physical activity on mucociliary clearance is unresolved, with
27      previously discussed studies (U.S. Environmental Protection Agency, 1996) indicating either no
28      effect or an increased clearance rate with exercise.  There are no data concerning changes in
29      A region clearance with increased activity levels. Breathing with an increased tidal volume was
30      noted to increase the rate of particle clearance from the A region, and this was suggested to result
31      from distension-related evacuation of surfactant into proximal airways, resulting in a facilitated
        April 2002                                6-51        DRAFT-DO NOT QUOTE OR CITE

-------
 1      movement of particle-laden macrophages or uningested particles because of the accelerated
 2      motion of the alveolar fluid film (John et al., 1994).
 3
 4      6.3.4.4 Respiratory Tract Disease
 5           Various respiratory tract diseases are associated with clearance alterations. Evaluation of
 6      clearance in individuals with lung disease requires careful interpretation of results, because
 7      differences in deposition of particles used to assess clearance function may occur between
 8      normal individuals and those with disease; this would impact directly on the measured clearance
 9      rates, especially in the tracheobronchial tree. Earlier studies reported in the 1996 PM AQCD
10      (U.S. Environmental Protection Agency,  1996) noted findings of (a) slower nasal mucociliary
11      clearance in humans with chronic sinusitis, bronchiectasis, rhinitis, or cystic fibrosis and (b)
12      slowed bronchial mucus transport associated with bronchial carcinoma, chronic bronchitis,
13      asthma, and various acute respiratory infections.  However, a recent study by Svartengren et al.
14      (1996a) concluded, based on deposition and clearance patterns, that particles cleared equally
15      effectively from the small ciliated airways of healthy humans and those with mild to moderate
16      asthma; but, this similarity was ascribed to effective therapy for the asthmatics.
17           In another study, Svartengren  et al. (1996b) examined clearance from the TB region in
18      adults with chronic bronchitis who inhaled 6-//m Teflon particles.  Based on calculations,
19      particle deposition was assumed to be in small ciliated airways at low flow and in larger airways
20      at higher flow.  The results were compared to those obtained in healthy subjects from other
21      studies. At low flow, a larger fraction of particles was retained over 72 h in people with chronic
22      bronchitis compared to healthy subjects, indicating that clearance resulting from spontaneous
23      cough could not fully compensate for impaired mucociliary transport in small airways.  For larger
24      airways, patients with chronic bronchitis cleared a larger fraction of the deposited particles over
25      72 h than did healthy subjects, but this was reportedly because of differences in deposition
26      resulting from airway obstruction.
27           An important mechanism of clearance  from the tracheobronchial region, under some
28      circumstances, is cough. Although  cough can be a reaction to an inhaled stimulus, in most
29      individuals with respiratory infections and disease, spontaneous coughing also serves to clear the
30      upper bronchial airways by dislodging mucus from the airway surface. Recent studies confirm
31      that this mechanism likely plays a significant role in clearance for people with mucus

        April 2002                                 6-52         DRAFT-DO NOT QUOTE OR CITE

-------
 1      hypersecretion, at least for the upper bronchial tree, and for a wide range of deposited particle
 2      sizes (0.5 to 5 //m) (Toms et al., 1997; Groth et al., 1997). There appears to be a general trend
 3      towards an association between the extent (i.e., number) of spontaneous coughs and the rate of
 4      particle clearance, with faster clearance being associated with a greater number of coughs (Groth
 5      et al., 1997). Thus, recent evidence continues to support cough as an adjunct to mucociliary
 6      movement in the removal of particles from the lungs of individuals with COPD.  However, some
 7      recent evidence suggests that, like mucociliary function, cough-induced clearance may become
 8      depressed with worsening airway disease. Noone et al. (1999) found that the efficacy of
 9      clearance via cough in patients with primary ciliary dyskinesia (who rely on coughing for
10      clearance because of immotile cilia) correlated with lung function (FEV1), in that decreased
11      cough clearance was associated with decreased percentage of predicted FEV1.
12           Earlier reported studies (U.S. Environmental Protection Agency, 1996) indicated that rates
13      of A region particle clearance were reduced in humans with chronic obstructive lung disease and
14      in laboratory animals with viral infections; whereas the viability and functional activity of
15      macrophages were impaired in human asthmatics and in animals with viral-induced lung
16      infections. However, any modification of functional properties of macrophages appears to be
17      injury-specific, in that they reflect the nature and anatomic pattern of disease.
18           One factor that may affect clearance of particles is the integrity of the epithelial surface
19      lining of the lungs.  Damage or injury to the epithelium may result from disease or from the
20      inhalation of chemical irritants.  Earlier studies performed with particle instillation had shown
21      that alveolar epithelial damage in mice at the time of deposition resulted in increased
22      translocation of inert carbon to pulmonary interstitial macrophages (Adamson and Hedgecock,
23      1995).  A similar response was observed in a more recent assessment (Adamson and Prieditis,
24      1998), whereby silica (<0.3 //m) was instilled into a lung having alveolar epithelial damage (as
25      evidenced by increased permeability) and particles were noted to reach the interstitium and
26      lymph nodes.
27
28
29      6.4  PARTICLE OVERLOAD
30           Experimental studies using some laboratory rodents have employed high exposure
31      concentrations of relatively nontoxic, poorly soluble particles. These particle loads interfered
        April 2002                                6-53        DRAFT-DO NOT QUOTE OR CITE

-------
 1      with normal clearance mechanisms, producing clearance rates different from those that would
 2      occur at lower exposure levels.  Prolonged exposure to high particle concentrations is associated
 3      with a phenomenon that has been termed particle "overload", defined as the overwhelming of
 4      macrophage-mediated clearance by the deposition of particles at a rate that exceeds the capacity
 5      of that clearance pathway. It has been suggested that, in the rat, overload is more dependent
 6      upon the volume rather than the mass of particles (Tran et al., 2000) and that volumetric
 7      overloading will begin when particle retention approaches 1 mg particles/g lung tissue (Morrow,
 8      1988).  Overload is a nonspecific effect noted in experimental studies using many different kinds
 9      of poorly soluble particles and results in A region clearance slowing or stasis, with an associated
10      chronic inflammation and aggregation of macrophages in the lungs  and increased translocation of
11      particles into the interstitium.
12          The relevance of lung overload to humans exposed to poorly soluble, nonfibrous particles
13      remains unclear. Although it is likely to be of little relevance for most "real world" ambient
14      exposures, it may be of concern in interpreting some long-term experimental exposure data and,
15      perhaps, also for occupational exposures.  For example, it has been  suggested that a condition
16      called progressive massive fibrosis, which is unique to humans, has features indicating that dust
17      overload is a factor in its pathogenesis (Green, 2000).  This condition is associated with
18      cumulative dust exposure and impaired clearance and can occur following high exposure
19      concentrations associated with occupational situations. In addition, any relevance to humans is
20      clouded by the suggestion that macrophage-mediated clearance is normally slower, and perhaps
21      of less relative importance in overall clearance, in humans than in rats (Morrow,  1994), and that
22      there can be significant differences in macrophage loading between species. On the other hand,
23      overload may be a factor in individuals with compromised lungs even under normal exposure
24      conditions. Thus,  it has been hypothesized (Miller et al.,  1995) that localized overload of particle
25      clearance mechanisms in people with compromised lung status may occur, whereby clearance is
26      overwhelmed and  results in morbidity or mortality from particle exposure.
27
28
29
        April 2002                                6-54        DRAFT-DO NOT QUOTE OR CITE

-------
 1      6.5 COMPARISON OF DEPOSITION AND CLEARANCE PATTERNS
 2          OF PARTICLES ADMINISTERED BY INHALATION AND
 3          INTRATRACHEAL INSTILLATION
 4           The most relevant exposure route by which to evaluate the toxicity of paniculate matter is
 5      inhalation.  However, many toxicological studies deliver particles by intratracheal instillation.
 6      This latter technique has been used because it is easy to perform, requires significantly less effort,
 7      cost, and amount of test material than does inhalation, and can deliver a known, exact dose of a
 8      toxicant to the lungs. Because particle disposition is a determinant of dose, it is important to
 9      compare deposition and clearance of particles delivered by these two routes in order to evaluate
10      the relevance of studies using instillation.  However, in most instillation studies, the effect of this
11      route of administration on particle deposition and clearance per se was not examined. Although
12      these parameters were evaluated in some studies, it has been very difficult to compare particle
13      deposition/clearance between different inhalation and instillation studies because  of differences
14      in experimental procedures and in the manner by which particle deposition/clearance was
15      quantitated.  A recent paper provides a detailed evaluation of the role of instillation in respiratory
16      tract dosimetry and toxicology studies (Driscoll et al., 2000); and a short summary derived from
17      this paper is provided below in this section.
18           The pattern of initial regional deposition is strongly influenced by the exposure technique
19      used. Furthermore, the patterns within specific respiratory tract regions also are influenced in
20      this regard.  Depending on particle size,  inhalation results in varying degrees of deposition within
21      the ET airways, a region that is completely bypassed by instillation.  Thus, differences in amount
22      of particles deposited in the lower airways will occur between the two procedures, especially for
23      those particles in the coarse mode. This is important if inhaled particles in ambient air affect the
24      upper respiratory tract and such responses  are then involved in the evaluation of health outcomes.
25           Exposure technique also influences the intrapulmonary distribution of particles, which
26      potentially would affect routes and rates of ultimate clearance from the lungs and  dose delivered
27      to specific sites within the respiratory tract or to extrapulmonary organs.  Intratracheal instillation
28      tends to disperse particles fairly evenly within the TB region but can result in heterogeneous
29      distribution in the A region; whereas inhalation tends to produce a more homogeneous
30      distribution throughout the major conducting airways as well as the A region for the same
31      particles. Thus, inhalation results in a randomized distribution of particles within the lungs;

        April 2002                                6-55        DRAFT-DO NOT QUOTE OR CITE

-------
 1      whereas intratracheal instillation produces an heterogeneous distribution, in that the periphery of
 2      the lung receives little particle load and most of the instilled particles are found in regions that
 3      have a short path length from the major airways.  Furthermore, inhalation results in greater
 4      deposition in apical areas of the lungs and less in basal areas; whereas intratracheal instillation
 5      results in less apical than basal deposition. Thus, toxicological effects from instilled materials
 6      may not represent those which would occur following inhalation, due to differences in sites of
 7      initial deposition following exposure.  In addition, instillation studies generally deliver high
 8      doses to the lungs, much higher than those which would occur with realistic inhalation exposure.
 9      This would also clearly affect the initial dose delivered to target tissue and its relevance to
10      ambient exposure.
11           Comparison of the kinetics of clearance of particles administered by instillation or
12      inhalation have shown similarities, as well as differences, in rates for different clearance phases,
13      depending on the exposure technique used (Oberdorster et al., 1997). However, some of the
14      differences in kinetics may be explained by differences  in the initial  sites of deposition.  One of
15      the maj or pathways of clearance involves particle uptake and removal via pulmonary
16      macrophages.  Dorries and Valberg (1992) noted that inhalation resulted in a lower percentage of
17      particles recovered in lavaged cells and a more even distribution of particles among
18      macrophages.  More individual cells received measurable amounts of particles via inhalation than
19      via intratracheal instillation; whereas with the latter, many cells received little or no particles and
20      others received very high burdens.  Furthermore,  with intratracheal instillation,  macrophages at
21      the lung periphery contained few, if any, particles; whereas  cells in the regions of highest
22      deposition were overloaded, reflecting the heterogeneity of particle distribution when particles
23      are administered via instillation.  Also, both the relative number of particles phagocytized by
24      macrophages as well as the percentage of these cells involved in phagocytosis is affected by the
25      burden of administered particles, which is clearly different in instillation and inhalation (Suarez
26      et al., 2001). Thus, when guinea pigs were administered latex microspheres (1.52-3.97 //m
27      MMAD) by inhalation or instillation, the percentage of cells involved in phagocytosis, as well as
28      the amount of particles per cell, were both significantly higher with the latter route. The route of
29      exposure, therefore, influences particle distribution in the macrophage population and could, by
30      assumption, influence clearance pathways and clearance kinetics.


        April  2002                                  6-56         DRAFT-DO NOT QUOTE OR CITE

-------
 1           In summary, inhalation may result in deposition within the ET region, the extent of which
 2      depends on the size of the particles used.  Of course, intratracheal instillation bypasses this
 3      portion of the respiratory tract and delivers particles directly to the tracheobronchial tree.
 4      Although some studies indicate that short (0 to 2 days) and long (100 to 300 days postexposure)
 5      phases of clearance of insoluble particles delivered either by inhalation or intratracheal
 6      instillation are similar, other studies indicate that the percentage retention of particles delivered
 7      by instillation is greater than that for inhalation at least up to 30 days postexposure.  Thus, there
 8      is some inconsistency in this regard.
 9           Perhaps the most consistent conclusion regarding differences between inhalation and
10      intratracheal instillation is related to the intrapulmonary distribution of particles.  Inhalation
11      generally results in a fairly homogeneous distribution of particles throughout the lungs.  On the
12      other hand, instillation results in a heterogeneous distribution, especially within the alveolar
13      region, and focally high concentrations of particles. The bulk of instilled material penetrates
14      beyond the major tracheobronchial airways, but the lung periphery is often virtually devoid of
15      particles. This difference is reflected in particle burdens within macrophages, with those from
16      animals inhaling particles having more homogeneous burdens and those  from animals with
17      instilled particles showing groups of cells with no particles and others with heavy burdens. This
18      difference impacts on clearance pathways, dose to cells and tissues, and systemic absorption.
19      Exposure method, thus, clearly influences dose distribution.
20
21
22      6.6 MODELING THE DISPOSITION OF PARTICLES IN THE
23          RESPIRATORY TRACT
24      6.6.1 Modeling Deposition, Clearance, and Retention
25           Over the years, mathematical models for predicting deposition, clearance and, ultimately,
26      retention of particles in the respiratory tract have been developed. Such models help interpret
27      experimental data and can be used to make dosimetry predictions for cases where data are not
28      available.  In fact, model  predictions described below are estimates based on the best available
29      models at the time of publication and, except where noted, have not been verified by
30      experimental data.

        April 2002                                6-57        DRAFT-DO NOT QUOTE OR CITE

-------
 1           A review of various mathematical deposition models was given by Morrow and Yu (1993)
 2      and in U.S. Environmental Protection Agency (1996). There are three major elements involved
 3      in mathematical modeling.  First, a structural model of the airways must be specified in
 4      mathematical terms. Second, deposition efficiency in each airway must be derived for each of
 5      the various deposition mechanisms. Finally, a computational procedure must be developed to
 6      account for the transport and deposition of the particles in the airways.  As noted earlier, most
 7      models are deterministic, in that particle deposition probabilities are calculated using anatomical
 8      and airflow information on an airway generation by airway generation basis. Other models are
 9      stochastic, whereby modeling is performed using individual particle trajectories and finite
10      element simulations of airflow.
11           Recent reports involve modeling the deposition of ultrafine particles and deposition at
12      airway bifurcations. Zhang and Martonen (1997) used a mathematical model to simulate
13      diffusion deposition of ultrafine particles in the human upper tracheobronchial tree and compared
14      the results to those in a hollow cast obtained by Cohen et al. (1990). The model results were in
15      good agreement with experimental data.  Zhang and Martonen (1997) studied the inertial
16      deposition of particles in symmetric three-dimensional models of airway bifurcations,
17      mathematically examining effects of geometry and flow.  They developed equations for use in
18      predicting deposition based on Stokes numbers, Reynolds numbers, and bifurcation angles for
19      specific inflows.
20           Models for deposition, clearance, and  dosimetry of the respiratory tract of humans have
21      been available for the past four decades.  For example, the International Commission on
22      Radiological Protection (ICRP)  has recommended three different mathematical models during
23      this time period (International Commission on Radiological Protection,  1960, 1979, 1994).
24      These models make it possible to calculate the mass deposition and retention in different parts of
25      the respiratory tract and provide, if needed, mathematical descriptions of the translocation of
26      portions of the  deposited material to other organs and tissues beyond the respiratory tract.
27      A somewhat simplified variation of the 1994 ICRP dosimetry model was used by Snipes et al.
28      (1997) to predict average particle deposition in the ET, T and A regions and retention patterns in
29      the A region, under a repeated exposure situation for two characterized environmental  aerosols
30      obtained from Philadelphia, PA and Phoenix, AZ.  Both of these aerosols had both fine and
31      coarse particles. They found similar retention for the fine particles in both aerosols, but

        April 2002                                 6-58        DRAFT-DO NOT QUOTE OR CITE

-------
 1      significantly different retention for the coarse-mode particles. Because the latter type dominated
 2      the aerosol in the Phoenix sample, this type of evaluation can be used to improve understanding
 3      of the relationship between exposures to ambient PM and retention patterns that affect health
 4      endpoints in residents of areas in which the particle distributions and, therefore, the particle
 5      chemistry may differ.
 6           A morphological model based on laboratory data from planar gamma camera and single-
 7      photon emission tomography images has been developed (Martonen et al., 2000). This model
 8      defines the parenchymal wall in mathematical terms, divides the lung into distinct left and right
 9      components, derives a set of branching angles from experimental measurements, and confines
10      the branching network within the left and right components (so there is no overlapping  of
11      airways). The authors conclude that this more physiologically realistic model can be used to
12      calculate PM deposition patterns for risk assessment.
13           Musante and Martonen (2000c) developed an age-dependent theoretical model to predict
14      dosimetry in the lungs of children.  The model comprises dimensions of individual airways and
15      geometry of branching airway networks within developing lungs and breathing parameters as a
16      function of age.  The model suggests that particle size, age, and activity level markedly affect
17      deposition patterns of inhaled particles. Simulations thus far predict a lung deposition fraction of
18      38% in an adult and 73% (nearly twice as high) in a 7-mo-old for 2 //m particles inhaled during
19      heavy breathing.  The authors conclude that this model will be useful for estimating dose
20      delivered to sensitive subpopulations, such as children.
21           Segal et al. (2000a) developed a computer model, noted earlier, for airflow and particle
22      motion in the lungs of children to study how airway disease, specifically cancer, affects inhaled
23      PM deposition. The model considers how tumor characteristics (size and location) and
24      ventilatory parameters (breathing rates and tidal volumes) influence particle trajectories and
25      deposition patterns. The findings indicate that PM may be deposited on the upstream surfaces of
26      tumors because of enhanced efficiency of inertial impaction. Also, submicron particles and
27      larger particles, respectively, may be deposited on the downstream surfaces of tumors because of
28      enhanced efficiency of diffusion and sedimentation.  The mechanisms of diffusion and
29      sedimentation are functions of the particle residence times in airways.  Eddies downstream of
30      tumors would trap particles and allow more time for deposition to occur by diffusion and


        April 2002                                 6-59        DRAFT-DO NOT QUOTE OR CITE

-------
 1      sedimentation. The authors conclude that particle deposition is complicated by the presence of
 2      airway disease, but that the effects are systematic and predictable.
 3           Segal et al. (2000b) have used a traditional mathematical model based on Weibel's lung
 4      morphology and calculated total lung deposition fraction of 1 to 5 //m diameter particles in
 5      healthy adults. Airway dimensions were scaled by individual lung volume.  Deposition
 6      predictions were made with both plug flow and parabolic flow profiles in the airways.  The
 7      individualized airway dimension improved the accuracy of the predicted values when compared
 8      with experimental data. There were significant differences, however, between the model
 9      predictions and experimental data depending on the flow profiles used, indicating that use of
10      more realistic parameters is essential to improving the accuracy of model predictions.
11           Broday and Georgopoulos (2001) presented a model that solves a variant of the general
12      dynamic equation for size evolution of respirable particles within human tracheobronchial
13      airways. The model considers polydisperse aerosols with respect to size but heterosperse with
14      respect to thermodynamic state and chemical composition.  The aerosols have an initial bimodal
15      lognormal size distribution that evolves with time in response to condensation-evaporation and
16      deposition processes. Simulations reveal that submicron size particles grow rapidly and cause
17      increased number and mass fractions of the particle population to be found in the intermediate
18      size range. Because deposition by diffusion decreases with increasing size, hygroscopic fine
19      particles may persist longer in the inspired air than nonhygroscopic particles of comparable initial
20      size distribution.  In contrast, the enhanced deposition probability of hygroscopic particles
21      initially from the intermediate size range increases their fraction deposited in the airways.  The
22      model demonstrates that the combined effect of growth and deposition tends to decrease the
23      nonuniformity of the persistent aerosol, forming an aerosol which is characterized by size
24      distribution of smaller variance. These factors also alter the deposition profile along airways.
25           Lazaridis et al. (2001) developed a deposition model for humans that was designed to better
26      describe the dynamics of respirable particles within the airways.  The model took into account
27      alterations in aerosol particle size and mass distribution that may result from processes such as
28      nucleation, condensation, coagulation, and gas phase chemical reactions. The airway geometry
29      used was the regular dichotomous model of Weibel,  and it incorporated the influences of airway
30      boundary layers on particle dynamics, although simplified velocity profiles were used so as to
31      maintain a fairly uncomplicated description of respiratory physiology. Thus, this model was

        April 2002                                6-60        DRAFT-DO NOT QUOTE OR CITE

-------
 1      considered to be an improvement over previous models which did not consider either the effects
 2      of boundary layers on both the airborne and deposited particles or the effects of gas-phase
 3      transport processes, because it can account for the polydispersity, multimodality, and
 4      heterogeneous composition of common ambient aerosols. The authors indicate that the model
 5      predictions were both qualitatively and quantitatively consistent with experimental data for
 6      particle deposition within the  TB and A regions.
 7           Another respiratory tract dosimetry model was developed concurrently with the ICRP
 8      model by the National Council on Radiation Protection and Measurements (NCRP, 1997).
 9      As with the ICRP model (International Commission on Radiological Protection, 1994), the
10      NCRP model addresses inhalability of particles, revised subregions of the respiratory tract,
11      dissolution-absorption as an important aspect of the model, and body size and age. The NCRP
12      model defines the respiratory tract in terms of a naso-oro-pharyngo-laryngeal (NOPL) region, a
13      tracheobronchial (TB) region, a pulmonary (P) region, and lung-associated lymph nodes (LN).
14      Deposition and clearance are calculated separately for each of these regions.  As with the 1994
15      ICRP model, inhalability of aerosol particles is considered, and deposition in the various regions
16      of the respiratory tract is modeled using methods that relate to mechanisms of inertial impaction,
17      sedimentation, and diffusion.
18           Fractional deposition in  the NOPL region was developed from empirical relationships
19      between particle diameter and air flow rate.  Deposition in the TB and P regions were projected
20      from model calculations, based on geometric or aerodynamic particle diameter and physical
21      deposition mechanisms such as impaction, sedimentation, diffusion, and interception.
22      Deposition in the TB and P regions used the lung model of Yeh and Schum (1980) with a method
23      of calculation similar to that of Findeisen (1935) and Landahl (1950).  This method was modified
24      to accomodate an adjustment of lung volume and substitution of realistic deposition equations.
25      These calculations were based on air flow information and idealized morphometry and used a
26      typical pathway model.  Comparison of regional deposition fraction predictions between the
27      NCRP and ICRP models was  provided in U.S. Environmental Protection Agency (1996).  The
28      definition of inhalability was that of the American Conference of Governmental Industrial
29      Hygenists (1985).  Breathing frequency, tidal volume, and functional residual capacity were the
30      ventilatory factors used to model  deposition.  These were related to body weight and to three
31      levels of physical activity, namely low activity, light exertion, and heavy exertion.

        April 2002                                6-61        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Clearance from all regions of the respiratory tract was considered to result from
 2      competitive mechanical and absorptive mechanisms. Mechanical clearance in the NOPL and TB
 3      regions was considered to result from mucociliary transport. This was represented in the model
 4      as a series of escalators moving towards the glottis and where each airway had an effective
 5      clearance velocity.  Clearance from the P region was represented by fractional daily clearance
 6      rates to the TB region, the pulmonary LN region, and the blood.  A fundamental assumption in
 7      the model was that the rates for absorption into blood were the same in all regions of the
 8      respiratory tract; the rates of dissolution-absorption of particles and their constituents were
 9      derived from clearance data primarily from laboratory animals.  The effect of body growth on
10      particle deposition also was considered in the model, but particle clearance rates were assumed to
11      be independent of age. Some consideration for compromised individuals was incorporated into
12      the model by altering normal rates for the NOPL and TB regions.
13           Mathematical deposition models for a number of nonhuman species have been developed;
14      these were discussed previously in the 1996 PM AQCD (U.S. Environmental Protection Agency,
15      1996).  Despite difficulties, modeling studies in laboratory animals remain a useful step in
16      extrapolating exposure-dose-response relationships from laboratory animals to humans.
17           Respiratory-tract clearance begins immediately upon deposition of inhaled particles. Given
18      sufficient time, the deposited particles may be removed completely by these clearance processes.
19      However, single inhalation exposures may be the exception rather than the rule. It generally is
20      accepted that repeated or chronic exposures are common for environmental aerosols. As a result
21      of such exposures, accumulation of particles may occur.  Chronic exposures produce respiratory
22      tract burdens of inhaled particles that continue to increase with time until the rate of deposition is
23      balanced by the rate of clearance.  This is defined as the "equilibrium respiratory tract burden".
24           It is important to evaluate these accumulation patterns, especially when assessing ambient
25      chronic exposures, because they dictate what the equilibrium respiratory tract burdens of inhaled
26      particles will be for a specified exposure atmosphere. Equivalent concentrations can be defined
27      as "species-dependent concentrations of airborne particles which, when chronically inhaled,
28      produce equal lung deposits of inhaled particles per gram of lung during a specified exposure
29      period" (Schlesinger et al., 1997). Available data and approaches to evaluate exposure
30      atmospheres that produce similar respiratory tract burdens in laboratory animals and humans
31      were discussed in detail in the 1996 PM AQCD.

        April 2002                                 6-62        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Several laboratory animal models have been developed to help interpret results from
 2      specific studies that involved chronic inhalation exposures to nonradioactive particles (Wolff
 3      et al., 1987; Strom et al., 1988;  Stober et al.,  1994). These models were adapted to data from
 4      studies involving high level chronic inhalation exposures in which massive lung burdens of low
 5      toxicity, poorly soluble particles were accumulated. Koch and Stober (2001) further adapted
 6      clearance models for more relevant particle deposition in the pulmonary region.  They published
 7      a pulmonary retention model that accounts for dissolution and macrophage-mediated removal of
 8      deposited polydisperse aerosol particles. The model provides a mathematical solution for the
 9      size distribution of particles in the surfactant layer of the alveolar surface and in the cell plasma
10      of alveolar macrophages and accounts for the different kinetics and biological effects in the two
11      compartments. It does not, however, account for particle penetration to the lung interstitium and
12      particle clearance by the lymph system.
13           The multiple-path models of Anjilvel and Asgharian (1995) for rat lung and its extension
14      by Subramaniam et al. (1999) for human lungs describe a method for calculating a deposited
15      fraction for a specific size distribution based  on a summary of published data on regional
16      deposition of different size particles. The method is based on constructing nomograms that are
17      used to estimate alveolar deposition fractions for three species (human, monkey, and rat). The
18      data are then incorporated into a regression model that calculates more exact deposition fractions
19      in the whole lung for monodisperse and polydisperse aerosols for ultrafme through coarse
20      particle sizes.  The model is somewhat constrained at present because of limitations in the
21      underlying deposition database.
22           Tran et al. (1999) used a mathematical  model of clearance and retention in the A region of
23      rats lungs to determine the extent to which a  sequence of clearance mechanisms and pathways
24      could explain experimental data obtained from inhalation studies using relatively insoluble
25      particles.  These pathways were phagocytosis by macrophages with subsequent clearance,
26      transfer of particles into the interstitium and to lymph nodes, and overloading of defense
27      mechanisms.  The model comprised a description of the complete defense system in this region,
28      using both clearance and transfer processes represented by sets of equations.  The authors suggest
29      that the model could be used to examine the consistency of various hypotheses concerning the
30      fate of inhaled particles and could be used  for species other than the rat with appropriate scaling.


        April 2002                                 6-63        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Hofmann et al. (2000) used three different morphometric models of the rat lung to compute
 2      particle deposition in the acinar (alveolar) airways:  the multipath lung model (MPL) with a fixed
 3      airway geometry; the stochastic lung (SL) model with a randomly selected branching structure;
 4      and a hybrid of the MPL and SL models.  They calculated total and regional deposition for a
 5      range of particle sizes during quiet and heavy breathing. Although the total bronchial and acinar
 6      deposition fractions were similar for the three models, the SL and the hybrid models predicted a
 7      substantial variation in particle deposition among different acini. Acinar deposition variances in
 8      the MPL model were consistently smaller than in the SL and the hybrid lung models. The
 9      authors conclude that the similarity of acinar deposition variations in the latter two models and
10      their independence of the breathing pattern suggest that the heterogeneity of the acinar airway
11      structure is primarily responsible for the heterogeneity of acinar particle deposition.
12           The combination of MPL and SL models developed for the human lung takes into
13      consideration both intra- and inter-human variability in airway structure. The models also have
14      been developed to approximately the same level of complexity for laboratory animals and,
15      therefore, can be readily used for interspecies extrapolation (Asgharian et al., 1999).  A variation
16      of these models will soon be developed for inclusion of the airway geometry of children. By the
17      incorporation of particle clearance in the TB region (Asgharian et al., 2001) and hopefully in the
18      alveolar region (Koch and Stober, 2001), this suite of models should prove to be very useful in
19      better predicting PM dosimetry in humans.
20
21      6.6.2 Models To Estimate Retained Dose
22           Models have been used routinely to express retained dose in terms of temporal patterns for
23      A region retention of acutely inhaled materials.  Available information for a variety of
24      mammalian species, including humans, can be used to predict deposition patterns in the
25      respiratory tract for inhalable aerosols  with reasonable degrees of accuracy.  Additionally,
26      alveolar clearance data for non-human mammalian species commonly used in inhalation studies
27      are available from numerous experiments that involved inhaled radioactive particles.
28           An important factor in using models to predict retention patterns in laboratory animals or
29      humans is the dissolution-absorption rate of the inhaled material. Factors that affect the
30      dissolution of materials or the leaching of their constituents in physiological fluids and the
31      subsequent absorption of these constituents are not fully understood. Solubility is known to be
        April 2002                                6-64        DRAFT-DO NOT QUOTE OR CITE

-------
 1      influenced by the surface-to-volume ratio and other surface properties of particles (Mercer, 1967;
 2      Morrow, 1973). The rates at which dissolution and absorption processes occur are influenced by
 3      factors that include the chemical composition of the material. Temperature history of materials is
 4      also an important consideration for some metal oxides.  For example, in controlled laboratory
 5      environments, the solubility of oxides usually decreases when the oxides are produced at high
 6      temperatures, which generally results in compact particles having small surface-to-volume ratios.
 7      It is sometimes possible to accurately predict dissolution-absorption characteristics of materials
 8      based on physical/chemical considerations, but predictions for in-vivo dissolution-absorption
 9      rates for most materials, especially if they contain multivalent cations or anions, should be
10      confirmed experimentally.
11           Phagocytic cells, primarily macrophages, clearly play a role in dissolution-absorption of
12      particles retained in the respiratory tract (Kreyling, 1992). Some particles dissolve within the
13      phagosomes because of the acidic milieu in those organelles (Lundborg et al., 1984,  1985), but
14      the dissolved material may remain associated with the phagosomes or other organelles in the
15      macrophage rather than diffuse out of the macrophage to be absorbed and transported elsewhere
16      (Cuddihy, 1984).  This same phenomenon has been reported for organic materials. For example,
17      covalent binding of benzo[a]pyrene or metabolites to cellular macromolecules resulted in an
18      increased alveolar retention time for that compound after inhalation exposures of rats (Medinsky
19      andKampcik, 1985). Understanding these phenomena and recognizing species similarities and
20      differences are important for evaluating alveolar retention and clearance processes and for
21      interpreting the results of inhalation studies.
22           Dissolution-absorption of materials in the respiratory tract is clearly dependent on the
23      chemical and physical attributes of the material.  Although it is  possible to predict rates of
24      dissolution-absorption, it is prudent to determine this important clearance parameter
25      experimentally. It is important to understand the impact of this clearance process for the lungs,
26      tracheobronchial lymph nodes, and other body organs that might receive particles  or their
27      constituents that enter the circulatory system from the lung.
28           Insufficient data were available to adequately model long-term retention of particles
29      deposited in the conducting airways of any mammalian  species at the time of the 1996 PM
30      AQCD, and this still remains the case.  Additional research must be done to provide the
31      information needed to properly evaluate retention of particles in conducting airways. However,

        April 2002                                6-65        DRAFT-DO NOT QUOTE OR CITE

-------
 1      a number of earlier studies, discussed in the 1996 document and in Section 6.2.2.2 herein, noted
 2      that some particles were retained for relatively long times in the tracheobronchial regions,
 3      effectively contradicting the general conclusion that almost all inhaled particles that deposit in
 4      the TB region clear within hours or days. These studies have demonstrated that variable portions
 5      of the particles that deposit in, or are cleared through, the TB region are retained with half times
 6      on the order of weeks or months.  Long-term retention and clearance patterns for particles that
 7      deposit in the ET and TB regions must continue to be thoroughly evaluated because of the
 8      implications of this information for respiratory tract dosimetry and risk assessment.
 9           Model projections are possible for the A region using the cumulative information in the
10      scientific literature relevant to deposition, retention, and clearance of inhaled particles.
11      Clearance parameters for six laboratory animal species were summarized in U.S. Environmental
12      Protection Agency (1996).  Nikula et al. (1997) evaluated results in rats and monkeys exposed to
13      high levels of either diesel soot or coal dust.  Although the total amount of retained material was
14      similar in both species, the rats retained a greater portion in the lumens of the alveolar ducts and
15      alveoli than did monkeys; whereas the monkeys retained a greater portion of the material in the
16      interstitium. The investigators concluded that intrapulmonary retention patterns in one species
17      may not be predictive of those in another species at high levels of exposure, but this may not be
18      the case at lower levels  of exposure.
19           The influence of exposure concentration on the pattern of particle retention in rats (exposed
20      to diesel soot) and humans (exposed to coal dust) was examined by Nikula et al. (2000) using
21      histological lung sections obtained from both  species.  The exposure concentrations for  diesel
22      soot were 0.35, 3.5, or 7.0 //g/m3, and exposure duration was 7 h/day, 5 days/week for 24 mo.
23      The human lung sections were obtained from  nonsmoking nonminers, nonsmoking coal miners
24      exposed to levels <2 //g dust/m3 for 3 to 20 years, or nonsmoking miners exposed to <10 //g/m3
25      for 33 to 50 years. In both  species, the amount of retained material (using morphometric
26      techniques based on the volume density of deposition) increased with increasing dose (which is
27      related to exposure duration and concentration). In rats, the  diesel exhaust particles were found
28      to be primarily in the lumens of the alveolar duct and alveoli; whereas in humans, retained dust
29      was found primarily in the interstitial tissue within the respiratory acini.
30
31

        April 2002                                 6-66        DRAFT-DO NOT  QUOTE OR CITE

-------
 1      6.6.3 Fluid Dynamics Models for Deposition Calculations
 2           The available models developed to simulate particulate deposition in the lung are based on
 3      simplifying assumptions about the morphometry of the lung and the fluid dynamics of inspired
 4      air through a branching airway system. All of the approaches, whether analytic, symmetric, or
 5      multiple-path models, simulate particle behavior in an "idealized" respiratory system with
 6      homogeneous geometry and flow profile and can only predict average regional and total
 7      dosimetry in the lung. As new models are developed, they will better predict particle deposition
 8      patterns in a more realistic airway geometry under realistic flow conditions that can result in local
 9      inhomogeneities of particle deposition and the formation of hot-spots. One example is the model
10      of ventilation distribution in the human lung developed by Chang and Yu (1999). This model
11      was designed as an improvement over those that assumed uniform ventilation in the lungs,
12      because it better simulated the effect of airway dynamics on the distribution of ventilation under
13      different conditions which may occur in the various lobes of the lungs and under various
14      inspiratory flow rates. The authors indicate that the results of the model compared favorably
15      with experimental  data and that the model will be incorporated into a particle deposition model
16      that will  allow for the evaluation of the nonuniformity of deposition within the lungs  resulting
17      from the physiological situation of nonuniform distribution of ventilation. Computational fluid
18      dynamics (CFD) modeling adds another step to better model development by providing increased
19      ability to predict local airflow and particle deposition patterns and provide a better representation
20      of extrathoracic deposition in the human respiratory tract. The CFD models developed to date,
21      however, also are limited in scope because they are unable to simulate flow in the more complex
22      gas exchange regions. Due to a lack of more realistic simulations for the lower airways, they
23      impose another "idealized" boundary condition at the distal end of the human respiratory tract.
24           Airflow patterns within the lung are determined by the interplay of structural and
25      ventilatory conditions.  These flow patterns govern the deposition kinetics of entrained particles
26      in the inspired air.  A number of CFD software programs are available to simulate airflow
27      patterns in the lung by numerically solving the Navier-Stokes equations (White, 1974). The CFD
28      modeling requires  a computer reconstruction of the appropriate lung region and the application of
29      boundary conditions.  The flow field resulting from the CFD modeling is represented by velocity
30      vectors in the grid  points  of a two- or three-dimensional mesh. Numerical models of particle
31      deposition patterns are computed by  simulating the trajectories of particles introduced into these
        April 2002                                6-67        DRAFT-DO NOT QUOTE OR CITE

-------
 1      flow streams after solving for the particles' equation of motion.  Such CFD models have been
 2      developed for different regions of the respiratory tract, including the nasal cavity (Yu et al., 1998;
 3      Sarangapani and Wexler, 2000); larynx (Martonen et al. 1993; Katz et al., 1997; Katz, 2001);
 4      major airway bifurcations (Gradon and Orlicki, 1990; Balashazy and Hofmann, 1993a,b, 1995,
 5      2001; Heistracher and Hofmann, 1995; Lee et al., 1996; Zhang et al., 1997, 2000, 2001, 2002;
 6      Comer et al., 2000, 2001a,b); and alveoli (Tsuda et al., 1994a,b;  Chantal, 2001).
 7           Kimbell (2001) has recently reviewed the literature on CFD models of the upper respiratory
 8      tract (URT).  Most of these models have focused on characterizing the airflow patterns in the
 9      URT and have not included simulation of particulate dosimetry.  Keyhani et al. (1995) were the
10      first to use computer-aided tomography (CAT) scans of the human nasal cavity to construct an
11      anatomically accurate three-dimensional airflow model of the human nose.  Subramaniam et al.
12      (1998) used MRI scan data to extend these CFD studies to include the nasopharynx.  However,
13      neither of these studies investigated particle deposition in the upper respiratory tract.
14           Yu et al. (1998) have developed a three-dimensional CFD model of the entire human upper
15      respiratory tract, including the nasal  airway, oral airway, laryngeal airway, and the first two
16      generations of the tracheobronchial airway. They have used this CFD model to investigate the
17      effect of breathing pattern, i.e., nasal breathing, oral breathing, and simultaneous nasal and oral
18      breathing, on airflow and ultrafme particle deposition. They concluded that the ultrafme particle
19      deposition simulated using the CFD  model was in reasonable agreement with the corresponding
20      experimental measurements.  In a study led by Sarangapani and Wexler (2000), an upper
21      respiratory tract CFD model that included the nasal cavity, nasopharynx, pharynx,  and larynx was
22      developed to study the deposition efficiency of hygroscopic and non-hygroscopic particles in this
23      region.  They used the CFD model to simulate the temperature and water vapor conditions in the
24      upper airways and predicted high relative humidity conditions in this region. They also
25      simulated particle trajectories for 0.5 //m, 1 //m, and 5 //m particles under physiologically
26      realistic flow rates.  The predictions  of the CFD model indicated that high relative humidity
27      conditions contribute to rapid growth of hygroscopic particles and would dramatically alter the
28      deposition characteristics of ambient hygroscopic aerosols.
29           Stapleton et al. (2000) investigated deposition of a polydisperse  aerosol (MMD = 4.8 //m
30      and GSD = 1.65) in a replica of a human mouth and throat, using both experimental results and
31      3-D CFD simulation.  They found that CFD results were comparable with experimental results

        April 2002                                6-68        DRAFT-DO NOT QUOTE OR CITE

-------
 1      for a laminar flow case but were more than 200% greater for a turbulent flow case.  The results
 2      suggest that accurate predictions of particle deposition in a complex airway geometry requires a
 3      careful evaluation of geometric and fluid dynamic factors in developing CFD models.
 4           Due to the complex structural features and physiological conditions of the human laryngeal
 5      region, only a limited number of modeling studies have been conducted to evaluate laryngeal
 6      fluid dynamics and particle deposition. A high degree of inter-subject variability, a compliant
 7      wall that presents challenges in setting appropriate boundary conditions, and a complex turbulent
 8      flow field are some of the difficulties encountered in developing CFD models of the laryngeal
 9      airways. Martonen et al. (1993) investigated laryngeal airflow using a two-dimensional  CFD
10      model and concluded that laryngeal morphology exerts a pronounced influence on regional flow,
11      as well as fluid motion in the trachea and the main bronchi. In this study, the glottal aperture
12      (defined by the geometry of the vocal folds) was allowed to change in a prescribed manner with
13      the volume of inspiratory flow (Martonen and Lowe, 1983), and three flow rates corresponding
14      to different human activity were examined.
15           In a subsequent CFD analysis, a three-dimensional model of the larynx based on
16      measurements of human replica laryngeal casts (Martonen  and Lowe,  1983; Katz and Martonen,
17      1996; Katz et al., 1997) simulated the flow field in the larynx and trachea under steady
18      inspiratory flow conditions at three flow rates.  They observed that the complex geometry
19      produces jets, recirculation zones, and circumferential  flow that may directly influence particle
20      deposition at select sites within the larynx and  tracheobronchial airways.  The primary
21      characteristics of the simulated flow field were a central jet penetrating into the trachea created
22      by the ventricular and vocal folds, a recirculating zone downstream of the vocal folds, and a
23      circumferential secondary flow. Recently, a computational model for fluid dynamics and particle
24      motion for inspiratory flow through the human larynx and trachea has been described (Katz,
25      2001). This model calculates the trajectory of single particles introduced  at the entrance to the
26      larynx using a stochastic model for turbulent fluctuations incorporated into the particles'
27      equation of motion and time-averaged flow fields in the larynx and trachea. The effects of flow
28      rate and initial particle location on overall deposition were presented in the form of probability
29      density histograms of final particle deposition sites.  At present, however, there are no
30      experimental data to validate results of such modeling.


        April 2002                                6-69        DRAFT-DO NOT QUOTE OR CITE

-------
 1           A number of CFD models have been developed to study fluid flow and particle deposition
 2      patterns in airway bifurcations.  The bifurcation geometries that have been modeled include:
 3      two-dimensional (Li and Ahmadi, 1995); idealized three-dimensional using circular airways
 4      (Kinsara et al., 1993) or square channels (Asgharian and Anjilvel, 1994); symmetric bifurcations
 5      (Balashazy and Hofmann, 1993a,b); or physiologically realistic asymmetric single (Balashazy
 6      and Hofmann, 1995; Heistracher and Hofmann, 1995) and multiple bifurcation models (Lee
 7      et al., 1995; Heistracher and Hofmann, 1997; Comer et  al., 2000, 2001; Zhang et al., 2000, 2001,
 8      2002), with anatomical irregularities such as cartilaginous rings (Martonen et al., 1994a) and
 9      carinal ridge (Martonen et al., 1994b; Comer et al., 2001a) shapes incorporated. The CFD flow
10      simulations in the bifurcating geometry models show distinct asymmetry in the axial (primary)
11      and radial (secondary) velocity profile in the daughter and parent airway during inspiration and
12      expiration, respectively. In a systematic investigation of flow patterns in airway bifurcations,
13      numerical simulations were performed to study primary flow (Martonen et al., 200la), secondary
14      currents (Martonen et al., 200Ib), and localized flow conditions (Martonen et al., 200Ic) for
15      different initial flow rates. The effects of inlet conditions, Reynolds numbers, ratio of airway
16      diameters, and branching angles with respect to intensity of primary flow, vortex patterns of the
17      secondary currents, and reverse flow in the parent-daughter transition region were investigated.
18      These simulated flow patterns match experimentally-observed flow profiles in airway
19      bifurcations (Schroter and Sudlow, 1969).
20           Gradon and Orlicki (1990) computed the local deposition flux of submicron size particles
21      in a three-dimensional bifurcation model for both inhalation and exhalation; and they found
22      enhanced deposition in the carinal ridge region during inspiration and in the central zone of the
23      parent airway during expiration. Numerical models of particle deposition in symmetric three-
24      dimensional bifurcations were developed by Balashazy  and Hofmann (1993a,b), and these were
25      subsequently extended to incorporate effects of asymmetry in  airway branching (Balashazy and
26      Hofmann,  1995) and physiologically realistic shapes of the bifurcation transition zone and the
27      carinal ridge (Heistracher and Hofmann, 1995; Balashazy and Hofmann, 2001). In these
28      numerical models, three-dimensional airflow patterns were computed by finite difference or
29      finite volume methods, and the trajectories of particles entrained in the airstream were simulated
30      using Monte Carlo techniques considering the simultaneous effects of gravitational settling,
31      inertial impaction, Brownian motion, and interception.  The spatial deposition pattern of inhaled

        April 2002                                6-70        DRAFT-DO NOT QUOTE OR CITE

-------
 1      particles was examined for a range of particle sizes (0.01-10 //m) and flow rates (16-32 L/min
 2      minute volume) by determining the intersection of particle trajectories with the surrounding
 3      surfaces.  The overall deposition rates derived using the CFD models correspond reasonably with
 4      experimental data (Kim and Iglesias, 1989).  These simulations predict deposition hot spots at the
 5      inner side of the daughter airway downstream of the carinal ridge during inspiration,
 6      corresponding to the secondary fluid motion of the inhaled air stream. During exhalation, the
 7      CFD models predict enhanced deposition at the top and bottom parts of the parent airway,
 8      consistent with secondary motion in the exhaled air stream.  These studies indicate that
 9      secondary flow patterns within the bifurcating geometry play a dominant role in determining
10      highly non-uniform local particle deposition patterns.
11          Zhang et al. (1997) numerically simulated particle deposition in three-dimensional
12      bifurcating airways (having varying bifurcation angles) due to inertial impaction during
13      inspiration for a wide range of Reynolds numbers (100-1000). Inlet velocity profile, flow
14      Reynolds number, and bifurcation angle had a substantial  effect on particle deposition efficiency.
15      Based on the simulated results, equations were derived for particle deposition efficiency as a
16      function of nondimensional parameters, such as Stokes number, Reynolds number, and
17      bifurcation angle, and were shown to compare favorably with available experimental results.
18      More recently, Comer et al. (2000) have estimated the deposition efficiency of 3, 5, and 7 //m
19      particles in a three-dimensional double bifurcating airway model for both in-plane and out-of-
20      plane configurations for a wide range of Reynolds numbers (500-2000). They demonstrated
21      deposition in the first bifurcation to be higher than in the second bifurcation, with deposition
22      mostly concentrated near the carinal region.  The non-uniform flow generated by the first
23      bifurcation had a dramatic effect on the deposition pattern in the second bifurcation.  Based on
24      these results, they concluded that use of single bifurcation models are inadequate to capture the
25      complex fluid-particle interactions that occur in multigeneration airway systems.
26          Comer et al. (2001a) further investigated detailed characteristics of the axial and secondary
27      flow in a double bifurcation airway model using 3-D CFD simulation. Effects of carina shape
28      (sharp vs. rounded) and bifurcation plane (planar vs. non-planar) were examined. Particle
29      trajectories and deposition patterns were subsequently investigated in the  same airway model
30      (Comer et al, 2001b). There was a highly localized deposition at and near the carina  both in the
31      first and second bifurcation, and deposition efficiency was much lower in the second bifurcation

        April 2002                                 6-71        DRAFT-DO NOT QUOTE OR CITE

-------
 1      than in the first bifurcation as demonstrated in the earlier study (Comer et al, 2000). They found
 2      that deposition patterns were not much different between the sharp vs. rounded carina shape at
 3      Stokes numbers of 0.04 and 0.12.  However, deposition patterns were altered significantly for
 4      these particles when the bifurcation plane was rotated, suggesting that a careful consideration of
 5      realistic airway morphology is important for accurate prediction of particle deposition by CFD
 6      modeling.
 7           Zhang et al (2000, 2001) extended the studies of Comer et al. described above and
 8      investigated effects of angled inlet tube as well as asymmetric  flow distribution between daughter
 9      branches. The flow asymmetry caused uneven deposition between downstream daughter
10      branches. Also noted was that the absolute deposition amount was higher, but deposition
11      efficiency per se was lower in the high flow branch than in the low flow branch. The intriguing
12      relationship between flow asymmetry and deposition was in fact consistent with experimental
13      data of Kim et al. (1999), indicating that the CFD model could correctly simulate complicated
14      airflow and particle dynamics that may occur in the respiratory airways.
15           Most CFD models use constant inspiratory  or expiratory flows for simplicity and practical
16      reasons.  However, the respiratory airflow is cyclic, and such flow characteristics cannot be fully
17      described by constant flows.  Recent studies of Zhang et al. (2002) investigated particle
18      deposition in a triple bifurcation airway model under cyclic flow conditions mimicking resting
19      and light activity breathing. Deposition dose was obtained for every mm square area.  They
20      found that deposition patterns were similar to those obtained with constant flows. However,
21      deposition efficiencies were greater with  the cyclic flows than  constant flows, and the difference
22      could be as high as 50% for 0.02 < mean Stk < 0.12 during normal breathing.  The CFD results
23      are qualitatively comparable to experimental data (Kim et al, 1991) that showed about 25%
24      increase in deposition with cyclic flows.  With further improvement of airway morphology and
25      computational scheme, CFD modeling could be a valuable tool for exploring the microdosimetry
26      in the airway structure.
27           Current CFD models of the acinar region are limited due to the  complex and dynamic
28      nature of the gas exchange region. Flow  simulation in a linearly increasing volume of a spherical
29      truncated two-dimensional alveolus model show  distinct velocity maxima in the alveolar ducts
30      close to the entrance and exit points of the alveolus and a radial velocity profile in the interior
31      space of the alveolus (Tsuda et al., 1996). This is in contrast to simulations based on a rigid

        April 2002                                6-72        DRAFT-DO NOT QUOTE OR CITE

-------
 1      alveolus (Tsuda, 1994a,b) and suggests that a realistic simulation of the flow pattern in the acinar
 2      region should involve application of time-dependent methods with moving boundary conditions.
 3      Nonuniform deposition patterns, with higher deposition near the alveolar entrance ring, have
 4      been predicted using numerical models (Tsuda, 1994a,b, 1996).
 5           Recent studies of Chantal (2001) examined aerosol transport and deposition in 6-generation
 6      alveolated ducts using 2-D computer simulation. Particle trajectories and deposition patterns
 7      were obtained for one complete breathing cycle (2 s inspiration and 2 s expiration). There were
 8      large non-uniformities in deposition between generations, between ducts of a given generation,
 9      and within each alveolated duct,  suggesting that local deposition dose can be much greater than
10      the mean acinar dose.
11
12
13      6.7 SUMMARY AND CONCLUSIONS
14           An understanding of biological effects of inhaled particulate matter and underlying
15      mechanisms of action requires knowledge of the dosimetry of such material. This is because the
16      dose of particles delivered to a target site or sites of concern, rather than the actual exposure
17      concentration, is the proximal cause of the biological response. Such information is also critical
18      for extrapolation of effects found in controlled exposure studies of animals to those observed in
19      human clinical studies and, also, for relating effects in potentially susceptible persons to those in
20      normal, healthy persons.  Dosimetry involves delineation of the processes of particle deposition,
21      translocation, and clearance. While the current understanding of basic mechanisms of particle
22      dosimetry, clearance, and retention has not changed since the 1996 PM ACQD (U.S.
23      Environmental Protection Agency,  1996), additional information has become available on the
24      role of certain biological  determinants of these processes, such as gender and age; and there has
25      been an expansion of previous knowledge about the relationship between regional deposition and
26      translocation in regard to specific particle size ranges of significance to ambient particulate
27      exposure scenarios. There also has been significant improvement  in the mathematical and CFD
28      modeling of particle dosimetry in the respiratory tract of humans.  Although the models have
29      become more sophisticated and versatile, validation of the models is still needed.
30           One of the areas that has improved since the 1996 PM ACQD is consideration of specific
31      and relevant ambient size particle ranges in deposition studies. One such size mode is the
        April 2002                                6-73        DRAFT-DO NOT QUOTE OR CITE

-------
 1      ultrafine. While further information on respiratory deposition for this size mode is still needed,
 2      there has been an improvement in the understanding of total deposition as a function of particle
 3      size and breathing pattern and of certain aspects of regional deposition of ultrafme particles.
 4      This new information indicates that the ET region, especially the nasal passages, is a very
 5      efficient "filter" for these particles, reducing the amount which would be available for deposition
 6      in the TB and A regions of the respiratory tract. Within the thoracic region, the deposition
 7      distribution of ultrafme particles is highly skewed towards the proximal airway regions and
 8      resembles that of coarse particles. In other words, deposition patterns of ultrafme particles are
 9      very much like those of coarse particles.  Another example involves studies which attempt to
10      evaluate the contribution of fine- and coarse-mode particles to deposition in various parts of the
11      respiratory tract, although there have been only a few of these.
12           It always has been clear that certain host factors affect deposition, and there has been
13      improvement since the 1996 PM AQCD in the understanding of some of these factors,
14      specifically gender and age. Recent information suggests that there are significant gender
15      differences in the homogeneity of deposition as well as the deposition rate, and this could affect
16      susceptibility. In regard to age, recent evaluations employed both mathematical models as well
17      as experimental studies, and most involved comparison of deposition in children compared to
18      adults.  These studies generally indicate that children would receive greater doses of particles per
19      lung surface area than would adults. Unfortunately, deposition studies in another potentially
20      susceptible population, namely the elderly, are still lacking although there have been a number of
21      studies examining effects of chronic pulmonary disease on deposition.  These studies confirmed
22      that significant increases in deposition in obstructed lungs could occur.
23           Once deposited on airway surfaces, particles are subjected to translocation and clearance.
24      While the general pathways of clearance have been known for years, recent information has
25      improved the understanding of translocation of particles within size ranges which may be of
26      specific concern for ambient exposures. One such size mode, as noted above, is the ultrafme;
27      and recent studies indicate that ultrafme particles can be rapidly cleared from the lungs into the
28      systemic circulation and reach extrapulmonary organs.  This provides a mechanism whereby
29      inhaled particles may affect cardiovascular function, as noted in various epidemiological studies
30      (see Chapter 8).


        April 2002                                  6-74        DRAFT-DO NOT QUOTE OR CITE

-------
 1           As with experimental studies, the major improvements in mathematical modeling of
 2     dosimetry involve evaluation of realistic size modes for ambient conditions, as well as
 3     improvements in the precision of these models for more realistic depictions of respiratory tract
 4     airflow patterns and detailed airway structures that may result in deposition "hot spots".  These
 5     improvements include more detailed evaluations of enhanced deposition at airway bifurcations,
 6     use of parameters that allow determination of age differences in dosimetry, and improvement in
 7     the modeling of clearance mechanisms.
 8           Thus, in general, while our understanding of specific aspects of particle dosimetry has
 9     improved since the 1996 PM AQCD, there are still areas in need of further evaluation. These
10     areas of research include dosimetry in susceptible humans, better models for extrapolation
11     between animals used in inhalation studies and humans, and better understanding of differences
12     in the manner in which particles of different and relevant ambient size modes are handled
13     following deposition.  This latter research need is important for determining the potential of
14     various particle types to exert effects systemically, rather than just locally within the respiratory
15     tract.
       April 2002                                6-75        DRAFT-DO NOT QUOTE OR CITE

-------
  1       REFERENCES

  2       Adamson, I. Y. R.; Bowden, D. H. (1981) Dose response of the pulmonary macrophagic system to various
  3             particulates and its relationship to transepithelial passage of free particles. Exp. Lung Res. 2: 165-175.
  4       Adamson, I. Y. R.; Hedgecock, C. (1995) Patterns of particle deposition and retention after instillation to mouse
  5             lung during acute injury and fibrotic repair. Exp. Lung Res. 21: 695-709.
  6       Adamson, I. Y. R.; Prieditis, H. (1998) Silica deposition in the lung during epithelial injury potentiates fibrosis and
  7             increases particle translocation to lymph nodes. Exp. Lung Res. 24: 293-306.
  8       American Conference of Governmental Industrial Hygienists (ACGIH). (1985) Particle size-selective sampling in
  9             the workplace: report of the ACGIH technical committee on air sampling procedures. Cincinnati, OH:
10             American Conference of Governmental Industrial Hygienists.
11       Anjilvel, S.; Asgharian, B. (1995) A multiple-path model of particle deposition in the rat lung. Fundam. Appl.
12             Toxicol. 28: 41-50.
13       Asgharian, B.; Anjilvel, S. (1994) Inertia! and gravitational deposition of particles in a square cross section
14             bifurcating airway. Aerosol Sci. Technol. 20: 177-193.
15       Asgharian, B.; Wood, R.; Schlesinger, R. B.  (1995) Empirical modeling of particle deposition in the alveolar region
16             of the lungs: a basis for interspecies extrapolation. Fundam. Appl. Toxicol. 27: 232-238.
17       Asgharian, B.; Miller, F. J.; Subramaniam, R. P. (1999) Dosimetry software to predict particle deposition in humans
18             and rats. CUT Activities 19(3). Available: http://www.ciit.org/ACT99/ACTIVITIESMAR99/mar99.html
19             [19 December, 2001].
20       Asgharian, B.; Hofmann, W.; Miller, F. J. (2001) Mucociliary clearance of insoluble particles from the
21             tracheobronchial airways of the human lung. Aerosol Sci. 32: 817-832.
22       Balashazy, I.;  Hofmann, W. (1993a) Particle deposition in airway bifurcations-I: inspiratory flow. J. Aerosol Sci.
23             24:745-772.
24       Balashazy, I.;  Hofmann, W. (1993b) Particle deposition in airway bifurcations-II: expiratory flow. J. Aerosol Sci.
25             24:773-786.
26       Balashazy, I.;  Hofmann, W. (1995) Deposition of aerosols in asymmetric airway bifurcations. J. Aerosol Sci.
27             26:273-292.
28       Balashazy, I.;  Hofmann, W. (2001) Fluid dynamics and related particle deposition patterns in human airway
29             bifurcations. In: Martonen, T. B., ed.,  Medical Applications of Computer Modeling: The Respiratory System.
30             WIT Press, Southampton, Boston, p. 84-108.
31       Bennett, W. D.; Ilowite, J. S. (1989) Dual pathway clearance of 99mTc-DTPA from the bronchial mucosa. Am. Rev.
32             Respir. Dis. 139:  1132-1138.
33       Bennett, W. D.; Zeman, K. L. (1998) Deposition of fine particles in children spontaneously breathing at rest.
34             Inhalation Toxicol. 10: 831-842.
35       Bennett, W. D.; Zeman, K. L.; Kim, C. (1996) Variability of fine particle deposition in healthy adults: effect of age
36             and gender. Am. J. Respir. Crit. Care Med. 153: 1641-1647.
37       Bennett, W. D.; Zeman, K. L.; Kang, C. W.;  Schechter, M. S. (1997) Extrathoracic deposition of inhaled, coarse
3 8             particles (4.5 ,am) in children vs. adults. In: Cherry, N.; Ogden, T., eds. Inhaled particles VIII: proceedings of
39             an international symposium on inhaled particles organised by the British Occupational Hygiene Society;
40             August 1996;  Cambridge, UK. Ann. Occup. Hyg. 41(suppl. 1): 497-502.
41       Bennett, W. D.; Zeman, K. L.; Kim, C.; Mascarella, J. (1997b) Enhanced deposition of fine particles in COPD
42             patients spontaneously breathing at rest. Inhalation Toxicol. 9: 1-14.
43       Bennett, W. D.; Scheuch, G.; Zeman, K. L.; Brown, J. S.; Kim, C.; Heyder, J.; Stahlhofen, W. (1998) Bronchial
44             airway  deposition and retention of particles in inhaled boluses: effect of anatomic dead space. J. Appl.
45             Physiol. 85: 685-694.
46       Bennett, W. D.; Scheuch, G.; Zeman, K. L.; Brown, J. S.; Kim, C.; Heyder, J.; Stahlhofen, W. (1999) Regional
47             deposition and retention of particles in shallow, inhaled boluses: effect if lung volume. J. Appl. Physiol.
48             86: 168-173.
49       Brauer, M.; Avila-Casado, C.; Fortoul, T. I.;  Vedal, S.; Stevens, B.; Churg, A. (2001) Air pollution and retained
50             particles in the lung. Environ. Health Perspect. 109: 1039-1043.
51       Broday, D. M.; Georgopoulos, P. G. (2001) Growth and deposition of hygroscopic paniculate matter in the human
52             lungs. Aerosol Sci. Technol. 34: 144-159.
53       Brown, J. S.; Kirby, Z. L.; Bennett, W. D. (2001) Regional deposition of coarse particles and ventilation distribution
54             in healthy subjects and patients with cystic fibrosis. J. Aerosol Med. 14: 443-454.

         April 2002                                      6-76         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Camner, P.; Anderson, M; Philipson, K.; Bailey, A.; Hashish, A.; Jarvis, N.; Bailey, M.; Svartengren, M. (1997)
  2             Human bronchiolar deposition and retention of 6-, 8-, and 10-/u particles. Exp. Lung Res. 23:  517-535.
  3       Chang, Y. H.; Yu, C. P. (1999) A model of ventilation distribution in the human lung. Aerosol Sci. Technol.
  4             30:309-319.
  5       Cheng, Y.-S.; Smith, S. M.; Yeh, H.-C.; Kim, D.-B.; Cheng, K.-H.; Swift, D. L. (1995) Deposition of ultrafine
  6             aerosols and thoron progeny in replicas of nasal airways of young children. Aerosol. Sci. Technol.
  7             23:541-552.
  8       Cheng, K.-H.; Cheng, Y.-S.; Yeh, H.-C.; Guilmette, R. A.; Simpson, S. Q.; Yang, Y.-H.; Swift, D. L. (1996) In vivo
  9             measurements of nasal airway dimensions and ultrafine aerosol deposition in the human nasal and oral
10             airways. J. Aerosol Sci. 27: 785-801.
11       Cheng, K.-H.; Cheng, Y.-S.; Yeh, H.-C.; Swift, D. L. (1997) An experimental method for measuring aerosol
12             deposition efficiency in the human oral airway. Am. Ind. Hyg.  Assoc. J. 58: 207-213.
13       Churg, A.; Brauer, M. (1997) Human lung parenchyma retains PM2 5. Am. J. Respir. Crit. Care Med.
14             155:2109-2111.
15       Churg, A.; Vedal, S. (1996) Carinal and tubular airway particle concentrations in the large airways of non-smokers
16             in the general population: evidence for high particle concentration at airway carinas. Occup. Environ. Med.
17             53:553-558.
18       Cohen, B. S.; Harley, N. H.; Schlesinger, R. B.; Lippmann, M. (1988) Nonuniform particle deposition on
19             tracheobronchial airways: implications for lung  dosimetry. In: Dodgson, J.; McCallum, R. L; Bailey, M. R.;
20             Fisher, D. R., eds. Inhaled particles VI: proceedings of an international symposium and workshop on lung
21             dosimetry; September 1985; Cambridge, United Kingdom. Ann. Occup. Hyg. 32 (suppl. 1): 1045-1053.
22       Cohen, B. S.; Sussman, R. G.; Lippmann, M.  (1990) Ultrafine particle deposition in a human tracheobronchial cast.
23             Aerosol Sci. Technol. 12: 1082-1091.
24       Cohen, B. S.; Xiong, J. Q.; Fang, C.-P.; Li, W. (1998) Deposition of charged particles on lung airways. Health
25             Phys. 74: 554-560.
26       Comer, J. K.; Kleinstreuer, C.; Hyun, S.; Kim, C. S. (2000) Aerosol transport and deposition in sequentially
27             bifurcating airways. J. Biomech. Eng. 122: 152-158.
28       Comer, J. K.; Kleinstreuer, C.; Zhang, Z. (2001a) Flow structures and particle deposition patterns in double
29             bifurcation airway models. Part 1. Air flow fields. J. Fluid Mech. 435: 24-54.
30       Comer, J. K.; Kleinstreuer, C.; Kim, C. S. (200Ib) Flow structures and particle deposition patterns in double
31             bifurcation airway models. Part 2. Aerosol transport and deposition. J. Fluid Mech. 435: 55-80.
32       Crystal, R. G., West, J. B.; Barnes, P. J.; Weibel, E. R., eds. (1997) The lung: scientific foundations. Volume 1.
33             Section III: major components. 2nd ed. Philadelphia, PA: Lippincott-Raven; chapters 30-67, pp. 445-991.
34       Cuddihy, R. G. (1984) Mathematical models for predicting clearance  of inhaled radioactive materials. In: Smith, H.;
3 5             Gerber, G., eds. Lung modelling for inhalation of radioactive materials: proceedings of a meeting jointly
3 6             organized by the Commission of the European Communities and the National Radiological Protection Board;
37             March; Oxford, United Kingdom. Luxembourg: Commission of the European Communities; pp. 167-179;
38             report no. EUR 9384 EN.
39       Cuddihy, R. G.; Yeh, H. C. (1988) Respiratory tract clearance of particles and substances dissociated from particles.
40             In: Mohr, U.; Dungworth, D.; Kimmerle, G.; Lewkowski, J.; McClellan, R.; Stober, W., eds. Inhalation
41             toxicology: the design and interpretation of inhalation studies and their use in risk assessment. New York,
42             NY:  Springer-Verlag; pp. 169-193.
43       Darquenne, C. (2001) A realistic two-dimensional model of aerosol transport and deposition in the alveolar zone of
44             the human lung. J. Aerosol Sci. 32: 1161-1174.
45       Dorries, A.  M.; Valberg, P. A. (1992) Heterogeneity of phagocytosis for inhaled versus instilled material. Am. Rev.
46             Respir. Dis. 146:831-837.
47       Driscoll, K. E.; Costa, D. L.; Hatch, G.; Henderson, R.; Oberdorster, G.; Salem, H.; Schlesinger, R. B. (2000)
48             Intratracheal instillation as an exposure technique for the evaluation of respiratory tract toxicity: uses and
49             limitations. Toxicol. Sci. 55: 24-35.
50       Falk, R.; Philipson, K.; Svartengren, M.; Jarvis, N.; Bailey, M.;  Camner, P. (1997) Clearance of particles from small
51             ciliated airways. Exp. Lung Res. 23: 495-515.
52       Falk, R.; Philipson, K.; Svartengren, M.; Bergmann, R.; Hofmann, W.; Jarvis, N.; Bailey, M.; Camner, P. (1999)
5 3             Assessment of long-term bronchiolar clearance of particles from measurements of lung retention and
54             theoretical estimates of regional deposition. Exp. Lung Res. 25: 495-516.
         April 2002                                      6-77         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Ferin, J. (1977) Effect of particle content of lung on clearance pathways. In: Sanders, C. L.; Schneider, R. P.; Dagle,
  2             G. E.; Ragan, H. A., eds. Pulmonary macrophages and epithelial cells: proceedings of the sixteenth annual
  3             Hanford biology symposium; September 1976; Richland, WA. Oak Ridge, TN: Energy Research and
  4             Development Administration; pp. 414-423. Available from: NTIS, Springfield, VA; CONF-760927.
  5             (ERDA symposium series 43).
  6       Ferin, J.; Feldstein, M. L. (1978) Pulmonary clearance and hilar lymph node content in rats after particle exposure.
  7             Environ. Res. 16: 342-352.
  8       Ferin, J.; Oberdorster, G.; Penney, D. P. (1992) Pulmonary retention of ultrafine and fine particles in rats. Am. J.
  9             Respir. Cell Mol. Biol. 6: 535-542.
10       Findeisen, W. (1935) Uber das Absetzen kleiner, in der Luft suspendierter Teilchen in der menschlichen Lunge bei
11             der Atmung [The deposition of small airborne particles in the human lung during respiration]. Pfluegers
12             Arch. Gesamte Physiol. Menschen Tiere 236: 367-379.
13       Foster, W. M.; Langenback, E.; Bergofsky, E. H. (1980) Measurement of trachea! and bronchial mucus velocities in
14             man: relation to lung clearance. J. Appl. Physiol.: Respir. Environ. Exercise Physiol. 48: 965-971.
15       Frampton, M. W.; Chalupa, D.; Morrow, P. E.; Gibb, F. R.; Oberdorster, G.; Speers, D. M.; Zareba, W.; Utell, M. J.
16             (2000) Deposition and effects of inhaled ultrafine carbon particles in healthy subjects at rest. Presented at:
17             PM2000: paniculate matter and health—the scientific basis for regulatory decision-making, specialty
18             conference & exhibition; January; Charleston, SC. Pittsburgh, PA: Air & Waste Management Association.
19       Fry, F. A.; Black, A. (1973) Regional deposition and clearance of particles in the human nose. J. Aerosol Sci.
20             4: 113-124.
21       Gehr, P.; Schurch, S.; Berthaiume, Y.; ImHof, V.; Geiser, M. (1990) Particle retention in airways by surfactant.
22             J. Aerosol Med. 3:27-43.
23       Gehr, P.; ImHof, V.; Geiser, M.; Schurch, S. (1991) The fate of particles deposited in the intrapulmonary
24             conducting airways. J. Aerosol Med. 4: 349-362.
25       Gradofi, L.; Orlicki, D. (1990) Deposition and inhaled aerosol particles in a generation of the tracheobronchial tree.
26             J. Aerosol Sci. 21:3-19.
27       Green, F. H. Y. (2000) Pulmonary responses to inhaled poorly soluble paniculate in the human. In: Gardner, D. E.,
28             ed. ILSI Risk Science Institute Workshop: The Relevance of the Rat Lung Response to Particle Overload for
29             Human Risk Assessment; March, 1998. Inhalation Toxicol. 12: 59-95.
30       Groth, M. L.; Macri, K.; Foster, W. M. (1997) Cough and mucociliary transport of airway paniculate in chronic
31             obstructive lung disease. In: Cherry, N.; Ogden, T., eds. Inhaled Particles VIII: proceedings of an
32             international symposium on inhaled particles organised by the British Occupational Hygiene Society;
33             August 1996; Cambridge, UK. Ann. Occup. Hyg. 41(suppl.): 515-521.
34       Guilmette, R. A.; Cheng, Y. S.; Griffith, W. C. (1997) Characterising the variability in adult human nasal airway
35             dimensions. In: Cherry, N.; Ogden, T., eds. Inhaled Particles VIII: proceedings of an international
36             symposium on inhaled particles organised by the British Occupational Hygiene Society; August 1996;
37             Cambridge, UK. Ann. Occup. Hyg. 41(suppl. 1): 491-496.
38       Harmsen, A. G.; Muggenburg, B. A.; Snipes, M. B.; Bice, D. E. (1985) The role of macrophages in particle
39             translocation from lungs to lymph nodes.  Science (Washington, DC) 230: 1277-1280.
40       Heistracher, T.; Hofmann, W. (1995) Physiologically realistic models of bronchial airway bifurcation. J. Aerosol
41             Sci. 26: 497-509.
42       Heistracher, T.; Hofmann, W.  (1997) Flow and deposition patterns in successive airway bifurcations. In: Cherry,
43             N.; Ogden, T., eds. Inhaled Particles VIII: proceedings of an international symposium on inhaled particles
44             organised by the British Occupational Hygiene Society; August 1996; Cambridge,  UK. Ann. Occup.  Hyg.
45             41(suppl.): 537-542.
46       Henshaw, D. L.; Fews, A. P. (1984) The microdistribution of alpha emitting particles in the human lung.
47             In: Smith, H.; Gerber, G., eds. Lung modelling for inhalation of radioactive materials: proceedings of a
48             meeting jointly organized by the Commission of the European Communities and the National Radiological
49             Protection Board; March; Oxford, United Kingdom. Luxembourg: Commission of the European
50             Communities; pp. 199-218; report no. EUR 9384 EN.
51       Heyder, J.; Rudolf,  G. (1977) Deposition of aerosol particles in the human nose. In: Walton, W. H.; McGovern, B.,
52             eds. Inhaled particles IV: proceedings of an international symposium, part 1; September 1975; Edinburgh,
53             United Kingdom. Oxford, United Kingdom: Pergamon Press, Ltd.; pp. 107-126.
54       Killer, F. C. (1991) Health implications of hygroscopic particle growth in the human respiratory tract. J. Aerosol
55             Med. 4: 1-23.


         April 2002                                     6-78        DRAFT-DO NOT  QUOTE OR CITE

-------
  1       Hofmann, W.; Bergmann, R. (1998) Predictions of particle deposition patterns in human and rat airways. Inhalation
  2             Toxicol. 10: 557-583.
  3       Hofmann, W.; Martonen, T. B.; Graham, R. C. (1989a) Predicted deposition of nonhygroscopic aerosols in the
  4             human lung as a function of subject age. J. Aerosol Med. 2: 49-68.
  5       Hofmann, W.; Koblinger, L.; Martonen, T. B. (1989b) Structural differences between human and rat lungs:
  6             implications for Monte Carlo modeling of aerosol deposition. In: Mahaffey, J. A., ed. 26th Hanford life
  7             sciences symposium, modeling for scaling to man: biology, dosimetry, and response. Health Phys.
  8             57(suppl. 1): 41-47.
  9       Hofmann, W.; Balashazy, L; Heistracher, T.; Koblinger, L. (1996) The significance of particle deposition patterns in
10             bronchial airway bifurcations for extrapolation modeling. Aerosol Sci. Technol. 25: 305-327.
11       Hofmann, W.; Bergmann, R.; Koblinger, L. (1999) Characterization of local particle deposition patterns in human
12             and rat lungs by different morphometric parameters. J. Aerosol Sci. 30: 651-667.
13       Hofmann, W.; Asgharian, B.; Bergmann, R.; Anjilvel, S.; Miller, F. J. (2000) The effect of heterogeneity of lung
14             structure on particle deposition in the rat lung. Toxicol. Sci. 53: 430-437.
15       Hsieh, T. H.; Yu, C. P. (1998) Two-phase pulmonary clearance of insoluble particles in mammalian species.
16             Inhalation Toxicol. 10: 121-130.
17       Huchon, G. J.; Montgomery, A. B.; Lipavsky, A.; Hoeffel, J. M.; Murray, J. F. (1987) Respiratory clearance of
18             aerosolized radioactive solutes of varying molecular weight. J. Nucl. Med. 28: 894-902.
19       International Commission on Radiological Protection. (1960) Report of Committee II on permissible dose for
20             internal radiation (1959). Health Phys. 3.
21       International Commission on Radiological Protection. (1979) Limits for intakes of radionuclides by workers.
22             Oxford, United Kingdom: Pergamon Press; ICRP publication 30, part 1.
23       International Commission on Radiological Protection. (1994) Human respiratory tract model for radiological
24             protection: a report of a task group of the International Commission on Radiological Protection. Oxford,
25             United Kingdom: Elsevier Science Ltd. (ICRP publication 66; Annals of the ICRP: v. 24, nos.  1-3).
26       Jaques, P. A.; Kim, C. S. (2000) Measurement of total lung deposition of inhaled ultrafine particles in healthy men
27             and women. Inhalation Toxicol. 12: 715-731.
28       John, J.; Wollmer, P.; Dahlback, M.; Luts, A.; Jonson, B. (1994) Tidal volume and alveolar clearance of insoluble
29             particles. J. Appl. Physiol.  76: 584-588.
3 0       Katz, I. M. (2001) Computer modeling of fluid dynamics and particle motion in the larynx and trachea. In:
31             Martonen, T. B., ed. Medical Applications of Computer Modeling: The Respiratory System. Boston, MA:
32             WIT Press, pp. 47-63.
33       Katz, I. M.; Martonen, T. B. (1996) Three-dimensional fluid particle trajectories in the human larynx and trachea.
34             J. Aerosol Med. 9:513-20.
35       Katz, I. M., Martonen, T. B., and Flaa, W. (1997) Three-dimensional computational study of inspiratory aerosol
36             flow through the larynx: the effect of glottal aperture modulation. J. Aerosol Sci. 28: 1073-83.
37       Katz, I. M.; Davis, B. M.; Martonen, T. B. (1999) A numerical study of particle motion within the human larynx
38             and trachea. J. Aerosol Sci. 30: 173-183.
39       Kesavanathan, J.; Swift, D.  L. (1998) Human nasal passage particle deposition: the effect of particle size, flow rate,
40             and anatomical factors. Aerosol Sci. Technol. 28: 457-463.
41       Keyhani, K.; Scherer, P. W.; Mozell, M. M. (1995) Numerical simulation of airflow in the human nasal cavity.
42             J. Biomech. Eng. 117: 429-41.
43       Kim, C. S. (1989) Aerosol deposition in the lung with obstructed airways. J. Aerosol Med. 2: 111-120.
44       Kim, C. S. (2000) Methods  of calculating lung delivery and deposition of aerosol particles. Respir. Care
45             45:695-711.
46       Kim, C. S.;  Fisher, D. M. (1999) Deposition characteristics of aerosol particles in sequentially bifurcating airway
47             models. Aerosol Sci. Technol. 31:198-220.
48       Kim, C. S.;  Garcia, L. (1991) Particle deposition in cyclic bifurcating tube flow. Aerosol Sci. Technol. 14: 302-315.
49       Kim, C. S.;  Hu, S. C. (1998) Regional deposition of inhaled particles in human lungs: comparison between men and
50             women. J. Appl. Physiol. 84: 1834-1844.
51       Kim, C. S.;  Iglesias, A. J. (1989) Deposition of inhaled particles in bifurcating airway models: I. inspiratory
52             deposition. J. Aerosol Med. 2: 1-14.
53       Kim, C. S.;  Jaques, P. A. (2000) Respiratory dose of inhaled ultrafine particles in healthy adults. Phil. Trans. Roy.
54             Soc.  London A 358: 2693-2705.
         April 2002                                      6-79          DRAFT-DO NOT QUOTE OR CITE

-------
  1       Kim, C. S.; Kang, T. C. (1997) Comparative measurement of lung deposition of inhaled fine particles in normal
  2             subjects and patients with obstructive airway disease. Am. J. Respir. Crit. Care Med. 155: 899-905.
  3       Kim, C. S.; Iglesias, A. J.; Sackner, M. A. (1987) Mucus clearance by two-phase gas-liquid flow mechanism:
  4             Asymmetric periodic flow model. J. Appl. Physiol. 62: 959-971.
  5       Kim, C. S.; Eldridge, M. A.; Garcia, L.; Wanner A. (1989) Aerosol deposition in the lung with asymmetric airways
  6             obstruction: in vivo observation. J. Appl. Physiol. 67: 2579-2585.
  7       Kim, C. S.; Hu, S. C.; DeWitt, P.; Gerrity, T. R. (1996) Assessment of regional deposition of inhaled particles in
  8             human lungs by serial bolus delivery method. J. Appl. Physiol. 81: 2203-2213.
  9       Kim, C. S.; Hu, S. C.; Jaques, P.; Ding, J.; DeWitt, P. (2000) Measurement of regional deposition dose of inhaled
10             particles in the size range of 0.04-5 /j. in human lungs. Presented at: PM2000: paniculate matter and
11             health~the scientific basis for regulatory decision-making, specialty conference & exhibition; January;
12             Charleston, SC. Pittsburgh, PA: Air & Waste Management Association.
13       Kimbell, J. S. (2001) Computational fluid dynamics of the extrathoracic airways. In: Martonen, T. B., ed. Medical
14             Applications of Computer Modeling:  The Respiratory System. Boston, MA: WIT Press; pp. 11-45.
15       Kinsara, A. A.; Tompson, R. V.; Loyalka, S. K. (1993) Computational flow and aerosol concentration profiles in
16             lung bifurcations. Health Phys. 64:  13-22.
17       Koblinger, L.; Hofmann, W. (1985) Analysis of human lung morphometric data for stochastic aerosol deposition
18             calculations. Phys. Med. Biol. 30: 541-556.
19       Koblinger, L.; Hofmann, W. (1988) Stochastic morphological model of the rat lung. Anat. Rec. 221: 533-539.
20       Koch, W.; Stober, W. (2001) A simple pulmonary retention model accounting for dissolution and
21             macrophage-mediated removal of deposited polydisperse particles. Inhalation Toxicol. 13: 129-148.
22       Kohlhaufl, M.; Brand, P.; Scheuch, G.; Meyer, T. S.; Schulz, H.; Haussinger, K.; Heyder, J. (1999) Increased fine
23             particle deposition in women with asymptomatic nonspecific airway hyperresponsiveness. Am. J. Respir.
24             Crit. Care Med. 159: 902-906.
25       Kreyling, W. G. (1992) Intracellular particle dissolution in alveolar macrophages. Environ. Health Perspect.
26             97:121-126.
27       Kreyling, W. G.;  Blanchard, J. D.; Godleski, J. J.; Haeussermann, S.; Heyder, J.; Hutzler, P.; Schulz, H.; Sweeney,
28             T. D.; Takenaka, S.; Ziesenis, A. (1999) Anatomic localization of 24- and 96-h particle retention in canine
29             airways. J. Appl. Physiol. 87: 269-284.
30       LaBelle, C. W.; Brieger, H. (1961) Patterns and mechanisms in the elimination of dust from the lung. In: Davies,
31             C. N., ed. Inhaled particles and vapours: proceedings of an international symposium; March-April 1960;
32             Oxford,  United Kingdom. New York, NY: Pergamon Press; pp. 356-368.
33       Landahl, H. D.  (1950) On the removal of air-borne droplets by the human respiratory tract: I. the lung. Bull. Math.
34             Biophys. 12: 43-56.
35       Lay, J. C.; Berry, C. R.; Kim, C. S.; Bennett, W. D. (1995) Retention of insoluble particles after local intrabronchial
36             deposition in dogs. J. Appl. Physiol. 79: 1921-1929.
37       Lay, J. C.; Bennett, W. D.; Kim, C. S.; Devlin, R. B.; Bromberg, P. A. (1998) Retention and intracellular
3 8             distribution of instilled iron oxide particles in human alveolar macrophages.  Am. J. Respir. Cell Mol. Biol.
39             18:687-695.
40       Lazaridis, M.; Broday, D. M.; Hov, 0.; Georgopoulos, P. G. (2001) Integrated exposure and dose modeling and
41             analysis  system. 3. Deposition of inhaled particles in the human respiratory tract. Environ. Sci. Technol.
42             35: 3727-3734.
43       Lee, J. W.; Goo, J. H.; Chung, M. K. (1996) Characteristics of inertial deposition in a double bifurcation. J. Aerosol
44             Sci. 27:  119-138.
45       Lehnert, B. E.;  Morrow, P. E. (1985) Association of 59iron oxide with alveolar macrophages during alveolar
46             clearance.  Exp. Lung Res. 9: 1-16.
47       Lehnert, B. E.;  Valdez, Y. E.; Bomalaski, S. H.  (1988) Analyses of particles in the lung free cell, tracheobronchial
48             lymph nodal, and pleural space compartments following their deposition in the lung as related to lung
49             clearance mechanisms. In: Dodgson, J.; McCallum, R. I.; Bailey, M. R.; Fisher, D. R., eds. Inhaled particles
50             VI:  proceedings of an international symposium and workshop on lung dosimetry; September 1985;
51             Cambridge, United Kingdom. Ann. Occup. Hyg. 32(suppl.  1): 125-140.
52       Leikauf, G.; Yeates, D. B.; Wales, K. A.; Spektor, D.; Albert, R. E.; Lippmann, M. (1981) Effects of sulfuric acid
53             aerosol on respiratory mechanics and mucociliary particle clearance in healthy nonsmoking adults. Am. Ind.
54             Hyg. Assoc. J. 42: 273-282.
         April 2002                                      6-80         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Leikauf, G. D.; Spektor, D. M; Albert, R. E.; Lippmann, M. (1984) Dose-dependent effects of submicrometer
  2             sulfuric acid aerosol on particle clearance from ciliated human lung airways. Am. Ind. Hyg. Assoc. J.
  3             45:285-292.
  4       Lennon, S.; Zhang, Z.; Lessmann, R.; Webster, S. (1998) Experiments on particle deposition in the human upper
  5             respiratory system. Aerosol Sci. Technol. 28: 464-474.
  6       Li, A.; Ahmadi, G. (1995) Computer simulation of particle deposition in the upper tracheobronchial tree. Aerosol
  7             Sci. Tech. 23:201-223.
  8       Li, W.; Perzl, M.; Heyder, J.; Langer, R.; Brain, J. D.; Englmeier, K. H.; Niven, R. W.; Edwards, D. A. (1996)
  9             Aerodynamic and aerosol particle deaggregation phenomena in model oral-pharyngeal cavities. J. Aerosol
10             Sci. 27: 1269-1286.
11       Lundborg, M.; Lind, B.; Camner, P. (1984) Ability of rabbit alveolar macrophages to dissolve metals. Exp. Lung
12             Res. 7:11-22.
13       Lundborg, M.; Eklund, A.; Lind, B.; Camner, P. (1985) Dissolution of metals by human and rabbit alveolar
14             macrophages. Br. J. Ind. Med. 42:  642-645.
15       Madl, A. K.; Wilson, D. W.; Segall, H. J.; Pinkerton, K. E. (1998) Alteration in lung particle translocation,
16             macrophage function, and microfilament arrangement in monocrotaline-treated rats. Toxicol. Appl.
17             Pharmacol.  153:28-38.
18       Martonen, T. B.; Lowe, J.  (1983) Assessment of aerosol deposition patterns in human respiratory tract casts.
19             In: Marple, V. A.; Liu, B. Y. H., eds. Aerosols in the Mining and Industrial Work Environments Vol.  1
20             Fundamentals and Status, p. 151-164. Ann Arbor, MI: Ann Arbor Science Publishers.
21       Martonen, T. B.; Zhang, Z.; Lessmann, R. (1993) Fluid dynamics of the human larynx and upper tracheobronchial
22             airways. Aerosol Sci. Technol. 19: 133-156.
23       Martonen, T. B.; Yang, Y.; Xue, Z. Q. (1994a) Influences of cartilagious rings on tracheobronchial fluid dynamics.
24             Inhalation Toxicol.  6: 185-203.
25       Martonen, T. B.; Yang, Y.; Xue, Z. Q. (1994b) Effects of carinal ridge shapes on lung airstreams. Aerosol Sci.
26             Technol. 21: 119-136.
27       Martonen, T. B.; Schroeter, J. D.; Hwang, D.; Fleming, J. S.; Conway, J. H. (2000) Human lung morphology
28             models for particle deposition studies. In: Grant, L. D., ed. PM2000: paniculate matter and health.
29             Inhalation Toxicol.  12(suppl. 4): 109-121.
30       Martonen, T. B.; Guan, X.; Schreck, R. M. (2001a) Fluid dynamics in airway bifurcations: I. Primary flows.
31             Inhalation Toxicol.  13: 261-79.
32       Martonen, T. B.; Guan, X.; Schreck, R. M. (200Ib) Fluid dynamics in airway bifurcations: II. Secondary currents.
33             Inhalation Toxicol.  13: 281-9.
34       Martonen, T. B.; Guan, X.; Schreck, R. M. (2001c) Fluid dynamics in airway bifurcations: III. Localized flow
35             conditions. Inhalation Toxicol. 13: 291-305.
36       Matsui, H.; Randell, S. H.; Peretti, S. W.; Davis, C. W.; Boucher, R. C. (1998) Coordinated clearance of periciliary
37             liquid and mucus from airway surfaces. J. Clin. Invest. 102: 1125-1131.
3 8       Medinsky, M. A.; Kampcik, S. J.  (1985) Pulmonary retention of [14C]benzo/o/pyrene in rats as influenced by the
39             amount instilled. Toxicology 35: 327-336.
40       Mercer, T. T. (1967) On the role of particle size in the dissolution of lung burdens. Health Phys. 13:  1211-1221.
41       Miller, F. J.;  Anjilvel, S.; Menache, M. G.; Asgharian, B.; Gerrity, T. R.. (1995) Dosimetric issues relating to
42             paniculate toxicity.  Inhalation Toxicol. 7: 615-632.
43       Morrison, D.; Skwarski, D.; Millar, A. M.; Adams, W.; MacNee, W. (1998) A comparison of three methods  of
44             measuring 99mTc-DTPA lung clearance and their repeatability. Eur. Respir. J. 11:  1141-1146.
45       Morrow, P. E. (1973)  Alveolar clearance of aerosols. Arch. Intern. Med. 131: 101-108.
46       Morrow, P. E. (1977)  Clearance kinetics of inhaled particles. In: Brain, J. D.; Proctor, D. F.; Reid, L. M., eds.
47             Respiratory defense mechanisms (in two parts), part II. New York, NY: Marcel Dekker, Inc.; pp. 491-543.
48             (Lenfant, C., ed. Lung biology in health and disease: v. 5).
49       Morrow, P. E. (1986)  Factors determining hygroscopic aerosol deposition in airways. Physiol. Rev. 66: 330-376.
50       Morrow, P. E. (1988)  Possible mechanisms to explain dust overloading of the lungs. Fundam. Appl. Toxicol.
51             10:369-384.
52       Morrow, P. E. (1994)  Mechanisms and significance of "particle overload." In: Mohr, U.; Dungworth, D. L.;
53             Mauderly, J. L.; Oberdorster, G., eds. Toxic and carcinogenic effects of solid particles in the respiratory tract:
54             [proceedings of the  4th international inhalation symposium]; March 1993; Hannover, Germany. Washington,
55             DC: International Life Sciences Institute Press; pp. 17-25.


         April 2002                                     6-81          DRAFT-DO NOT QUOTE OR CITE

-------
  1       Morrow, P. E.; Yu, C. P. (1993) Models of aerosol behavior in airways and alveoli. In: Moren, F.; Dolovich, M. B.;
  2             Newhouse, M. T.; Newman, S. P., eds. Aerosols in medicine: principles, diagnosis and therapy. 2nd rev. ed.
  3             Amsterdam, The Netherlands: Elsevier; pp. 157-193.
  4       Morrow, P. E.; Gibb, F. R.; Gazioglu, K. M. (1967) A study of paniculate clearance from the human lungs.
  5             Am. Rev. Respir. Dis. 96: 1209-1221.
  6       Musante, C. J.; Martonen, T. B. (1999) Predicted deposition patterns of ambient paniculate air pollutants in
  7             children's lungs under resting conditions. In: Proceedings of the third colloquium on paniculate air pollution
  8             and human health; June; Durham, NC. Irvine, CA: University of California, Air Pollution Health Effects
  9             Laboratory, p. 7-15 - 7-20.
10       Musante, C. J.; Martonen, T. B. (2000a) Computer simulations of particle deposition in the developing human lung.
11             J. Air Waste Manage. Assoc. 50: 1426-1432.
12       Musante, C. J.; Martonen, T. B. (2000b) An extrapolation model to aid in toxicological studies of paniculate air
13             pollutants. Presented at: PM2000: paniculate matter and health—the scientific basis for regulatory
14             decision-making, specialty conference & exhibition; January; Charleston, SC. Pittsburgh, PA: Air & Waste
15             Management Association.
16       Musante, C. J.; Martonen, T. B. (2000c) Paniculate matter deposition in the lungs of children and adults:
17             predictions of an age-dependent computer model. Presented at: PM2000: paniculate matter and health—the
18             scientific basis for regulatory decision-making,  specialty conference & exhibition; January; Charleston, SC.
19             Pittsburgh, PA: Air & Waste Management Association.
20       National Council on Radiation Protection and Measurements. (1997) Deposition, retention and dosimetry of inhaled
21             radioactive substances. Bethesda, MD: National Council on Radiation Protection and Measurements; report
22             no. 125.
23       Naumann, B. D.; Schlesinger, R. B. (1986) Assessment of early alveolar particle clearance and macrophage
24             function following an acute inhalation of sulfuric acid mist. Exp. Lung Res. 11: 13-33.
25       Nemmar, A.; Vanbilloen, H.; Hoylaerts, M. F.; Hoet, P. H. M.; Verbruggen, A.; Nemery, B. (2001) Passage of
26             intratracheally instilled ultrafine particles from the lung into the systemic circulation in hamster. Am. J.
27             Respir. Crit. Care Med.  164: 1665-1668.
28       Nemmar, A.; Hoet, H. M.; Vanquickenborne, B.; Dinsdale, D.; Thomeer,  M.; Hoylaerts, M. F.; Vanbilloen, H.;
29             Mortelmans, L.; Nemery, B. (2002) Passage of inhaled particles into the blood circulation in humans.
30             Circulation 105: 411-414.
31       Nikula, K. J.;  Avila, K. J.; Griffith, W. C.; Mauderly, J. L. (1997) Lung tissue responses and sites of particle
32             retention differ between rats and cynomolgus monkeys exposed chronically to diesel exhaust and coal dust.
33             Fundam. Appl. Toxicol. 37:  37-53.
34       Nikula, K. J.;  Vallyathan, V.; Green, F.  H. Y.; Hahn, F. F. (2001) Influence of exposure concentration or dose on
3 5             the distribution of paniculate material in rat and human lungs. Environ. Health Perspect. 109: 311-318.
36       Noone, P. G.; Bennett, W. D.; Regnis, J. A.; Zeman, K. L.; Carson, J. L.;  King, M.; Boucher, R. C.; Knowles, M. R.
37             (1999)  Effect of aerosolized uridine-5'-triphosphate on airway clearance with cough in patients with primary
38             ciliary dyskinesia. Am. J. Respir. Crit. Care Med.: 160: 144-149.
39       Oberdorster, G. (1993) Lung dosimetry: pulmonary clearance of inhaled particles. Aerosol Sci. Technol.
40             18:279-289.
41       Oberdorster, G.; Ferin, J.; Gelein, R.; Soderholm, S. C.; Finkelstein, J. (1992) Role of the alveolar macrophage in
42             lung injury: studies with ultrafine particles. Environ. Health Perspect. 97: 193-199.
43       Oberdorster, G.; Cox, C.; Gelein, R. (1997) Intratracheal instillation versus intratracheal inhalation of tracer
44             particles for measuring lung clearance function. Exp. Lung Res. 23: 17-34.
45       Oldham, M. J.; Mannix, R. C.; Phalen, R. F. (1997) Deposition of monodisperse particles in hollow models
46             representing adult and child-size  tracheobronchial airways. Health  Phys. 72: 827-834.
47       Passali, D.; Bianchini Ciampoli, M. (1985) Normal values of mucociliary transport time in young subjects. Int. J.
48             Pediatr. Otorhinolaryngol. 9: 151-156.
49       Fatten, J. S. (1996) Mechanisms of macromolecule absorption by the lungs. Adv. Drug Delivery Rev. 19: 3-36.
50       Peterson, B. T.; Dickerson, K. D.; James, H. L.; Miller, E. J.; McLarty, J.  W.; Holiday, D. B. (1989) Comparison of
51             three tracers for detecting lung epithelial injury in anesthetized sheep. J. Appl. Physiol. 66: 2374-2383.
52       Phalen, R. F.; Oldham, M. J. (2001) Methods for modeling particle deposition as a function of age. Respir. Physiol.
53             128: 119-130.
54       Pritchard, J. N.; Jefferies, S. J.; Black, A. (1986) Sex differences in the regional deposition of inhaled particles in
55             the 2.5-7.5 ^m size range. J. Aerosol Sci.  17: 385-389.


         April 2002                                       6-82         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Raabe, O. G.; Al-Bayati, M. A.; Teague, S. V.; Rasolt, A. (1988) Regional deposition of inhaled monodisperse,
  2             coarse, and fine aerosol particles in small laboratory animals. In: Dodgson, J.; McCallum, R. I.; Bailey,
  3             M. R.; Fischer, D. R., eds. Inhaled particles VI: proceedings of an international symposium and workshop
  4             on lung dosimetry; September 1985; Cambridge, United Kingdom. Ann. Occup. Hyg. 32(suppl. 1): 53-63.
  5       Radford, E. P.; Martell, E. A.  (1977) Polonium-210: lead-210 ratios as an index of residence time of insoluble
  6             particles from cigarette smoke in bronchial epithelium. In: Walton, W. H.; McGovern, B., eds. Inhaled
  7             particles IV: proceedings of an international symposium, part 2; September 1975; Edinburgh, United
  8             Kingdom. Oxford, United Kingdom: Pergamon Press, Ltd.; pp. 567-581.
  9       Rasmussen, T. R.; Andersen, A.; Pedersen, O. F. (2000) Particle deposition in the nose related to nasal cavity
10             geometry. Rhinology 38: 102-107.
11       Roy, M. (1989) Lung clearance modeling on the basis of physiological and biological parameters. Health Phys.
12             57(suppl. 1): 255-262.
13       Rutland, J.; Cole, P. J. (1981) Nasal mucociliary clearance and ciliary beat frequency in cystic fibrosis compared
14             with sinusitis and bronchiectasis. Thorax 36: 654-658.
15       Salma, I.; Balashazy, I.;  Winkler-Heil, R.; Hofmann, W.; Zaray, G. (2002) Effect of particle mass size distribution
16             on the deposition of aerosols in the human respiratory system. J. Aerosol Sci. 33: 119-132.
17       Sarangapani, R.; Wexler, A. S. (2000) Modeling particle deposition in extrathoracic airways. Aerosol Sci. Technol.
18             32:72-89.
19       Scheuch, G.; Stahlhofen, W. (1988) Particle deposition of inhaled aerosol boluses in the upper human airways.
20             J. Aerosol Med. 1: 29-36.
21       Scheuch, G.; Philipson, K.; Falk, R.; Svartengren, M.; Stahlhofen, W.; Camner, P. (1995) Retention of particles
22             inhaled inboli with and without induced bronchoconstriction. Exp. Lung Res. 21: 901-916.
23       Schlesinger, R. B. (1985) Clearance from the respiratory tract. Fundam. Appl. Toxicol. 5: 435-450.
24       Schlesinger, R. B. (1988) Biological disposition of airborne particles: basic principles and application to vehicular
25             emissions. In: Watson,  A. Y.; Bates, R. R.; Kennedy, D., eds. Air pollution, the automobile, and public
26             health. Washington, DC: National Academy Press; pp. 239-298.
27       Schlesinger, R. B. (1989) Deposition and clearance of inhaled particles. In: McClellan, R. O.; Henderson, R. F.,
28             eds. Concepts in inhalation toxicology. New York, NY: Hemisphere Publishing Corp.; pp. 163-192.
29       Schlesinger, R. B. (1995) Deposition and clearance of inhaled particles. In: McClellan, R. O.; Henderson, R. F.,
30             eds. Concepts in inhalation toxicology. 2nd ed. Washington, DC: Taylor & Francis; pp.  191-224.
31       Schlesinger, R. B.; Ben-Jebria, A.; Dahl, A. R.; Snipes, M. B.; Ultman, J. (1997) Disposition of inhaled toxicants.
32             In: Massaro, E. J., ed. Handbook of human toxicology. Boca Raton, FL: CRC Press; pp. 493-550.
33       Schroter, R. C.; Sudlow, M. F. (1969) Flow patterns in models of the human bronchial airways. Respir. Physiol.
34             7:341-55.
35       Segal, R. A.; Guan, X.; Shearer, M; Martonen, T. B. (2000a) Mathematical model of airflow in the lungs of children
36             I: effects of tumor sizes and locations. J. Theor. Med. 2:199-213.
37       Segal, R. A.; Martonen,  T. B.; Kim, C. S. (2000b) Comparison of computer simulations to total lung deposition to
38             human subject data in healthy test subjects. J. Air Waste Manage.  Assoc. 50: 1262-1268.
39       Smaldone, G. C.; Perry,  R. J.; Bennett, W. D.; Messina, M. S.; Zwang, J.; Ilowite, J. (1988) Interpretation of
40             "24 hour lung retention" in studies of mucociliary clearance. J. Aerosol Med. 1:  11-20.
41       Snipes, M. B. (1989) Long-term retention and clearance of particles inhaled by mammalian species. CRC Crit. Rev.
42             Toxicol. 20: 175-211.
43       Snipes, M. B.; Clem, M. F. (1981) Retention of microspheres in the rat lung after intratracheal instillation.
44             Environ. Res. 24: 33-41.
45       Snipes, M. B.; McClellan, R. O.; Mauderly, J. L.; Wolff, R. K. (1989) Retention patterns for inhaled particles in the
46             lung: comparisons between laboratory animals and humans for chronic exposures. Health Phys.
47             57(suppl. 1): 69-78.
48       Snipes, M. B.; James, A. C.; Jarabek, A. M. (1997) The 1994 ICRP66 human respiratory tract dosimetry model as a
49             tool for predicting lung burdens from exposures to environmental  aerosols. Appl. Occup. Environ. Hyg. 12:
50             547-554.
51       Stahlhofen, W.; Gebhart, J.; Rudolf, G.; Scheuch, G.; Philipson, K. (1986a) Clearance from the human airways of
52             particles of different sizes deposited from inhaled aerosol boli. In:  Aerosols: formation and reactivity,
53             proceedings of the second international aerosol conference; September; Berlin, Federal Republic of
54             Germany. Oxford, United Kingdom: Pergamon Press; pp. 192-196.
         April 2002                                      6-83          DRAFT-DO NOT QUOTE OR CITE

-------
  1       Stahlhofen, W.; Gebhart, J.; Rudolf, G.; Scheuch, G. (1986b) Measurement of lung clearance with pulses of
  2             radioactively-labelled aerosols. J. Aerosol Sci. 17: 333-336.
  3       Stanley, P. J.; Wilson, R.; Greenstone, M. A.; Mackay, I. S.; Cole, P. J. (1985) Abnormal nasal mucociliary
  4             clearance in patients with rhinitis and its relationship to concomitant chest disease. Br. J. Dis. Chest
  5             79:77-82.
  6       Stapleton, K. W.; Guentsch, E.; Hoskinson, M. K.; Finlay, W. H. (2000) On the suitability of K-e turbulence
  7             modeling for aerosol deposition in the mouth and throat: a comparison with experiment. J. Aerosol Sci.
  8             31:739-749.
  9       Stober, W.; Morrow, P. E.; Koch, W.; Morawietz, G. (1994) Alveolar clearance and retention of inhaled insoluble
10             particles in rats simulated by a model inferring macrophage particle load distributions. J. Aerosol Sci.
11             25:975-1002.
12       Strom, K. A.; Chan, T. L.; Johnson, J. T. (1988) Pulmonary retention of inhaled submicron particles in rats: diesel
13             exhaust exposures and lung retention model. In: Dodgson, J.; McCallum, R. L; Bailey, M. R.; Fischer, D. R.,
14             eds. Inhaled particles VI: proceedings of an international symposium and workshop on lung dosimetry;
15             September 1985; Cambridge, United Kingdom.  Ann. Occup. Hyg. 32(suppl. 1): 645-657.
16       Suarez, S.; Kazantseva, M.; Bhat, M.; Costa, D.; Hickey, A. J. (2001) The influence of suspension nebulization or
17             instillation on particle uptake by guinez pig alveolar macrophages. Inhalation Toxicol. 13: 773-788.
18       Subramaniam, R. P.; Richardson, R. B.; Morgan, K. T.; Guilmette, R. A.; Kimbell, J. S. (1998) Computational fluid
19             dynamics simulations of inspiratory airflow in the human nose and nasopharynx. Inhalation Toxicol.
20             10:91-120.
21       Subramaniam, R. P.; Asgharian, B.; Leendert, V.; Cassee, F. R.; Freijer, J. L; Miller, F. J.; Rombout, P. J. A. (1999)
22             A software package for multiple-path modeling of paniculate matter deposition in human and rat lungs.
23             In: Phalen, R.; Bell, Y., eds. Proceedings of Third Colloquium on Paniculate Air Pollution and Human
24             Health. APHEL report no. 99-01, pp. 7.21-7.30. Irvine, CA: Air Pollution Health Effects  Laboratory,
25             University of California.
26       Svartengren, K.; Lindestad, P.; Svartengren, M.; Philipson, K.; Bylin, G.; Camner, P. (1995) Added external
27             resistance reduces oropharyngeal deposition and increases lung deposition of aerosol particles in asthmatics.
28             Am.  J. Respir. Crit. Care Med. 152: 32-37.
29       Svartengren, K.; Philipson, K.; Svartengren, M.; Anderson, M.; Camner, P. (1996a) Tracheobronchial deposition
30             and clearance in small airways in asthmatic subjects. Eur. Respir. J. 9: 1123-1129.
31       Svartengren, K.; Ericsson, C. H.; Svartengren, M.; Mossberg, B.; Philipson, K.; Camner, P. (1996b) Deposition and
32             clearance in large and small airways in chronic bronchitis. Exp. Lung Res. 22: 555-576.
33       Svartengren, K.; Philipson, K.; Svartengren, M.; Camner, P. (1998) Effect of adrenergic stimulation on clearance
34             from small ciliated airways in healthy subjects. Exp. Lung Res. 24:  149-158.
35       Svartengren, M.; Svartengren, K.; Aghaie, F.; Philipson, K.; Camner, P. (1999) Lung deposition and extremely slow
36             inhalations of particles. Limited effect of induced airway obstruction. Exp. Lung Res. 25: 353-366.
37       Swift, D. L.; Strong, J. C. (1996) Nasal deposition of ultrafine 218Po  aerosols in human subjects. J. Aerosol Sci.
38             27: 1125-1132.
39       Takahashi,  S.; Asaho, S.; Kubota, Y.; Sato, H.; Matsuoka, O. (1987) Distribution of 198Au and 133Ba in thoracic and
40             cervical lymph nodes of the rat following the intratracheal instillation of 198Au-colloid and 133BaSO4.
41             J. Radial Res. 28:227-231.
42       Takahashi,  S.; Kubota, Y.; Hatsuno, H. (1992) Effect of size on the movement of latex particles in the respiratory
43             tract following local administration. Inhalation Toxicol. 4:  113-123.
44       Takenaka, S.; Karg, E.; Roth, C.; Schulz, H.; Ziesenis, A.; Heinzmann, U.; Schramel, P.; Heyder, J. (2001)
45             Pulmonary and systemic distribution of inhaled ultrafine silver particles in rats. Environ. Health Perspect.
46             109(suppl. 4): 547-551.
47       Toms, N.; Hasani, A.; Pavia, D.; Clarke, S. W.; Agnew, J. E. (1997) Effect of mucus hypersecretion on initial time
48             course of inert particle clearance from the lung. In: Cherry, N.; Ogden,  T., eds. Inhaled Particles VIII:
49             proceedings of an international symposium on inhaled particles organised by the British Occupational
50             Hygiene Society; August 1996; Cambridge, UK. Ann. Occup. Hyg. 41(suppl.): 509-514.
51       Tran, C. L.; Jones, A. D.; Cullen, R. T.; Donaldson, K. (1999) Exploration of the mechanisms of retention and
52             clearance of low-toxicity particles in the rat lung using a mathematical model. Inhalation Toxicol.
53             11: 1077-1108.
         April 2002                                      6-84         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Tran, C. L.; Buchanan, D.; Cullen, R. T.; Searl, A.; Jones, A. D.; Donaldson, K. (2000) Inhalation of poorly soluble
  2             particles. II. Influence of particle surface area on inflammation and clearance. Inhalation Toxicol. 12: 1113-1126.
  3       Tsuda, A.; Butler, J. P.; Fredberg, J. J. (1994a) Effects of alveolated duct structure on aerosol kinetics. I.
  4             Diffusional deposition in the absence of gravity. J. Appl. Physiol. 76: 2497-2509.
  5       Tsuda, A.; Butler, J. P.; Fredberg, J. J. (1994b) Effects of alveolated duct structure on aerosol kinetics. II.
  6             Gravitational sedimentation and inertial impaction. J. Appl. Physiol. 76: 2510-2516.
  7       Tsuda, A.; Henry, F. S.; Otani, Y.; Haber, S.; Butler, J. P. (1996) Aerosol transport and deposition in the
  8             rhythmically expanding pulmonary acinus. J. Aerosol. Med. 9: 389-408.
  9       U.S. Environmental Protection Agency. (1996) Air quality criteria for paniculate matter. Research Triangle Park,
10             NC: National Center for Environmental Assessment-RTF Office; report nos. EPA/600/P-95/00laF-cF. 3v.
11       Venkataraman, C.; Kao, A. S. (1999) Comparison of particle lung doses from the fine and coarse fractions of urban
12             PM-10 aerosols. Inhalation Toxicol. 11: 151-169.
13       Wagner, E. M; Foster, W. M. (2001) The role of the bronchial vasculature in soluble particle clearance. Environ.
14             Health Perspect. 109(suppl. 4): 563-565.
15       White, F. M. (1974) Viscous fluid flow. New York, NY: McGraw-Hill, pp. 65-72.
16       Wolff, R. K.; Henderson, R. F.; Snipes, M. B.; Griffith, W. C.; Mauderly, J. L.; Cuddihy, R. G.; McClellan, R. O.
17             (1987) Alterations in particle accumulation and clearance in lungs of rats chronically exposed to diesel
18             exhaust. Fundam. Appl. Toxicol. 9: 154-166.
19       Xu, G. B.; Yu, C. P. (1986) Effects of age on deposition of inhaled aerosols in the human lung. Aerosol Sci.
20             Technol. 5: 349-357.
21       Yeates, D. B.; Aspin, M. (1978) A mathematical description of the airways of the human lungs. Respir. Physiol.
22             32:91-104.
23       Yeates, D. B.; Aspin, N.; Levison, H.; Jones, M. T.; Bryan, A. C. (1975) Mucociliary trachea! transport rates in
24             man. J. Appl. Physiol. 19: 487-495.
25       Yeates, D. B.; Pitt, B. R.; Spektor, D. M.; Karron, G. A.; Albert, R.  E. (1981) Coordination of mucociliary transport
26             in human trachea and intrapulmonary airways. J. Appl. Physiol.: Respir. Environ. Exercise Physiol.
27             51: 1057-1064.
28       Yeh, H.-C.; Schum, G. M. (1980) Models of human lung airways and their application to inhaled particle
29             deposition. Bull. Math. Biol. 42: 461-480.
30       Yu, G.; Zhang, Z.; Lessmann, R. (1996) Computer simulation of the flow field and particle deposition by diffusion
31             in a 3-D human airway bifurcation. Aerosol Sci. Technol. 25: 338-352.
32       Yu, G.; Zhang, Z.; Lessmann, R. (1998) Fluid flow and particle diffusion in the human upper respiratory system.
33             Aerosol Sci. Technol. 28: 146-158.
34       Zhang, Z.; Martonen, T. (1997) Deposition of ultrafine aerosols  in human tracheobronchial airways. Inhalation
35             Toxicol. 9:99-110.
36       Zhang, L.; Asgharian, B.; Anjilvel, S. (1997) Inertial deposition  of particles in the human upper airway bifurcations.
37             Aerosol Sci. Tech. 26: 97-110.
3 8       Zhang, Z.; Kleinstreuer, C.; Kim, C. S. (2000) Effects of asymmetric branch flow rates on aerosol deposition in
39             bifurcating airways. J. Med. Eng. Technol. 24: 192-202.
40       Zhang, Z.; Kleinstreuer, C.; Kim, C. S. (2001) Effects of curved inlet tube on air flow and particle deposition in
41             bifurcating lung models. J. Biomechanics 34: 659-669.
42       Zhang, Z.; Kleinstreuer, C.; Kim, C. S. (2002) Cyclic micron-size particle inhalation and deposition in a triple
43             bifurcation lung airway model. J. Aerosol Sci. 33: 257-281.
         April 2002                                      6-85          DRAFT-DO NOT QUOTE OR CITE

-------
 i          7.  TOXICOLOGY OF PARTICULATE MATTER IN
 2                 HUMANS AND LABORATORY ANIMALS
 3
 4
 5      7.1  INTRODUCTION
 6           Toxicological research on ambient particulate matter (PM) is used to address several
 7      related questions, including (1) does exposure to PM at relevant ambient concentrations cause
 8      toxicological effects, (2) what mechanisms may be involved in the toxicological response to PM
 9      exposure, (3) what factors affect individual or subpopulation susceptibility to the effects of PM
10      exposure, (4) what characteristics of PM (e.g., size, composition) contribute to the observed
11      toxicity, and (5) what are the combined effects of PM and gaseous co-pollutants in producing
12      toxic responses? A variety of research approaches are used to address these questions, including
13      studies of human volunteers exposed to PM under controlled conditions; in vivo studies of
14      laboratory animals such as nonhuman primates, dogs and rodent species; and in vitro studies of
15      tissue, cellular, genetic, and biochemical systems.  Similarly, a variety of exposure conditions are
16      employed, including whole body and nose-only inhalation exposures to laboratory generated PM
17      or concentrated ambient PM, tracheal or pulmonary instillation, nasal or nasopharyngeal
18      instillation, and in vitro exposure to test materials in solution or suspension.  The various
19      research approaches are targeted to test hypotheses and, ultimately, provide a scientific basis for
20      an improved understanding  of the role of PM in producing the health effects  identified by
21      epidemiological studies.
22           Because of the sparsity of toxicological data on ambient PM at the time the previous PM
23      Air Quality Criteria Document or "PM AQCD" (U.S. Environmental Protection Agency, 1996a)
24      was completed, the discussion of respiratory effects of PM was organized into specific chemical
25      components of ambient PM or model "surrogate" particles (e.g., acid aerosols, metals, ultrafme
26      particles, bioaerosols, "other particle  matter").  In this chapter, the conclusions of the 1996 PM
27      AQCD are summarized for each of these components. Since completion of the previous
28      document, there are many new studies demonstrating the potentially toxic effects of combustion-
29      related particles. The main reason for this increased interest in combustion particles is that these
        April 2002                                7-1        DRAFT-DO NOT QUOTE OR CITE

-------
 1     particles, along with the secondary aerosols that they form, are typically the dominant sources
 2     represented in the fine fraction of ambient PM.
 3           This chapter is organized as follows.  The respiratory effects of specific components of
 4     ambient PM or surrogate particles delivered by in vivo exposures of both humans and laboratory
 5     animals are discussed first (Section 7.2), followed by cardiovascular and systemic effects of
 6     particles (Section 7.3)  and effects in laboratory animal models that mimic human disease
 7     (Section 7.4). The in vitro exposure studies are discussed next (Section 7.5) because they
 8     provide valuable information on potential hazardous constituents and mechanisms of PM injury.
 9     The remaining section on exposure studies examines the health effects of mixtures of ambient
10     PM or PM surrogates with gaseous pollutants (Section 7.6). This organization provides the
11     underlying data for evaluation in the final section of this chapter (Section 7.7), but it may fail to
12     adequately convey the  extensive and intricate linkages among the pulmonary, cardiac, and
13     nervous systems, all of which may  be involved individually and in concert to represent the effects
14     of exposure to PM.
15
16
17     7.2  RESPIRATORY EFFECTS OF PARTICULATE MATTER IN
18           HEALTHY HUMANS AND LABORATORY ANIMALS: IN VIVO
19           EXPOSURES
20           The following sections assess the respiratory effects of controlled human exposure to
21     various types of PM and also review and evaluate controlled laboratory animal toxicology
22     studies. A discussion of related in  vitro studies using animal  or human respiratory cells can be
23     found in Section 7.5. The discussion focuses on  studies published since  completion of the
24     previous 1996 PM AQCD (U.S. Environmental Protection Agency, 1996a).
25           The biological responses occurring in the respiratory tract following controlled PM
26     inhalation include changes in pulmonary inflammation and systemic effects resulting from direct
27     effects on lung tissue.  The observed responses may be dependent on the physicochemical
28     characteristics of the PM, the exposure, and the health status of the host.  Many of the responses
29     are usually seen only at the higher concentrations characteristic of occupational and  laboratory
30     animal exposures and not at (typically much lower)  ambient particle concentrations.  Moreover,
31     there are substantial differences in the inhalability and deposition profiles of PM in humans and

       April 2002                                7-2         DRAFT-DO NOT QUOTE OR CITE

-------
 1      rodents (see Chapter 6 for details). Observed responses and dose-response relationships also are
 2      very dependent on the specific biological response being measured.
 3           Paniculate matter is a broad term that encompasses a myriad of physical and chemical
 4      species, some of which have been investigated in controlled laboratory animal or human studies
 5      (see Table 7-1). However, a full discussion of all types of particles that have been studied is
 6      beyond the scope of this chapter.  Thus, specific criteria were used to select topics for
 7      presentation.  High priority was placed on studies that may (1) elucidate health effects of major
 8      common constituents of ambient PM or (2) contribute to enhanced understanding of the
 9      epidemiological studies. Most studies have been designed to address the question of biologic
10      plausibility, rather than providing dose-response or risk assessment quantification.
11           Diesel particulate matter (DPM) generally fits the criteria; however, because it is described
12      in other documents in great detail (U. S. Environmental Protection Agency, 1999; Health Effects
13      Institute, 1995), it is not covered extensively in this chapter. Particles with high inherent toxicity,
14      such as silica, that are of concern primarily because of occupational exposure, are excluded from
15      this chapter and are discussed in detail in other documents and reports (U.S.  Environmental
16      Protection Agency, 1996b; Gift and Faust, 1997; Lippmann, 2000).
17           Most of the laboratory animal studies summarized here have used high particulate mass
18      concentrations administered by inhalation or by intratracheal instillation. The studies have used
19      doses that  are generally quite high when compared to ambient levels, even when laboratory
20      animal-to-human dosimetric differences are considered.  These high doses are necessary,
21      however, in laboratory animal studies that must explore potentially toxic effects using numbers
22      of subjects (animals) that are magnitudes fewer than those used in epidemiology studies. More
23      research on particle dose extrapolation is needed, therefore, to determine species differences and
24      the importance of exercise and other factors influencing particle deposition in humans that
25      together can account for a 50-fold or more difference in dose.
26           As mentioned earlier, the data available in the previous 1996 PM AQCD were from studies
27      that investigated the respiratory effects of specific components of ambient PM or surrogate
28      particles such as sulfuric acid droplets. More recently, pulmonary effects of controlled  exposures
29      to ambient PM have been investigated by the use of particles collected from  emission bag room
30      or ambient samplers (e.g., impactors; diffusion denuders) and by the use of aerosol concentrators
31      (e.g., Sioutas et al., 1995a,b, 2000; Gordon et al., 1998; Chang et al., 2000, Kim et al., 2000a,b).

        April 2002                                 7-3         DRAFT-DO NOT QUOTE OR CITE

-------
          TABLE 7-1. TYPES OF PARTICIPATE MATTER USED IN
                      TOXICOLOGICAL STUDIES
Source
category

Particle"

Source'

Label/Date1 Description

Referenced
Concentrated Ambient Particles
ambient


ambient



ambient


ambient



ambient

CAPs


CAPs



CAPs


CAPs



CAPs

New York, NY


Boston, MA



Boston, MA


Chapel Hill, NC;
Research Triangle
Park, NC

Los Angeles, CA

Gerber concentrator; 0.2 to 1.2 ,um
MMAD;
og = 0.2to3.9
Harvard concentrator;
0.2 to 0.3 ^m MMAD;
og = 2.9

1997 Harvard concentrator;
1998 0.23 to 0.34 //m MMAD;
og = 0.2to2.9
1997 Harvard concentrator; 0.65 ,um
1998 MMAD;
og = 2.35

Harvard concentrator; PM2 5

Gordon et al.
(1998; 2000)

Goldsmith et al.
(1998);
Clarke etal.
(1999; 2000a,b)
Godleski et al.
(2000)

Ohio et al.
(2000a);
Kodavanti et al.
(2000a)
Gong et al.,
(2000)
Ambient Particulate Matter Extracts
ambient
(aqueous
extracts)

ambient
(aqueous
extracts)
ambient
(aqueous
extracts)
ambient
(aqueous
extracts)
ambient
(aqueous
extracts)
ambient
(aqueous
extracts)

ambient
(aqueous
extracts)

ambient
(aqueous
extracts)
ambient
(aqueous
extracts)
urban PM (StL)



urban PM (Ott)


urban PM (Dus)


urban PM


urban PM (WDC)


urban PM



urban PM



urban CB & UCB


urban dust


St. Louis, MO



Ottawa, ONT


Dusseldorf,
Germany

Terni, Italy


Washington, DC


Provo, UT



Provo, UT



Edinburgh, UK


NIST;
Gaithersburg, MD

SRM 1648



EHC-93; videlon filter samples, mechanically
1993 sieved (36, 56, 80, 100, 300 ^m), and
stored at -80 °C






SRM 1649


1981,1982 TSP collected on glass-fiber filters,
suspended in aqueous medium,
centrifuged, lyophilized, and
resuspended in saline.
1986, 1987, TSP and PM10 collected on glass hi-
1988 vol filters, suspended, centrifuged,
lyophilized, and resuspended in
saline.






Dong et al.,
(1996); Becker
and Soukup
(1998)
Vincent et al.
(1997; 2001)

Costa and Dreher
(1997)

Fabiani et al.
(1997)

Becker and
Soukup (1998)

Kennedy et al.
(1998); Ohio
etal. (1999a,b)

Dye et al.,
(2001); Ohio and
Devlin, (2001)

Li etal. (1996;
1997)




April 2002
7-4
DRAFT-DO NOT QUOTE OR CITE

-------
             TABLE 7-1 (cont'd).  TYPES OF PARTICIPATE MATTER USED IN
                                        TOXICOLOGICAL  STUDIES
   Source
  category
Particle'
                      Source'
                                     Label/Date1
                                                             Description
                                  Referenced
 Complex Combustion-Related Particulate Matter
 stationary     coal fly ash (CFA)

 stationary     oil fly ash (OFA)


 stationary     residual oil fly ash
              (ROFA)
residential     domestic oil fly ash
              (DOFA)

residential     wood stove

mobile        diesel exhaust particles
              (DEP)

mobile        diesel particulate matter
              (DPM)
                 U.S. power plants

                 Niagra power
                 plant

                 variable
                                                                                     Watkinson et al.
                                                                                     (1998; 2000a,b);
                                                                                     Campen et al.
                                                                                     (2000);
                                                                                     Muggenburg
                                                                                     et al. (2000)
                 home oil-burning
                 furnace

                 Durham, NC
 Laboratory-Derived Surrogate Particulate Matter
ambient
(simulated)
ambient
(simulated)
stationary
(simulated)
mobile
natural
natural
inorganic
organic
acid aerosols
(e.g., H2S04)
bioaerosols (e.g.,
lipopolysaccharide, LPS)
inorganic metal oxides
(CdO, Fe2O, MnO2,
NiSO4, TiO2, V2O5, ZnO)
diesel soot NIST; SRM 1650
Gaithersburg, MD
Mt. St. Helens ash Ritzville, WA
(MSH)
coal dust
carbon black (CB)
synthetic polymer
microspheres (SPM)
See Table 7-2
See Table 7-7
See Table 7-3





 "Particle Notes:
     1.  See Tables 7-4, 7-5, 7-6 and 7-8 for description and additional information on studies using ambient PM and PM substitutes.
     2.  See Table 7-2 for description and additional information on studies using acid aerosols.
     3.  See Table 7-3 for description and additional information on studies using metal oxides.
     4.  See Table 7-7 for description and additional information on studies using ambient bioaerosols.
     5.  For additional information on Diesel PM (DPM) or Diesel exhaust particles (DEP), see U.S. Environmental Protection Agency
        (2000) and Health Effects Institute (1995).
     6.  UCB = fine or ultrafine urban carbon black particles.
 b Source Notes:
     1.  Aerosol concentrators (e.g., Harvard;  Gerber) were used to generate CAPs.
     2.  Particle samplers (e.g., impactors, diffusion denuders) were used to collect ambient PM.
     3.  NIST = National Institute of Standards and Technology.
 'Label/Date Notes:
     •   SRM = standard reference material.
     •   EHC = Environmental Health Center in Ottawa, Canada.
        Date of particle collection, when available.
 dReference: Not an exhaustive list; see text for details.
April 2002
                                     7-5
DRAFT-DO NOT QUOTE OR CITE

-------
 1      Some ambient PM has been standardized as a reference material and compared to existing dust
 2      and soot standards [e.g., National Institutes of Standards and Technology (NIST)]. Both ambient
 3      PM and CAPs have been used to investigate effects in normal and compromised laboratory
 4      animals and humans.
 5           Particles from ambient air samplers are collected on filters or other media, stored, and
 6      resuspended in an aqueous medium for use in experimental, tracheal installation, or in vitro
 7      studies.  The in vivo and in vitro studies discussed in this chapter have almost exclusively used
 8      PM10 or PM2 5 as particle size cutoffs for studying the adverse effects of ambient PM.  Studying
 9      only particles less than a certain size is justified based upon earlier interests in setting standards
10      for PM10 and PM25. In addition, the collection of these size fractions is made easier by the
11      widespread availability of ambient sampling equipment for PM10 and PM2 5. Unfortunately, the
12      study of other important size fractions, such as the coarse fraction (PM10_2 5) and PMl 0 has been
13      largely ignored and little toxicology data are available to specifically address these potentially
14      important particle sizes. Similarly, organic compounds make up 20 to 60% of the dry fine
15      particle mass of ambient PM (Chapter 3, Section 3.2), yet very little research has addressed the
16      mechanisms by which this organic fraction contributes to the adverse effects associated with
17      acute exposure to PM. The potential contribution of organics in mutagenesis and carcinogenesis
18      has been studied, but these findings are not discussed within the context of this chapter which is
19      focused on understanding the epidemiologic evidence of increased cardiopulmonary morbidity
20      and mortality associated with acute exposure to ambient PM.
21           Particle concentrators provide a technique for exposing animals or humans by inhalation to
22      concentrated ambient particles (CAPs) that are 5 to 10-fold higher than typical ambient PM
23      levels.  The development of particle concentrators has permitted the study of true ambient PM
24      under controlled conditions. This strength is somewhat weakened by the inability of CAPs
25      studies to precisely control the mass concentration and day-to-day variability in ambient particle
26      composition.  Nonetheless, these studies are invaluable in the attempt to understand the
27      biological mechanisms responsible for the cardiopulmonary response to inhaled PM. Because
28      the composition of concentrated ambient PM varies in both time and location, a thorough
29      physical-chemical characterization is necessary to compare results among studies or even among
30      exposures within studies or to link particle composition to effect.
31

        April 2002                                  7-6         DRAFT-DO NOT QUOTE OR CITE

-------
 1      7.2.1 Ambient Combustion-Related and Surrogate Particulate Matter
 2           In vivo toxicology studies utilizing inhalation exposure as a technique for measuring the
 3      respiratory effects of ambient particles in humans and laboratory animals have been conducted
 4      with CAPs (see Table 1) and with DPM. The majority of the in vivo exposures have utilized
 5      intratracheal instillation techniques.  Discussions on the pros and cons of this technique in
 6      comparison to inhalation are covered in Chapter 6 (Section 6.5), and these issues have also been
 7      reviewed elsewhere (Driscoll et al., 2000; Oberdorster et al.,  1997; Osier and Oberdorster, 1997).
 8      In most of the studies, PM samples were collected on filters, resuspended in a vehicle (usually
 9      saline), and a small volume of the suspension was instilled intratracheally into the animals. The
10      physiochemical characteristics of PM may be altered by deposition on a filter and resuspension in
11      an aqueous medium. In addition, the doses used in these instillation studies are generally high
12      compared to ambient concentrations, even when laboratory animal-to-human dosimetric
13      differences are considered.  Therefore, in terms of direct extrapolation to humans in ambient
14      exposure scenarios, greater importance should be placed on inhalation studies.  However,
15      delivery of PM by instillation has the advantages that much less material is needed and that the
16      delivered dose can be determined directly without extrapolating from estimates of lung
17      deposition.  Instillation studies have proven valuable in comparing the effects of different types
18      of PM and for investigating some of the mechanisms by which particles may cause lung injury
19      and inflammation. Tables 7-2, 7-3, and 7-4 outline studies in which various biological endpoints
20      were measured following exposures to CAPs, ambient PM extracts, complex combustion-related
21      PM, or laboratory-derived surrogate PM, respectively.
22           There were only limited data available from human studies or laboratory animal studies on
23      ultrafine particles and even less on coarse particles at the time of the release of the previous
24      criteria document (U.S. Environmental Protection Agency, 1996a).  In vitro studies have shown
25      that ultrafine particles have the capacity to cause injury to cells of the respiratory tract. High
26      levels of ultrafine particles, as metal or polymer "fume," are associated with toxic respiratory
27      responses in humans and other mammals. Such exposures are associated with cough, dyspnea,
28      pulmonary edema, and acute inflammation.  At concentrations less than 50 //g/m3, freshly
29      generated insoluble ultrafine PTFE fume particles can be severely toxic to the lung. However, it
30      was not clear what role in the observed effects was played by fume gases which adhered to the
31      particles. Newer data from controlled  exposures have demonstrated that particle composition, in
        April 2002                                 7-7         DRAFT-DO NOT QUOTE OR CITE

-------
                TABLE 7-2. RESPIRATORY EFFECTS OF AMBIENT PARTICIPATE MATTER
13.
to
o
o
to
















1
oo




o
W
p>
H
6
O


H
O
Cj
O
w
L-LJ
o

o
1 — I
H
W

Species, Gender,
Strain, Age, etc.

Rats, male S-D
200-225 g,
control and
SO2-treated


Rats, male S-D
60 days

Humans, healthy
nonsmokers;
21 M, 3F;
26.4±2.2yrold
Rats, S-D
60 days



Humans, healthy
nonsmokers;
18 to 40 yr old
Mongrel dogs,
some with balloon
occluded LAD
coronary artery
n= 14
Humans, healthy;
n=4, 19-41 yr old
Rats, male F 344
Hamsters, male,

8-mo-old Bi TO-2
Rats, male, 90 to
100-day-old S-D,
with or without
SO2-induced
bronchitis

Rats, Wis
(HAN strain)



Particle"

Concentrated
ambient particles
(CAPs)
(Boston)


Provo, UT,
TSP filters
(10 years old)
Provo, UT,
PM10 filters
(10 years old)

Provo, UT,
TSP filters
(10 years old),
soluble and
insoluble extracts
CAPs
(Chapel Hill)

CAPs
(Boston)



CAPs
(LA)
CAPs
(NY)


CAPs
(RTF)




Ambient PM
Edinburgh, CB,
CB Ultrafine
(UCB)

Exposure
Technique

Inhalation;
Harvard/EPA
fine particle
concentrator;
animals restrained
in chamber
Intratracheal
instillation

Intrabronchial
instillation


Intratracheal
instillation



Inhalation


Inhalation via
tracheostomy



Inhalation

Inhalation



Inhalation





Intratracheal
instillation



Concentration

206,733, 607 //g/m3 for
Days 1-3; 29 °C,
59% RH



0.25, 1.0, 2.5, 5.0 mg of
PM extract in 0.3 mL
saline
500 Mg of PM extract in
10 mL saline


100-1000//gofPM
extract in 0.5 mL saline



23.1 to 311.1 //g/m3


69-828 //g/m3




148-246 Mg/m3

132 to 919 //g/m3



650 Mg/m3





50-125 //gin 0.2 mL




Particle Size Exposure Duration Effect of Particles

0.18//m 5 h/day for 3 days PEF and TV increased in CAPS exposed animals.
og = 2.9 Increased protein and % neutrophils and
lymphocytes in lavage fluid after CAPS exposure.
Responses were greater in SO2-bronchitis animals.
No changes in LDH. No deaths occurred.

N/A 24 h Inflammation (PMN) and pulmonary injury was
produced by particles collected while the steel
mill was in operation
N/A 24 h BAL Inflammation (PMN) and pulmonary injury was
produced by particles collected while the steel
mill was in operation

N/A 24 h Inflammation (PMN) and lavage fluid protein was
greater with the soluble fraction containing more
metal (Zn, Fe, Cu).


0.65 /mi 2 h; analysis at 18 h Increased BAL neutrophils in both bronchial and
og = 2.35 alveolar fractions

0.23 to 0.34 /mi 6 h/day x 3 days Decreased respiratory rate and increased lavage fluid
og = 0.2 to 2.9 neutrophils in normal dogs.



PM2 5 2 h No significant changes in lung function, symptoms,
SaO2, or Holter ECGs were observed.
0.2 to 1.2 /mi Ix3hor3x6h No inflammatory responses, no cell damage
og = 0.2 to 3.9 responses, no PFT changes.


6 h/day x 2-3 days No significant changes in healthy rats; increased
BALF protein and neutrophil influx in bronchitic
rats; responses were variable between exposure
regimens.


PM10 Sacrificed at 6 h Increased PMN, protein, and LDH following PM10;
CB = (200-500 nm) greater response with ultrafme CB but not CB;
UCB = 20 nm decreased GSH level in BAL; free radical activity
(deplete supercoil DNA); leukocytes from treated
animals produced greater NO and TNF.
Reference

Clarke et al.
(1999)




Dye etal. (2001)


Ohio and Devlin
(2001)


Ohio et al.
(1999a)



Ohio et al.
(2000a)

Godleski et al.
(2000)



Gong et al.
(2000)
Gordon et al.
(2000)


Kodavanti et al.
(2000a)




Li etal. (1996,
1997)



"See Table 1 for details

-------
>
to
o
o
to
                 TABLE 7-3. RESPIRATORY EFFECTS OF COMPLEX COMBUSTION-RELATED PARTICIPATE MATTER
Species, Gender, Strain,
Age, etc.
Particle"
Exposure
Technique
Concentration
Particle Size
Exposure
Duration
Effect of Particles
Reference
          Hamsters, Syrian golden,
          male, 90-125g
Kuwaiti oil fire  Intratracheal
particles;        instillation
urban particles
from St. Louis,
MO
                   0.15, 0.75, and 3.75  Oil fire particles:       Sacrificed 1 and
                   mg/lOOg            -c 3.5 ,um, 10 days of    7 days post
                                      24-h samples          instillation
                                                        Increases in PMN, AM, albumin, LDH,
                                                        myeloperoxidase, and
                                                        p-N-acetylglucosaminidase;
                                                        acute toxicity of the particles found in the
                                                        smoke from the Kuwaiti oil fires is
                                                        comparable to that of urban particles.
                                                        Brain etal. (1998)
Mice, female, NMRI,
28-32g
















Rats, male, S-D,
60 days old





CFA Intratracheal
CMP instillation
we















Emission Intratracheal
source PM instillation
(ROFA,
DOFA, CFA)
Ambient
airshed PM
ROFA
CMP: 20 Mg N/A
arsenic/kg, or CMP:
100 mg
particles/kg,
WC alone
(100 mg/kg), CFA
alone (100 mg/kg
[i.e., 20 //g
arsenic/kg]), CMP
mixed with WC
(CMP, 13.6 mg/kg
[(i.e., 20 //g
arsenic/kg]), WC
(86. 4 mg/kg) and
Ca3(AsO4)2 mixed
with WC (20 ,ug
arsenic/kg), WC
(100 mg/kg)
Total mass: Emission PM:
2.5mg/rat 1. 78-4.17 ^m

Total transition Ambient PM:
metal: 46 Mg/rat 3. 27-4.09 ^m


1, 5, 30 days
posttreatment,
lavage for total
protein content,
inflammatory
cell number and
type, and TNF-
a production
particle
retention








Analysis at 24
and 96 h
following
instillation



          Rats, male WISTAR
          Bor:  WISW strain
Coal oil fly ash
Inhalation
(chamber)
0, 11, 32, and
103 mg/m3
1.9-2.6 Mm
og= 1.6-1.8
6 h/day,
5days/week,
4 weeks
                                                                                                                            Mild inflammation for WC; Ca3(AsO4)2       Broeckaert et al.
                                                                                                                            caused significant inflammation;            (1997)
                                                                                                                            CMP caused severe but transient
                                                                                                                            inflammation; CFA caused persistent
                                                                                                                            alveolitis; cytokine production was
                                                                                                                            upregulated in WC- and Ca3(AsO4) treated
                                                                                                                            animals after 6 and 30 days, respectively;
                                                                                                                            a 90% inhibition of TNF-a production still
                                                                                                                            was still observed at Day 30 after
                                                                                                                            administration of CMP and CFA;
                                                                                                                            a significant fraction persisted (10-15% of
                                                                                                                            the arsenic administered) in the lung of
                                                                                                                            CMP- and CFA-treated mice at Day 30.
                                                                                                                            Suppression of TNF-a production is
                                                                                                                            dependent on the slow elimination of the
                                                                                                                            particles and their metal content from the
                                                                                                                            lung
Increases in PMNs, albumin, LDH, PMN,    Costa and Dreher
and eosinophils following exposure to        (1997)
emission and ambient particles;
induction of injury by emission and
ambient PM samples is determined
primarily by constituent metals and their
bioavailability.

At the highest concentration, type II cell      Dormans et al.
proliferation and mild fibrosis occurred and   (1999)
increased perivascular lymphocytes were
seen. The main changes at the lowest
concentration were particle accumulation in
AM and mediastinal lymph nodes.
Lymphoid hyperplasia observed at all
concentrations. Effects increased with
exposure duration.

-------
TABLE 7-3 (cont'd). RESPIRATORY EFFECTS OF COMPLEX COMBUSTION-RELATED PARTICULATE MATTER
^
to
o
o
to















1
o




o

J>
"Tl
H
6
o

o
^^
H
O
CH
O
, 1
a
L-LJ
o
7^
o
H
w



Species, Gender, Strain,
Age, etc.

Rats, male, S-D,
60 days old

Rats, S-D, 5-day-old

Rats, male, S-D rats
60 days old

Rats, male, S-D,
5-day-old



Rats, male, S-D,
60 days old








Mice, female,
Balb/cJ7-15 weeks


Mice, female,
7-week-old Balb/cJ
(16-21 g)


Rats, male, S-D



Mice, normal and Hp,
105 days old




Mice, BALB/C,
2-day-old, sensitized to
ovalbumin (OVA)
Particle"

ROFA


ROFA

#6 ROFA,
volcanic ash

LowS
#6 ROFA,

volcanic ash
saline
Two ROFA
samples
Rl had 2x
saline-
leachable
sulfate, Ni, and
Vand40xFe
as R2; R2 had
3 1 x higher Zn

ROFA



ROFA lo-S
residual oil



ROFA



ROFA





Aerosolized
ROFA leachate

Exposure
Technique

Intratracheal
instillation

Intratracheal
Instillation
Intratracheal
Instillation

Intratracheal
Instillation



Intratracheal
instillation








Intratracheal
instillation


Inhalation and
intratracheal
instillation
challenge with
OVA
Intratracheal
instillation


Intratracheal
instillation




Nose-only
inhalation

Concentration Particle Size

8.33 mg/mL 1 .95 Mm MMAD
0.3 mL/rat

500 Mg/rat 1 .95 Mm MMAD

0.3,1.7 1.95 Mm
8.3 mg/mL og = 2.19
8.3 mg/mL 1.4 Mm
0.3,1.7, 1.95 Mm
8.3mg/kgBW og=1.95
in saline 1 .4 Mm
8.3 mg/kg BW
1 mL/kg BW
2.5mgin0.3mL Rl: 1.88 Mm,
MMAD
R2: 2.03 Mm,
MMAD






60 Mg in 50 ML < 2.5
(dose 3mg/kg)


158±3mg/m3 PM25 sample




500 Mg/animal 3. 6 Mm



50 Mg 1.95 Mm





50 mg/mL N/A


Exposure
Duration

Analysis at
24 and 96 h

24h

24 h


24 h




Analysis at
4 days








Analysis at
1-15 days after
instillation

1, 3, 8, 15 days
after
instillation


Analyzed
4 and 96 h
postexposure

Analysis at
24 h




30min


Effect of Particles

Increased PMNs, protein, LDH at both time points;
bioavailable metals were causal constituents of
pulmonary injury.
Increased neutrophilic inflammation was inhibited
by DMTU treatment, indicating role of ROS.
Plasma fibrinogen elevated after ROFA instillation
but not volcanic ash

Increased WBC count in ROFA-exposed rats
plasma fibrinogen increased 86% in ROFA rats at
highest concentration.


Four of the 24 animals treated with R2 or R2s
(supernatant) died; none in Rl s treated animals;
more AM, PMN, eosinophils protein, and LDH in
R2 and R2s animals; more focal alveolar lesions,
thickened alveolar septae, hyperplasia of type II
cells, alveolar fibrosis in R2 and R2s animals;
baseline pulmonary function and airway
hyperreactivity were worse in R2 and R2s groups.


ROFA caused increases in eosinophils, IL-4 and
IL-5 and airway responsiveness in ovalbumin-
sensitized and challenged mice.

Increased BAL protein and LDH at 1 and 3 days
but not at 15 days postexposure. Combined OVA
and ROFA challenge increased all damage markers
and enhanced allergen sensitization. Increased
methacholine response after ROFA.
Ferritin and transferrin were elevated; greatest
increase in ferritin, lactoferrin, transferrin occurred
24 h postexposure.

Diminished lung injury (e.g., decreased lavage
fluid ascorbate, protein, lactate dehydrogenase,
inflammatory cells, cytokines) in Hp mice lacking
transferrin; associated with increased metal storage
and transport proteins.

Increased airway response to methylcholine and
to OVA in ROFA exposed mice; increased airway
inflammation also.
Reference

Dreher et al.
(1997)

Dye etal. (1997)

Gardner et al.
(2000)

Gardner et al.
(2000)



Gavett et al.
(1997)








Gavett et al.
(1999)


Gavett et al.
(1999)



Ohio et al.
(1998b)


Ohio et al.
(2000b)




Hamada et al.
(1999)


-------
>
to
o
o
to
           TABLE 7-3 (cont'd).  RESPIRATORY EFFECTS OF COMPLEX COMBUSTION-RELATED PARTICIPATE MATTER
Species, Gender, Strain,
Age, etc. Particle*
Rats, male, S-D, ROFA
60 days old
Rats, S-D, 250 g FOFA
MCT




Exposure
Technique
Intratracheal
instillation
Inhalation





Concentration
l.OmginO.5 mL
saline
580 ± 110Mg/m3





Particle Size
1.95 Mm

2.06 Mm MMAD
og=1.57




Exposure
Duration
Analysis at
24 h
6 h/day for
3 days




Effect of Particles
Increased PMNs, protein.

Death occurred only in MCT rats exposed to
ROFA. Neutrophils in lavage fluid were
increased significantly in MCT rats exposed to
ROFA versus filtered air. MIP-2 mRNA
expression in lavage cells was induced in normal
animals exposed to fly ash.
Reference
Kadiiska et al.
(1997)
Killingsworth
etal. (1997)




         Rats, male, S-D and        ROFA         Intratracheal
         F-344 (60 days old)                      instillation
         Rats, male, S-D, WIS,       ROFA         Intratracheal
         and F-344 (60 days old)                   instillation
1.95 Mm          Sacrificed at     Increase in neutrophils in both S-D and F-344
og = 2.14         24 h            rats; a time-dependent increase in eosinophils
                                occurred in S-D rats but not in F-344 rats.

1.95 Mm          Sacrificed at 6,   Inflammatory cell infiltration, as well as alveolar,
og = 2.14         24, 48, and      airway, and interstitial thickening in all three rat
                 72 h; 1, 3, and    strains; a sporadic incidence of focal alveolar
                 12 weeks        fibrosis in S-D rats, but not in WIS and F-344
                                rats; cellular fibronectin (cFn)
                                mRNA isoforms EIIIA(+) were up-regulated in S-
                                D and WIS rats but not in F-344 rats. Fn mRNA
                                expression by macrophage, alveolar and airway
                                epithelium,  and within fibrotic areas in S-D rats;
                                increased presence of Fn EIIIA(+) protein in the
                                areas of fibrotic injury and basally to the airway
                                epithelium.
Kodavanti et al.
(1996)
Kodavanti et al.
(1997a)
Rats, male, S-D,
60 days old






Rats, male, S-D,
60 days old





ROFA Intratracheal
instillation
Fe2(S04)3,
VS04,
NiSO4



10 Intratracheal
compositionally instillation
different ROFA
particles from a
Boston power
plant

8.33 mg/kg 1.95 Mm
og = 2.14
ROFA-equivalent
dose of metals




0.833,3.33,8.3 1.99-2.59 Mm
mg/kg MMAD





Analysis at 3,
24, and 96 h,
postinstillation





Sacrificed at
24 h





ROFA-induced pathology lesions were as severe
as those caused by Ni. Metal mixture caused less
injury than ROFA or Ni alone; Fe was less
pathogenic. Cytokine and adhesion molecule
gene expression occurred as early as 3 h after
exposure. V-induced gene expression was
transient but Ni caused persistent expression and
injury.
ROFA-induced increases in BAL protein and
LDFI, but not PMN, were associated with water-
leachable total metal, Ni, Fe, and S; BALF
neutrophilic inflammation was correlated with V
but not Ni or S. Chemiluminescence signals in
vitro (AM) were greatest with ROFA containing
soluble V and less with Ni plus V.
Kodavanti et al.
(1997b)






Kodavanti et al.
(1998a)






-------
           TABLE 7-3 (cont'd). RESPIRATORY EFFECTS OF COMPLEX COMBUSTION-RELATED PARTICIPATE MATTER
 to
 o
 o
 to
 to
 H
 6
 o
 o
 H
O
o
HH
H
W
Species, Gender, Strain,
Age, etc. Particle"
Rats, male, S-D ROFA
60-day-old treated with
MCT (60 mg/kg)




Rats, male, WKY and ROFA
SH, 11-13 weeks old

Exposure
Technique
Intratracheal
instillation;

Nose-only
inhalation


Nose-only
Inhalation

Concentration
0, 0.83,
3.3 mg/kg

15 mg/m3



15 mg/m3


Exposure
Particle Size Duration
1.95,um 24-96 h
og = 2.14
6 h/day for
3 days analysis
at 0 or 18 h


1 .95 ^im 6 h/day x 3 day,
og = 2. 14 analysis at 0 or
18h
Effect of Particles
Both IT and IN rats showed inflammatory
responses (IL-6, MIP-2 genes upregulated).
58% of IT rats exposed to ROFA died within 96 h.
No mortality occurred by inhalation. ROFA
exacerbated lung lesions (edema, inflammation,
alveolar thickening) and gene expression in MCT
rats.
More pulmonary injury in SH rats. Increased
RBCs in BALF of SH rats. ROFA increased
airway reactivity to Ach in both SH and WKY
Reference
Kodavanti et al.
(1999)





Kodavanti et al.
(2000b)

          Rats, male, WKY and
          SH, 11-13 weeks old
                                                                               rats. Increased protein, albumin, and LDH in
                                                                               BALF after ROFA exposure (SH>WKY).
                                                                               Increased oxidative stress in SH rats. SH rats
                                                                               failed to increase glutathione. Inflammatory
                                                                               cytokine gene expression increased in both SH and
                                                                               WKY rats.

ROFA        Intratracheal       3.33                 1.95,um      1 and 4 days;     Increased BALF protein and LDH alveotis with
             Instillation         mg/mL/kg           og = 2.14      post instillation   macrophage accumulation in alveoli; increased
                                                                analysis at 6 or   neutrophils in BAL. Increased pulmonary protein
VSO4,                         1.5|/molkg                        24 h            leakage and inflammation in SH rats. Effects of
NiSO4, or                                                                       metal constituents of ROFA were strain specific;
saline                                                                          vanadium caused pulmonary injury only in WKY
                                                                               rats; nickel was toxic in both SH and WKY rats.
Kodavanti et al.
(2001)
Rats, Brown Norway
Rats, male, S-D,
60-day-old
Rats, male, S-D,
60-day-old
Rats, male, S-D;
60 days old
Rats, male, S-D,
60-day-old
ROFA
#6 ROFA from
Florida
NC ROFA;
Domestic oil
fly ash
#6 ROFA
(Florida)
NiSO4
VS04
ROFA
Intratracheal
instillation
Intratracheal
instillation
Intratracheal
instillation
Intratracheal
instillation
Intratracheal
instillation
200 Mg N/A
100 Mg
1000 Mg in 0.5 ml 1.95 ±0. 18 ^m
1000 Mg in
0.5 mL saline
3.3 mg/ml/kg; 1 .9 ,um
ROFA equivalent og = 2. 14
dose of metals
400-1000 Mg/mL N/A
N/A ROFA enhanced the response to house dust mite
(HDM) antigen challenge. Eosinophil numbers,
LDH, BAL protein, and IL-10 were increased with
ROFA + HDM versus HDM alone.
15 min to 24 h Production of acetaldehyde increased at 2 h
postinstillation.
15 min to 24 h ROFA induced production of acetaldehyde with a
peak at about 2 h. No acetaldehyde was seen in
plasma at any time. DOFA increased
acetaldehyde, as did V and Fe.
3 or 24 h Inflammatory and stress responses were
upregulated; the numbers of genes upregulated
were correlated with metal type and ROFA
12 h post-IT ROFA increased PGE2 via cycloxygenase
expression.
Lambert et al.
(1999)
Madden et al.
(1999)
Madden et al.
(1999)
Nadadur et al.
(2000); Nadadur
and Kodavanti
(2002)
Samet et al.
(2000)

-------
         TABLE 7-3 (cont'd).  RESPIRATORY EFFECTS OF COMPLEX COMBUSTION-RELATED PARTICIPATE MATTER
^
to
o
o
to







Species, Gender, Strain,
Age, etc.

Rats, male, S-D,
60-day-old

Rats, male, S-D;
60-day-old; WKY and
SH; cold-stressed SH,
ozone-exposed SH, and
MCT-treated SH
Particle"

LoS,
#6 ROFA

Ottawa dust,
ROFA, and
volcanic ash


Exposure
Technique

Intratracheal
instillation

Intratracheal
instillation, nose-
only inhalation


Concentration

500 Mg in 0.5 ml
saline

Dose: IT 0, 0.25,
1.0, and
2.5 mg/rat; INH
15 mg/m3

Exposure
Particle Size Duration

3.6 /j,m 1, 4, or24h


1.95 tan 6h/dayfor
3 -day
inhalation;
instillation -
96 h post-IT
Effect of Particles

Mild and variable inflammation at 4 h;
no pronounced inflammation until 24 h when there
were marked increases in P-Tyr and P-MARKS.
IT ROFA caused acute and dose-related increase
in pulmonary inflammation; no effect of volcanic
ash.


Reference

Silbajoris et al.
(2000)

Watkinson et al.
(2000a,b)



H
6
o

o
H
O
        "See Table 1 for details (CFA = Coal fly ash; CMP = Copper smelter dust; WC = Tungsten carbide; MCT = Monocrotaline; DOFA = Fly ash from a domestic oil-burning furnace; Fe2(SO4) = Iron

        sulfate; V SO4 = Vanadium sulfate; NiSO4 = Nickel sulfate; LoS = low sulfur)
o
HH
H
W

-------
>
to
o
o
to
                         TABLE 7-4.  RESPIRATORY EFFECTS OF SURROGATE PARTICIPATE MATTER
Species, Gender, Strain,
Age, etc.
Hamsters, Syrian golden
900 male, 900 female,
4-wks-old

Mice, C57B1/6J






Rats, male, F-344
200-230 g




Mice, male, C57BL/6J,
8 weeks and 8-mo-old



Rats, male, S-D,
MCT-treated

Rats, male, S-D
(200g)




Mice, male, Swiss-Webster,
6-8 weeks old (A/J, AKR/J,
B6C3F1/J, BALB/cJ,
C3H/HeJ-C3, C3HeOuJ,
CSTBL/6J-B6, SJL/J,
SWR/J, 129/J) strains
raised in a pathogen free
laboratory
Particle"
Toner
(carbon)
TiO2
Silica
PTFE
TiO2





PTFE Fumes





PTFE Fumes




Fluorescent
microspheres

Diesel,
SiO2,
carbon black



Carbon black
Regal 660

Carbon-
associated
so4-


Exposure
Technique
Nose-only
inhalation


Inhalation






Whole body
inhalation




Whole body
inhalation



Inhalation


Intratracheal
instillation




Nose only
inhalation






Concentration
1.5, 6.0, or
24 mg/m3
40 mg/m3
3 mg/m3
PTFE:
1.25, 2.5, or
5xl05
particles/cc
TiO2-F: 10 mg/m3
NiO: 5 mg/m3
Ni3S2: 0.5 mg/m3
1, 2.5, or 5 x 105
particles/cm3




1, 2.5, or 5 x 105
particles/cm3



3. 85 ±0.81
mg/m3

1 mg in 0.4 mL.





10 mg/m3
285 Mg/m3






Particle Size
4.0 Mm

1.1 A^m
1.4 ,um
PTFE: 18nm
TiO2-F: 200 nm
TiO2-D: 10 nm




18 nm





18 nm




1.38 ±0.10 Ann
og = 1.8 ±0.28

DEP Collected as
TSP - disaggregated
in solution by
sonication (20 nm);
SiO2 (7 nm);
carbon black
0.29 Aim
± 2.7 Aim






Exposure
Duration
3, 9, 15 mo
6 h/day
5days/week

30 min or
6 h/day,
5days/week,
6 mo



1 5 min,
analysis 4 h
postexposure



30-min
exposure,
analysis 6 h
following
exposure
3 h/day
x 3 days

Necropsy at 2,
7, 21, 42, and
84 days
postinstillation


4h







Effect of Particles
Retention increased with increased exposure.
Clearance halftimes retarded (males)


Effects on the epithelium caused by direct
interactions with particles, not a result of
macrophage-derived mediators, and suggest
a more significant role in the overall pulmonary
response than previously suspected; type II cell
growth factor production may be significant in
the pathogenesis of pulmonary fibrosis.
Increased PMN, mRNA of MnSOD and MT,
IL-la, IL-lp, IL-6, MIP-2, TNF-a mRNA of
MT and IL-6 expressed around all airways and
interstitial regions; PMN expressed IL-6, MT,
and TNF-a; AM and epithelial cells were
actively involved.
Increased PMN, lymphocytes, and protein
levels in old mice over young mice; increased
TNF-a mRNA in old mice over young mice;
no difference in LDH and p-Glucuronidase.

Monocrotaline-treated animals contained fewer
microspheres in their macrophages, probably
because of impaired chemotaxis.
Amorphous SiO2 increased permeability, and
neutrophillic inflammation. Carbon black
and DEP translocated to interstitum and lymph
nodes by 12 weeks.


Differences in inflammatory responses
(PMN) across strains. Appears to be genetic
component to the susceptibility.





Reference
Creutzenberg et al.
(1998)


Finkelstein et al.
(1997)





Johnston et al.
(1996)




Johnston et al.
(1998)



Madletal. (1998)


Murphy etal. (1998)





Ohtsuka et al.
2000a,b






       aSee Table 1 for details (PTFE = polytetrafluoroethylene; TiO2 = titanium oxide; SiO2 = silicon dioxide)

-------
 1      addition to particle size, may be responsible for the adverse health effects associated with
 2      ambient PM exposures.
 3           Toxicologic studies of other particulate matter species also were discussed in the previous
 4      criteria document (U.S. Environmental Protection Agency, 1996a). These studies included
 5      exposures to fly ash, volcanic ash, coal dust, carbon black, and miscellaneous other particles,
 6      either alone or in mixture.  Some of the particles discussed were considered to be models of
 7      "nuisance" or "inert" dusts (i.e., those having low intrinsic toxicity) and were used in instillation
 8      studies to delineate nonspecific particle effects from effects of known toxicants. A number of
 9      studies on "other PM" examined effects of up to 50,000 //g/m3 of respirable particles with
10      inherently low toxicity. Although there was no mortality, some mild pulmonary function
11      changes after exposure to 5,000 to 10,000 //g/m3 of inert particles were observed in rats and
12      guinea pigs.  Lung morphology studies revealed focal inflammatory responses, some  epithelial
13      hyperplasia, and fibrotic responses after exposure  to >5,000 //g/m3. Changes in macrophage
14      clearance after exposure to >10,000 //g/m3 were equivocal (no host defense effects).  In studies of
15      mixtures of particles and other pollutants, effects were variable depending on the toxicity of the
16      associated pollutant.  In humans, co-exposure to carbon particles appeared to increase responses
17      to formaldehyde but not to acid aerosol. None of the "other" particles mentioned above are
18      present in ambient air in more than trace quantities.  Thus, it was concluded that the relevance of
19      any of these studies to standard setting for ambient PM may be extremely limited (see Chapter 6,
20      Section 4, Particle Overload).
21
22      7.2.1.1  Ambient Particulate Matter
23           Studies that examined the acute effects of intratracheal instillation of ambient PM obtained
24      from specific ambient sources have  shown clearly that PM can cause lung inflammation and
25      injury. Costa and Dreher (1997) showed that instillation of relatively high concentrations of PM
26      samples from three emission sources (two oil and  one coal fly ash) and four ambient airsheds (St.
27      Louis, MO; Washington, DC; Dusseldorf, Germany; and Ottawa, Canada) resulted in increases in
28      lung polymorphonuclear leucocytes (PMNs) and eosinophils in rats 24 h after instillation.
29      Biomarkers of permeability (total protein and albumin) and cellular injury (LDH) also were
30      increased. This study demonstrated that the lung dose of bioavailable transition metal, not
31      instilled PM mass, was the primary determinant of the acute inflammatory response.  Kennedy et

        April 2002                                 7-15        DRAFT-DO NOT QUOTE OR CITE

-------
 1      al. (1998) reported a similar dose-dependent inflammation (i.e., increase in protein and PMN in
 2      lavage fluid, proliferation of bronchiolar epithelium, and intraalveolar hemorrhage) in rats
 3      instilled with water-extracted particles (TSP) collected in Provo, UT.  This study also indicated
 4      that the metal constituent, in this case PM-associated Cu, was a plausible cause of the outcome.
 5      Likewise, instillation of ambient PM10 collected in Edinburgh, Scotland, also caused pulmonary
 6      injury and inflammation in rats (Li et al., 1996, 1997). Brain et al. (1998) examined the effects
 7      of instillation of particles that resulted from the Kuwaiti oil fires in 1991 compared to urban
 8      particulate matter collected in St. Louis (NIST SRM 1648, collected in a bag house in early
 9      1980s) and  showed that on an equal mass basis, the acute toxicity of the Kuwaiti oil fire particles
10      was similar to that of urban particles collected in the United States.
11           Toxicological studies of ambient PM collected around Provo, UT (Utah Valley) in the late
12      1980s are particularly interesting (Ohio and Devlin, 2001; Dye et al., 2001; Wu et al., 2001;
13      Soukup et al., 2000; Frampton et al., 1999).  Epidemiological studies by Pope et al. (1989, 1991)
14      and reported in the previous PM AQCD (U.S. Environmental Protection Agency,  1996a) showed
15      that closure of an open-hearth steel mill  over the winter of 1987 was associated with reductions
16      in hospital admissions for respiratory diseases (see Chapter 8 for details of the epidemiology
17      studies). Ambient PM was collected near the steel mill during the winter of 1986 (before
18      closure), 1987 (during closure), and again in 1988 (after plant reopening). The glass hi-vol filters
19      were stored, folded PM-side inward, in plastic sleeves at room temperature and humidity (Dye et
20      al., 2001). A description of the in vivo studies follows; the in vitro studies are discussed in
21      Section 7.5.2.1.
22           Ohio and Devlin (2001) investigated the biologic effect of PM from the Utah Valley to
23      determine if the biological responses mirrored the epidemiological findings, with  greater injury
24      occurring after exposure to an equal mass of particles from those years in which the mill was in
25      operation. Aqueous extracts of the filters collected prior to closure of the  steel mill, during the
26      closure and after its reopening, were instilled through a bronchoscope into the lungs of
27      nonsmoking volunteers.  Twenty-four hours later, the same subsegment was lavaged. Exposure
28      to aqueous extracts of PM collected before closure and after reopening of the steel mill provoked
29      a greater inflammatory response than PM extract acquired during the plant shutdown.  These
30      results indicate that the pulmonary effects observed after experimental exposure of humans to the


        April 2002                                 7-16        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Utah Valley PM can be correlated with health outcomes observed in epidemiologic studies of the
 2      same material under normal exposure conditions.
 3           Dye et al (2001) examined the relationship between Utah Valley ambient PM and
 4      respiratory health effects. Sprague-Dawley rats were intratracheally instilled with equivalent
 5      masses of aqueous extracts from filters originally collected during the winter before, during, and
 6      after closure of the steel mill.  Twenty-four hours after instillation, rats exposed to extracts of
 7      particles collected when the plant was  open developed significant pulmonary injury and
 8      neutrophilic inflammation. Additionally, 50% of rats exposed to these extracts had increased
 9      airway responsiveness to acetylcholine, compared to 17 and 25% of rats exposed to saline or the
10      extracts of particles collected when the plant was closed.  By 96 hr, these effects were largely
11      resolved except for increases in lung lavage fluid neutrophils and lymphocytes in rats exposed to
12      PM extracts from prior to the plant  closing. Analogous effects were observed with lung
13      histologic assessment. Extract analysis demonstrated that nearly 70% of the mass in all three
14      extracts appeared to be sodium-based salts derived from the glass filter matrix. Extracts of
15      particles collected when the plant was  open contained more sulfate, cationic salts (i.e., calcium,
16      potassium, magnesium), and certain metals (i.e., copper, zinc, iron, lead, strontium, arsenic,
17      manganese, nickel).  Although total metal content was ~ 1% of the extracts by mass, the greater
18      quantity detected in the extracts of particles collected when the plant was open suggests that
19      metals may be important determinants of the observed pulmonary toxicity.  The authors conclude
20      that the pulmonary effects induced in rats by exposure to aqueous extracts of local ambient PM
21      filters were in good accord with the epidemiologic reports of adverse respiratory health effects in
22      Utah Valley residents.
23           The fact that instillation of ambient PM collected from different geographical areas and
24      from a variety of emission sources consistently caused pulmonary inflammation and injury tends
25      to corroborate epidemiological studies that report increased PM-associated respiratory effects in
26      populations living in many different geographical areas and climates. However, high-dose
27      instillation studies may produce different effects on the lung than inhalation exposures done at
28      more relevant concentrations.  This concern is somewhat diminished by the results of inhalation
29      studies of concentrated PM in healthy nonsmokers.
30           Ohio et al. (2000a) exposed 38 healthy volunteers exercising intermittently at moderate
31      levels of exertion for 2 h to either filtered air or particles concentrated from the air in Chapel

        April 2002                                 7-17       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Hill, NC (23 to 311 //g/m2). Analysis of cells and fluid obtained 18 h after exposure showed a
 2      mild increase in neutrophils in the bronchial and alveolar fractions of bronchoalveolar lavage
 3      (BAL) in subjects exposed to the highest quartile concentration of concentrated PM (mean of
 4      206.7 //g/m3).  Lavage protein did not increase, and there were no other indicators of pulmonary
 5      injury. No respiratory symptoms or decrements in pulmonary function were found after exposure
 6      to CAPs.
 7          The 38 human volunteers reported in Ohio (2000a) were also examined for changes in host
 8      defense and immune parameters in BAL and blood (Harder et al., 2001).  There were no changes
 9      in the number of lymphocytes or macrophages, subcategories of lymphocytes (according to
10      surface marker analysis by flow cytometry), cytokines IL-6 and IL-8, or macrophage
11      phagocytosis.  Similarly, there was no effect of concentrated ambient PM exposure on
12      lymphocyte subsets in blood. Thus, a mild inflammatory response to concentrated ambient PM
13      was not accompanied by an affect on immune defenses as determined by lymphocyte or
14      macrophage effects.
15          Other human inhalation studies with CAPs are limited by the small numbers of subjects
16      studied.  Petrovic et al., 1999 exposed four healthy volunteers (aged  18 to 40) under resting
17      conditions to filtered air and 3 concentrations of concentrated ambient PM (23 to 124 //g/m3) for
18      2 hours using a face mask. The exposure was followed by 30 minutes of exercise.  No cellular
19      signs of inflammation were observed in induced sputum samples collected at 2 or 24 hours after
20      exposure. There was a trend toward an increase in nasal lavage neutrophils although no
21      statistical significance was presented.  The only statistically significant change in pulmonary
22      function was a 6.4% decrease in thoracic gas volume after exposure to 124 //g/m3 PM versus a
23      5.6% increase after air. A similar, small pilot study has been reported (Gong et al., 2000) in
24      which no changes in pulmonary function or symptoms were observed in four subjects aged 19 to
25      41 after a 2 hour exposure to air or mean concentrations of 148 to 246 //g/m3 concentrated
26      ambient PM in Los Angeles, CA. Both of these laboratories are currently expanding on these
27      preliminary findings, but no data are available at this time.
28
29      7.2.1.2  Diesel Particulate Matter
30          Other controlled human exposures of ambient PM that may be relevant to this discussion
31      were the DPM studies previously examined in detail in separate assessment documents  (U.S.

        April 2002                                7-18        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Environmental Protection Agency, 2000; Health Effects Institute, 1995). Briefly, the data from
 2      work shift studies suggest that the principle noncancer human hazard from exposure to diesel
 3      exhaust (DE) includes increased acute sensory and respiratory symptoms (e.g., cough, phlegm,
 4      chest tightness, wheezing) that are more sensitive indicators of possible health risks from
 5      exposure to diesel exhaust than pulmonary function decrements. Immunological changes also
 6      have been demonstrated under short-term exposure scenarios to either diesel exhaust or DPM,
 7      and the evidence indicates that these immunological effects are caused by both the non-
 8      extractable carbon core and the adsorbed organic fraction of the diesel particle.  While noncancer
 9      effects from long-term exposure to DPM of several laboratory animal species include pulmonary
10      histopathology and chronic inflammation, noncancer effects in humans from long-term chronic
11      exposure to DPM are not evident. The mode of action of DPM is not completely understood but
12      the effects on the upper respiratory tract, observed in acute studies, suggest an irritant mechanism
13      while the effects on the lung, observed in chronic studies, indicate an underlying inflammatory
14      response. Currently available data suggest that the carbonaceous core of the diesel particle, or
15      metabolites of metal components of the particle, are possible causative agents for the noncancer
16      lung effects which are mediated, at least in part, by a progressive impairment of alveolar
17      macrophage function. The noncancer lung effects occur in response to DPM in several species
18      and occur in rats at doses lower than those inducing particle overload.
19           Diesel particulate matter, therefore, can be relevant to the urban environment, particularly
20      in urban micro-environments with heavy diesel engine traffic. The findings of controlled studies
21      on DPM are included here and in  Section 7.4.3 (allergic hosts/immunology).
22           Pulmonary function and inflammatory markers (as assayed in induced sputum samples or
23      BAL) have been studied in human subjects exposed to either resuspended or freshly generated
24      and diluted DPM. In a controlled human study, Sandstrom  and colleagues (Rudell et al., 1994)
25      exposed eight healthy subjects in an exposure chamber to diluted exhaust from a diesel engine
26      for 1 h with intermittent exercise.  Dilution of the diesel exhaust was controlled to provide a
27      median NO2 level of approximately 1.6 ppm. Median particle number was 4.3 x 106 /cm3, and
28      median levels of NO and CO were 3.7 and 27 ppm, respectively (particle size and mass
29      concentration were not provided). There were no effects on spirometry or on airway  closing
30      volume. Five of eight subjects experienced unpleasant smell, eye irritation, and nasal irritation
31      during exposure. BAL was performed 18 hours after exposure and was compared with a control

        April 2002                                 7-19         DRAFT-DO NOT QUOTE OR CITE

-------
 1      BAL performed 3 weeks prior to exposure. There was no control air exposure. Small yet
 2      statistically significant reductions were seen in BAL mast cells, AM phagocytic function, and
 3      lymphocyte CD4 to CD8+ cell ratios. A small increase in neutrophils was also observed. These
 4      findings suggest that diesel exhaust may induce mild airway inflammation in the absence of
 5      spirometric changes. Although this early study provided important information on the effect of
 6      diesel exhaust exposure in humans, only one exposure level was used, the number of subjects
 7      was low, and a limited range of endpoints was reported.  A number of follow-up studies have
 8      been done by the same and other investigators.
 9          Rudell et al. (1996) later exposed 12 healthy volunteers to diesel exhaust for 1 h in an
10      exposure chamber.  Light work on a bicycle ergometer was performed during exposure.
11      Random, double-blinded exposures included air, diesel exhaust, or diesel exhaust with particle
12      numbers reduced 46% by a particle trap.  The engine used was a new Volvo model 1990, a six-
13      cylinder direct-injection turbocharged diesel with an intercooler, which was run at a steady speed
14      of 900 rpm during the exposures. Comparison of this study with others is difficult because
15      neither exhaust dilution ratios nor particle concentrations were reported.  Carbon monoxide
16      concentrations of 27-30 ppm and NO of 2.6-2.7 ppm, however, suggested DPM concentrations
17      may have equaled several mg/m3.  The most prominent symptoms during exposure were
18      irritation of the eyes and nose accompanied by an unpleasant smell.  Both airway resistance and
19      specific airway resistance increased significantly during the exposures. Despite the 46%
20      reduction in particle numbers by the trap, effects on symptoms and lung function were not
21      significantly attenuated.
22          A follow-up study on the usefulness of a particle trap  confirmed the lack of effect of the
23      filter on diesel exhaust-induced symptoms (Rudell et al., 1999). In this study,  10 healthy
24      volunteers also underwent BAL 24 hours after exposure. Exposure to diesel exhaust produced
25      inflammatory changes in BAL as evidenced by increases in neutrophils and decreases in
26      macrophage phagocytic function in vitro. A 50% reduction in the particle number concentration
27      by the particle trap did not alter these cellular changes in BAL. Salvi et al. (1999) exposed
28      healthy human subjects to  diluted diesel exhaust (DPM = 300 //g/m3 ) for 1 h with intermittent
29      exercise.  As reported in the studies by Rudell and Sandstrom,  significant increases in neutrophils
30      and B lymphocytes, as well as histamine and fibronectin  in  airway lavage fluid, were not
31      accompanied by decrements in pulmonary function. Bronchial biopsies obtained 6 h after diesel

        April 2002                                7-20        DRAFT-DO NOT QUOTE OR CITE

-------
 1      exhaust exposure showed a significant increase in neutrophils, mast cells, and CD4+ and CD8+
 2      T lymphocytes, along with upregulation of the endothelial adhesion molecules ICAM-1 and
 3      vascular cell adhesion molecule-1 (VCAM-1) and increases in the number of leukoxyte function-
 4      associated antogen-1 (LFA-1+) in the bronchial tissue.  Importantly, extra-pulmonary effects
 5      were observed in these subjects.  Significant increases in neutrophils and platelets were observed
 6      in peripheral blood following exposure to diesel exhaust.
 7           In a follow-up investigation of potential mechanisms underlying the DE-induced airway
 8      leukocyte infiltration, Salvi et al. (2000) exposed healthy human volunteers to diluted DE on two
 9      separate occasions for 1 h each, in an exposure chamber.  Fiber-optic bronchoscopy was
10      performed 6 h after each exposure to obtain endobronchial biopsies and bronchial wash (BW)
11      cells. These workers  observed that diesel exhaust (DE) exposure enhanced gene transcription of
12      interleukin-8 (IL-8) in the bronchial tissue and BW cells and increased growth-regulated
13      oncogene-a protein expression and IL-8 in the bronchial epithelium; there was also a trend
14      toward an increase in interleukin-5 (IL-5) mRNA gene transcripts in the bronchial tissue.
15           Nightingale et al. (2000) have reported inflammatory changes in healthy volunteers
16      exposed to 200 //g/m3 resuspended DPM under resting conditions in a double-blinded study.
17      Small but statistically significant increases in neutrophils  and myeloperoxidase (an index of
18      neutrophil activation) were observed in sputum samples induced 4 hours after exposure to DPM
19      in comparison to air.  Exhaled carbon monoxide was measured as an index of oxidative stress
20      and was found to increase maximally at 1 hour after exposure.  These biochemical and cellular
21      changes occurred in the absence of any decrements in pulmonary function, thus  suggesting that
22      markers of inflammation are more sensitive than pulmonary function measurements.
23           Because of the considerable concern regarding the inhalation of ambient particles by
24      sensitive subpopulations, Sandstrom's laboratory also studied the effect of a 1 hour exposure to
25      300 //g/m3 DPM on 14 atopic asthmatics with stable disease on inhaled corticosteroid treatment
26      (Nordenhall et al., 2001). At 6 hours after exposure, there was a significant increase in IL-6 in
27      induced sputum.  At 24 hours after exposure, there was a  significant increase in  the nonspecific
28      airway responsiveness to inhaled methacholine. Although the exposure level was high relative to
29      ambient PM levels, these findings are important in terms of their relation to the  epidemiology
30      evidence of an increase in asthma morbidity associated with episodic exposure to ambient PM.


        April 2002                                7-21        DRAFT-DO NOT QUOTE OR CITE

-------
 1           The role of antioxidant defenses in protecting against acute diesel exhaust exposure has
 2      been studied. Blomberg et al. (1998) investigated changes in the antioxidant defense network
 3      within the respiratory tract lining fluids of human subjects following diesel exhaust exposure.
 4      Fifteen healthy, nonsmoking, asymptomatic subjects were exposed to filtered air or diesel
 5      exhaust (DPM 300 mg/m3) for 1 h on two separate occasions at least 3 weeks apart. Nasal lavage
 6      fluid and blood samples were collected prior to, immediately after, and 5.5 h post-exposure.
 7      Bronchoscopy was performed 6 h after the end of diesel exhaust exposure. Nasal lavage ascorbic
 8      acid concentration increased tenfold during diesel exhaust exposure, but returned to basal levels
 9      5.5 h post-exposure. Diesel exhaust had no significant effects on nasal lavage uric acid or GSH
10      concentrations and did not affect plasma, bronchial wash, or bronchoalveolar lavage antioxidant
11      concentrations or malondialdehyde or protein carbonyl concentrations. The authors concluded
12      that the acute increase in ascorbic acid in the nasal cavity induced by diesel exhaust may prevent
13      further oxidant stress in the respiratory tract of healthy individuals.
14
15      7.2.1.3 Complex Combustion-Related Particles
16           Because emission sources contribute to the overall ambient air particulate burden (Spengler
17      and Thurston, 1983), many of the studies investigating the response of laboratory animals to
18      particle exposures have used complex combustion-related particles for exposure (see Table 7-3).
19      For example, the residual oil fly ash (ROFA) samples used in toxicological studies have been
20      collected from a variety of sources such as boilers, bag houses used to control emissions from
21      power plants, and from the particles that are emitted downstream of the collection devices (see
22      Table 1).
23           ROFA has a high content of water soluble sulfate and metals, accounting for 82 to 92% of
24      water-soluble mass, while the water-soluble mass fraction in ambient air varies from  low teens to
25      more than 60% (Costa and Dreher, 1997; Prahalad et al., 1999). More than 90% of the metals in
26      ROFA are transition metals; whereas these metals are only a small subfraction of the total
27      ambient PM mass.  Thus, the dose of bioavailable metal that is delivered to the lung when ROFA
28      is instilled into a laboratory animal can be orders of magnitude greater than an ambient PM dose,
29      even under a worst-case scenario.
30           Intratracheal instillation of various doses of ROFA suspension has been shown  to produce
31      severe inflammation, an indicator of pulmonary injury that includes recruitment of neutrophils,

        April 2002                                 7-22        DRAFT-DO NOT QUOTE OR CITE

-------
 1      eosinophils, and monocytes into the airway.  The biological effects of ROFA in rats have been
 2      shown to depend on aqueous teachable chemical constituents of the particles (Dreher et al., 1997;
 3      Kodavanti et al., 1997b). A leachate prepared from ROFA, containing predominantly Fe, Ni, V,
 4      Ca, Mg, and sulfate, produced similar lung injury to that induced by the complete ROFA
 5      suspension (Dreher et al., 1997).  Depletion of Fe, Ni, and V from the ROFA leachate eliminated
 6      its pulmonary toxicity. Correspondingly, minimal lung injury was observed in animals exposed
 7      to saline-washed ROFA particles. A surrogate transition metal  sulfate solution containing Fe, V,
 8      and Ni largely reproduced the lung injury induced by ROFA.  Interestingly, ferric sulfate  and
 9      vanadium sulfate antagonized the pulmonary toxicity of nickel sulfate. Interactions between
10      different metals and the acidity of PM were found to influence the severity and kinetics of lung
11      injury induced by ROFA and its soluble transition metals.
12           To further investigate the response to ROFA with differing metal and sulfate composition,
13      male Sprague-Dawley rats (60 days old) were exposed to ROFA or metal sulfates (iron,
14      vanadium, and nickel, individually or in combination) (Kodavanti et al.,  1997b).  Transition
15      metal sulfate mixtures caused less injury than ROFA or Ni alone, suggesting metal interactions.
16      In addition,  this study showed that V-induced effects were less severe than that of Ni and were
17      transient. Ferric sulfate was least pathogenic. Cytokine gene expression was induced prior to the
18      pathology changes in the lung, and the kinetics of gene expression suggested persistent injury by
19      nickel sulfate.  Another study by the same investigators was performed using 10 different ROFA
20      samples collected at various sites within a power plant burning residual oil (Kodavanti et al.,
21      1998a).  Animals received intratracheal instillations of either saline (control), or a saline
22      suspension of whole ROFA (<3.0 //m MMAD) at three concentrations (0.833, 3.33, or
23      8.33 mg/kg). This study showed that ROFA-induced PMN influx was associated with its water-
24      teachable V content; however, protein leakage was associated with water-leachable Ni content.
25      ROFA-induced in vitro activation of alveolar macrophages (AMs) was highest with ROFA
26      containing teachable V but not with Ni plus V, suggesting that the potency and the mechanism of
27      pulmonary injury may differ between emissions containing bioavailable V and Ni.
28           Other  studies have shown that soluble metal components play an important role  in the
29      toxicity of emission source particles. Gavett et al. (1997) investigated the effects of two ROFA
30      samples of equivalent diameters, but having different metal and sulfate content, on pulmonary
31      responses in Sprague-Dawley rats. ROFA sample 1 (Rl) (the same emission particles used by

        April 2002                                7-23        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Dreher et al. [1997]) had approximately twice as much saline-leachable sulfate, nickel, and
 2      vanadium, and 40 times as much iron as ROFA sample 2 (R2); whereas R2 had a 31-fold higher
 3      zinc content. Rats were instilled with suspensions of 2.5 mg R2 in 0.3 mL saline, the supernatant
 4      of R2 (R2s), the supernatant of 2.5 mg Rl (Rls), or saline only. By 4 days after instillation, 4 of
 5      24 rats treated with R2s or R2 had died. None of those treated with Rls or saline died.
 6      Pathological indices, such as alveolitis, early fibrotic changes, and perivascular edema, were
 7      greater in both R2 groups. In surviving rats, baseline pulmonary function parameters and airway
 8      hyperreactivity to acetylcholine were significantly worse in the R2 and R2s groups than in the
 9      Rls groups. Other than BAL neutrophils, which were significantly higher in the R2 and R2s
10      groups, no other inflammatory cells (macrophages, eosinophils, or lymphocytes) or biochemical
11      parameters of lung injury were significantly different between the R2 and R2s groups and the
12      Rls group. Although soluble forms of zinc had been found in guinea pigs to produce a greater
13      pulmonary response than other sulfated metals (Amdur et al., 1978), and, although the level of
14      zinc was 30-fold  greater in R2 than Rl, the precise mechanisms by which zinc may induce such
15      responses are unknown. Nevertheless, these results show that the composition of soluble metals
16      and sulfate is critical in the development of airway hyperractivity and lung injury produced by
17      ROFA, albeit at high concentrations.
18           Reactive oxygen species may play an important role in  the in vivo toxicity of ROFA.  Dye
19      et al. (1997) pretreated rats with an intraperitoneal injection of saline or dimethylthiourea
20      (DMTU) (500 mg/kg), followed 30 min later by intratracheal instillation of either acidic saline
21      (pH = 3.3) or an acidified suspension of ROFA (500 //g/rat).  The systemic administration of
22      DMTU impeded  development of the cellular inflammatory response to ROFA but did not
23      ameliorate biochemical alterations in BAL fluid. In a subsequent study, these investigators
24      determined that oxidant generation, possibly induced by soluble vanadium compounds in ROFA,
25      is responsible  for the subsequent rat tracheal epithelial cells gene expression, inflammatory
26      cytokine production (MTP-2 and IL-6), and cytotoxicity (Dye et al., 1999).
27           In addition to transition metals, other components in fly ash also may cause lung injury.
28      The effects of arsenic compounds in coal fly ash or copper smelter dust on the lung integrity and
29      on the ex vivo release of TNFa by alveolar phagocytes were  investigated by Broeckaert et al.
30      (1997). Female Naval Medical Research Institute (NMRI) mice were instilled with different
31      particles normalized for the arsenic content (20 //g/kg body weight [i.e., 600 ng arsenic/mouse])

        April 2002                               7-24        DRAFT-DO NOT  QUOTE OR CITE

-------
 1      and the particle load (100 mg/kg body weight [i.e., 3 mg/mouse]). Mice received tungsten
 2      carbide (WC) alone, coal fly ash (CFA) alone, copper smelter dust (CMP) mixed with WC, and
 3      Ca3(AsO4)2 mixed with WC (see Table 7-2 for concentration details). Copper smelter dust
 4      caused a severe but transient inflammatory reaction; whereas a persisting alveolitis (30 days
 5      postexposure) was observed after treatment with coal fly ash. In addition, TNFa production in
 6      response to lipopolysaccharide (LPS) by alveolar phagocytes were significantly inhibited at Day
 7      1 but was still observed at 30 days after administration of CMP and CFA. Although arsenic was
 8      cleared from the lung tissue 6 days after Ca3(AsO4)2 administration, a significant fraction
 9      persisted (10 to 15% of the arsenic administered) in the lung of CMP- and CFA-treated mice at
10      Day 30.  It is possible that suppression of TNF-a production is dependent upon the slow
11      elimination of the particles and their metal content from the lung.
12           In summary, intratracheally instilled ROFA produced acute lung injury and inflammation.
13      The water soluble metals in ROFA appear to play a key role in the acute effects of instilled
14      ROFA. Although studies done with ROFA clearly  show that combustion generated particles
15      with a high metal content can cause substantial lung injury, there are still insufficient data to
16      extrapolate the high dose effects to the low levels of particle associated transition metals in
17      ambient PM.
18
19      7.2.2 Acid Aerosols
20           There have been extensive studies of the effects of controlled exposures to aqueous acid
21      aerosols on various aspects of lung function in humans and laboratory animals. Many of these
22      studies were  reviewed in the previous criteria document (U.S. Environmental Protection Agency
23      1996a) and in the Acid Aerosol Issue Paper (U.S. Environmental Protection Agency, 1989);
24      some of the more recent studies are summarized in  this document (Table 7-5). Methodology and
25      measurement methods for controlled human exposure studies have been reviewed elsewhere
26      (Folinsbeeetal., 1997).
27           The studies summarized in the previous document illustrate that aqueous acidic aerosols
28      have minimal effects on symptoms and mechanical lung function in young healthy adult
29      volunteers at concentrations as high as 1000 //g/m3. However, at concentrations as low as
30      100 //g/m3, acid aerosols can alter mucociliary clearance. Brief exposures (< 1 h) to low
31      concentrations («100 //g/m3) may accelerate clearance while longer (multihour) exposures to
        April 2002                                7-25         DRAFT-DO NOT QUOTE OR CITE

-------
>
to
o
o
to
              TABLE 7-5. RESPIRATORY EFFECTS OF ACID AEROSOLS IN HUMANS AND LABORATORY ANIMALS
to
H
6
o
o
H
O
Species, Gender, Strain
Age, etc.
Dogs, beagle, healthy;
n= 16

Humans, asthmatic;
13 M, 11 F
Rats, female, Fischer
344; Guinea Pigs,
female, Hartley
Humans, healthy
nonsmokers; 10 M,
2 F; 21-37 years old
Particle
Neutral sulfite
aerosol
Acidic sulfate
aerosol
H2SO4 aerosol
NH+4/SCr4
aerosol
H2SO4 aerosol
H2SO4 aerosol
Exposure
Technique
Inhalation
Inhalation
Inhalation by
face mask
Inhalation
Inhalation
Concentration Particle Size
1.5 mg/m3 1.0 Mm MMAD
og = 2.2
5. 7 mg/m3 1.1 Mm MMAD
og = 2.0
500 Mg/m3 9 Mm MMAD
7 Mm MMAD
94 mg/m3 0.80±1.89og
43 mg/m3 0.93 ±2. Hog
1,000 Mg/m3 0.8-0.9 Mm
MMAD
Exposure
Duration
16.5 h/day
for 13 mo
6 h/day for
13 mo
Ih
4h
3h
Effects of Particles
Long-term exposure to particle-associated sulfur and
hydrogen ions at concentrations close to ambient
levels caused only subtle respiratory responses and no
change in lung pathology.

Exposure to simulated natural acid fog did not induce
bronchoconstriction nor change bronchial
responsiveness in asthmatics.
Acid aerosol increased surfactant film compressibility
in guinea pigs.
No inflammatory responses; LDH activity in BAL was
elevated. Effect on bacterial killing by macrophages
was inconclusive; latex particle phagocytosis was
reduced 28%.
Reference
Heyderetal. (1999)

Leducetal. (1995)
Lee etal. (1999)
Zelikoffetal. (1997)
        H2SO4 = Sulfuric acid
        BAL = Bronchoalveolar lavage
        LDH = Lactate dehydrogenase
        MMAD = Mass median aerodynamic diameter
        MMD = Mass median diameter
        og = Geometric standard deviation
o
HH
H
W

-------
 1      higher concentrations (>100 //g/m3) can depress clearance. Asthmatic subjects appear to be more
 2      100 //g/m3, acid aerosols can alter mucociliary clearance.  Brief exposures (< 1 h) to low
 3      concentrations («100 //g/m3) may accelerate clearance while longer (multihour) exposures to
 4      higher concentrations (>100 //g/m3) can depress clearance. Asthmatic subjects appear to be more
 5      sensitive to the effects of acidic aerosols on mechanical lung function. Responses have been
 6      reported in adolescent asthmatics at concentrations as low as 68 //g/m3, and modest
 7      bronchoconstriction has been seen in adult asthmatics exposed to concentrations >400 //g/m3, but
 8      the available data are not consistent.
 9           A previously described, acid aerosol exposure in humans (1000 //g/m3 H2SO4) did not
10      result in airway inflammation (Frampton et al., 1992), and there was no evidence of altered
11      macrophage host defenses.  A more recent study by Zelikoff et al. (1997) compared the responses
12      of rabbits and humans exposed to similar concentrations of H2SO4 aerosol. For both rabbits and
13      humans, there was no evidence of PMN infiltration into the lung and no change in BAL fluid
14      protein level, although there was  an increase in LDH in rabbits but not in humans.  Macrophages
15      showed less antimicrobial activity in rabbits; insufficient data were available for humans.
16      Macrophage phagocytic activity was slightly reduced in rabbits but not in humans.  Superoxide
17      production by macrophages was somewhat depressed in both species.  No respiratory effects of
18      long-term exposure to acid aerosol were found in dogs (Heyder et al., 1999).  Thus, recent studies
19      have  not provided any additional  evidence to unequivocally demonstrate that relevant
20      concentrations of aqueous acid aerosols contribute  to the acute cardiopulmonary effects  of
21      ambient PM.
22
23      7.2.3 Metal Particles,  Fumes, and Smoke
24           Data from occupational and laboratory animal studies reviewed in the previous criteria
25      document (U. S. Environmental Protection Agency, 1996a) indicated that acute exposures to very
26      high levels (hundreds of//g/m3 or more) or chronic exposures to lower levels (up to 15 //g/m3) of
27      metallic particles could have an effect on the respiratory tract.  Therefore, it was concluded on
28      the basis of data available at that time that the metals at typical concentrations present in the
29      ambient atmosphere  (1 to 14 //g/m3) were  not likely to have a significant acute effect in healthy
30      individuals. The metals include arsenic, cadmium, copper, nickel, vanadium, iron, and zinc.
31      Other metals found at concentrations less than 0.5 //g/m3 were not reviewed in the previous
        April 2002                                7-27       DRAFT-DO NOT QUOTE  OR CITE

-------
 1      criteria document.  However, more recently published data from high-dose laboratory animal
 2      studies added to the existing PM data base indicate that particle-associated metals are among the
 3      potential causal components of PM.
 4           Since completion of the previous criteria document, only limited controlled human
 5      exposure studies have been performed with particles other than acid aerosols (see Table 7-6).
 6      Controlled inhalation exposure studies to high concentrations of two different fume particles,
 7      MgO and ZnO, demonstrate the differences in response based on particle metal composition
 8      (Kuschner et al., 1997).  Up to 6400 mg/m3/min cumulative dose of MgO had no effect on lung
 9      function (spirometry, DLCO), symptoms of metal  fume fever, or changes in inflammatory
10      mediators or cells recovered by BAL. However, lower concentrations of ZnO fume (165 to
11      1110 mg/m3/min) induced a neutrophilic inflammatory response in the airways 20 h
12      postexposure.  Lavage fluid PMNs, TNF-a, and IL-8 were increased by ZnO exposure.  Although
13      the concentrations used in these exposure studies exceed ambient levels by more than 1000-fold,
14      the absence  of a response to an almost 10-fold higher concentration of MgO compared with ZnO
15      indicates that metal composition, in addition to particle size (ultrafme/fine), is an important
16      determinant of the observed health responses to  inhaled PM.
17           Several metals, including zinc,  chromium, cobalt, copper, and vanadium, have been shown
18      to stimulate cytokine release in cultured human pulmonary cells.  Boiler makers, exposed
19      occupationally to approximately 400  to 500 //g/m3 of fuel oil ash, which contains high levels of
20      soluble metals, showed acute nasal inflammatory responses characterized by increased
21      myeloperoxidase (MPO) and IL-8 levels; these changes were associated with increased vanadium
22      levels in the upper airway (Woodin et al., 1998). Irsigler et al. (1999) reported that V2O5 can
23      induce asthma and bronchial hyperreactivity in exposed workers.
24           Autopsy data suggest that chronic exposure to urban air pollution leads to an increased
25      retention of metals in human tissues.  A comparison of autopsy cases in Mexico City from the
26      1950s with the  1980s indicated substantially higher (5- to 20-fold) levels of Cd, Co, Cu, Ni, and
27      Pb in lung tissue from the 1980s (Fortoul et al.,  1996).  Similar studies have examined metal
28      content in human blood and lung tissue (Tsuchiyama et al.,  1997; Osman et al., 1998) with
29      similar results.
30


        April 2002                                7-28        DRAFT-DO NOT QUOTE OR CITE

-------
TABLE 7-6. RESPIRATORY EFFECTS OF METAL PARTICLES, FUMES, AND SMOKE IN HUMANS AND
                              LABORATORY ANIMALS
to
o
o
to













1
to
**Q



O
l>
H
M
\^
o

0
H
/O
r*^x
o
H
W
O
^
o
HH
H
W


Species, Gender,
Strain, Age, etc.
Mice, Swiss



Rats, SD; 60
days old


Humans, healthy
nonsmokers;
12 M, 4 F;
18-35 years old
Humans,
vanadium plant
workers; 40 M;
19-60 years old
Humans, healthy
nonsmokers;
4 M, 2 F;
21-43 years old
Humans, healthy
nonsmokers;
27 M, 7 F;
20-36 years old
Rats, Fischer
344. (250 g)


Humans, healthy
nonsmokers;
8 M, 8 F;
18-34 years old

Mice, NMRI;
Mouse
peritoneal
macrophage
Particle
EHC-93
soluble
metal
salts
VS04
NiS04


Colloidal
iron oxide


VA



MgO
ZnO


FeA



FeA



FeA




Mn02



Exposure
Technique
Intratracheal
instillation


Inhalation



Bronchial
instillation


Inhalation



Inhalation



Intrapulmonary
instillation


Intratracheal
instillation


Inhalation




Intratracheal
instillation;
in vitro

Exposure
Concentration Particle Size Duration
ImginO.lml 0.8±0.4,um 3 days



0.3 - 2.4 mg/m3 N/A 6h/day x 4 days



5 mg in 10 mL 2.6 //m 1, 2, and 4 days
after instillation


0.05-1.53 N/A Variable
mg/m3


5.8-230 mg/m3 99% < 1.8 fj,m 15-45 min
29% < 0.1 fj.m


3 x 108 2.6 urn N/A
microspheres in
10 mL saline.

7.7 x 107 2.6 fj.m N/A
microspheres in
5 mL saline

12.7 mg/m3 1.5 ^m 30 min
ag = 2.1



0.037, 0. 12, 0.75, surface area of Sacrificed at
2.5 mg/animal 0.16, 0.5; 17, 5 days
62 m2/g

Effect of Particles
Solution containing all metal salts (Al, Cu, Fe, Pb,
Mg, Ni, Zn) or ZnCl alone increased BAL
inflammatory cells and protein.

V did not induce any significant changes in BAL or
HR; Ni caused delayed bradycardia, hypothermia, and
arrhythmogenesis at> 1.2 mg/m3; possible synergistic
effects were found.
L-ferritin increased after iron oxide particle exposure;
transferrin was decreased. Both lactoferrin and
transferrin receptors were increased.

12/40 workers had bronchial hyperreactivity that
persisted in some for up to 23 mo.


No significant differences in BAL inflammatory cell
concentrations, BAL interleukins (IL-1, IL-6, IL-8),
tumor necrosis factor, pulmonary function, or
peripheral blood neutrophils.
Transient inflammation induced initially (neutrophils,
protein, LDH, IL-8) was resolved by 4 days
postinstillation.

Transient inflammation at 1 day postinstillation.



No significant difference in 98mTc-DTPA clearance
half-times, DLCO, or spirometry



LDH, protein and cellular recruitment increased with
increasing surface area; freshly ground particles had
enhanced cytotoxicity.

Reference
Adamson et al.
(2000)


Campen et al.
(2001)


Ohio etal.
(1998a)


Irsigler et al.
(1999)


Kuschner et al.
(1997)


Lay etal. (1998)



Lay etal. (1998)



Lay etal. (2001)




Li son et al.
(1997)



-------
>
to
o
o
to
          TABLE 7-6 (cont'd). RESPIRATORY EFFECTS OF METAL PARTICLES, FUMES, AND SMOKE IN HUMANS AND
                                                               LABORATORY ANIMALS
Species, Gender,
Strain, Age, etc. Particle
Rats, WISTAR CdO Fume
Furth;
7-week-old,
Mice, C57BL6
and DBA3NCR
Rats, M, F344, TiO2
175-225 g
Rats, M. F344, TiO2
175-225 g

Rats NaVO3
VOSO4
V205
Humans, ROFA
boilermakers
(18 M), 26-61
years old, and
utility worker
controls (11 M),
30-55 years old
Exposure
Technique
Nose-only
Inhalation


Intratracheal
inhalation and
Intratracheal
instillation
Intratracheal
inhalation and
Intratracheal
instillation
Intratracheal
instillation
Inhalation of
fuel-oil ash

Concentration Particle Size
1.04 mg/m3 CMD = 0.008
Rats dose = ,um ag = 1.1
18.72,ug
Mouse dose =
4.59 Mg
Inhalation at Fine: 250 nm
125 mg/m3 for Ultrafine:
2 h; Instillation at 21 nm
500 //g for fine,
750 //g for
ultrafine
Inhalation at Fine: 250 nm
125 mg/m3 for Ultrafine:
2 h; Instillation at 21 nm
500 //g for fine,
750 //g for
ultrafine
21 or 210 Mg N/A
V/kg (NaVO3,
VOSO4 soluble)
42 or 420 //g
V/kg (V20S) less
soluble
0.4-0.47 mg/m3 10 ,um
0.1-0. 13 mg/m3

Exposure
Duration
1 x 3h


Inhalation
exposure, 2 h;
sacrificed at 0,
1, 3, and 7 days
postexposure for
both techniques
Inhalation
exposure, 2 h;
sacrificed at 0,
1, 3, and 7 days
postexposure for
both techniques
1 h or 10 days
following
instillation
6 weeks

Effect of Particles
Mice created more metallothionein than rats, which
may be protective of tumor formation.


Inflammation produced by intratracheal inhalation
(both severity and persistence) was less than that
produced by instillation; ultrafine particles produced
greater inflammatory response than fine particles for
both dosing methods.
MIP-2 increased in lavage cells but not in supernatant
in those groups with increased PMN (more in
instillation than in inhalation; more in ultrafine than in
fine); TNF-a levels had no correlation with either
particle size or dosing methods.
PMN influx was greatest following VOSO4, lowest for
V2O5; VOSO4 induced inflammation persisted longest;
MIP-2 and KC (CXC chemokines) were rapidly
induced as early as 1 h postinstillation and persisted
for 48 h; Soluble V induced greater chemokine mRNA
expression than insoluble V; AMs have the highest
expression level.
Exposure to fuel-oil ash resulted in acute upper airway
inflammation, possibly mediated by increased IL-8
and PMNs.

Reference
McKenna et al.
(1998)


Osier and
Oberdorster
(1997)
Osier et al.
(1997)

Pierce et al.
(1996)
Woodin et al.
(1998)

       CdO = Cadmium oxide
       Fe2O3 = Iron oxide
       MgO = Magnesium oxide
       MnO2 = Manganese oxide
       NaVO3 =
       TiO2 = Titanium oxide
       VOSO4 = Vanadium sulfate
       V2O5 = Vanadium oxide
       ZnO = Zinc oxide
BAL = Bronchoalveolar lavage
CMD = Count median diameter
IL = Interleukin
LDH = Lactate dehydrogenase
MIP-2 = Macrophage inflammatory protein-2
mRNA = Messenger RNA (ribonucleic acid)
N/A = Data not available

-------
 1           Iron is the most abundant of the elements that are capable of catalyzing oxidant generation
 2      and also is present in ambient urban particles. Lay et al.  (1998) and Ohio et al. (1998a) tested the
 3      hypothesis that the human respiratory tract will attempt to diminish the added, iron-generated
 4      oxidative stress.  They examined the cellular and biochemical response of human subjects,
 5      instilled via the intrapulmonary route, with a combination of iron oxyhydroxides that introduced
 6      an oxidative stress to the lungs.  Saline alone and iron-containing particles suspended in saline
 7      were instilled into separate lung segments of human subjects. Subjects underwent
 8      bronchoalveolar lavage at 1 to 91 days after instillation of 2.6-//m diameter iron oxide
 9      agglomerates. Lay and colleagues found iron-oxide-induced inflammatory responses in both the
10      alveolar fraction and the bronchial fraction of the lavage fluid at 1 day postinstillation.  Lung
11      lavage 24 h after instillation revealed decreased transferrin concentrations and increased ferritin
12      and lactoferrin concentrations, consistent with a host-generated response to decrease the
13      availability of catalytically reactive iron (Ohio et al., 1998a). Normal iron homeostasis returned
14      within 4 days of the  iron particle instillation. The same iron oxide preparation, which contained
15      a small amount of soluble iron, produced similar pulmonary inflammation in rats. In contrast,
16      instillation of rats with two iron oxide preparations that contained no soluble iron failed to
17      produce injury or inflammation, thus suggesting that soluble iron was responsible for the
18      observed intrapulmonary changes. Although the total dose of iron oxide delivered acutely to the
19      lung segments (approximately 5 mg or 2.1 x 108 particles) is considerably higher than would be
20      deposited in the lung at the concentrations of iron present in ambient urban air (generally less
21      than 1 //g/m3), only a small amount of the iron instilled in human subjects was "active."
22      Therefore, it is still not clear how the amount of active iron in the PM extract compares with the
23      iron found in ambient air particles.
24           In a subsequent inhalation study, Lay et al. (2001) studied the effect of iron oxide particles
25      on lung epithelial cell permeability. Healthy, nonsmoking human subjects inhaled 12.7 mg/m3
26      low- and high-solubility iron oxide particles (MMAD = 1.5 //m and og = 2.1) for 30 minutes.
27      Neither pulmonary function nor alveolar epithelial permeability, as assessed by pulmonary
28      clearance of technetium-labeled DPT A, was changed at 0.5 or 24 hours after exposure to either
29      type of iron oxide particle. Because the exposure concentration was so high, the data suggest that
30      metals may play little role in the adverse effects of ambient, urban PM.  Ohio et al. (2001) have
31      reported a case study, however, in which acute exposure to oil fly ash from a domestic oil-

        April 2002                                 7-31         DRAFT-DO NOT QUOTE OR CITE

-------
 1      burning stove produced diffuse alveolar damage, difficulty in breathing, and symptoms of angina.
 2      While steroid treatment led to rapid improvement in symptoms and objective measurements, this
 3      report suggests that the high metal content of oil fly ash can alter the epithelial cell barrier in the
 4      alveolar region.
 5
 6      7.2.4 Ambient Bioaerosols
 7           Ambient bioaerosols include fungal spores, pollen, bacteria, viruses, endotoxins, and plant
 8      and animal debris.  Such biological aerosols can produce various health effects including
 9      irritation, infection, hypersensitivity, and toxic response. Bioaerosols present in the ambient
10      environment have the potential to cause disease in humans under certain conditions. However, it
11      was concluded in the previous criteria document (U.S. Environmental Protection Agency, 1996a)
12      that bioaerosols, at the concentrations present in the ambient environment, would not contribute
13      to the observed effects of paniculate matter on human mortality and morbidity reported in PM
14      epidemiological studies. Moreover, bioaerosols generally represent a rather small fraction of the
15      measured urban ambient PM mass and are typically present even at lower concentrations during
16      the winter months when notable ambient PM effects have been demonstrated.  Bioaerosols tend
17      to be in the coarse fraction of PM, but some bioaerosols including nonagglomerated bacteria and
18      fragmented pollens, are found in the fine fraction.
19           More recent inhalation studies on ambient bioaerosols are summarized in Table 7-7.
20      In vitro studies on particle-associated endotoxin are discussed in Section 7.5.2.2.  Endotoxin,
21      a cell wall component of gram negative bacteria, is ubiquitous in the environment. Although
22      there is strong evidence that inhaled endotoxin plays a major role in the toxic effects of
23      bioaerosols encountered in the work place (Vogelzang et al., 1998;  Castellan et al., 1984, 1987),
24      it is not clear whether ambient concentrations of endotoxin are sufficient to produce toxic
25      pulmonary or systemic effects in healthy or sick individuals.
26           Michel et al. (1997) examined the dose-response relationship  to inhaled lipopolysaccharide
27      (LPS: the purified derivative of endotoxin) in normal healthy volunteers exposed to 0, 0.5, 5, and
28      50 //g of LPS.  Inhalation of 5 or 50 //g of LPS resulted in increased PMNs in blood and sputum
29      samples.  At the higher concentration, a slight (3%) but not significant decrease in FEVj was
30      observed. Cormier et al. (1998) reported an approximate 10% decline in FEVj and an increase in
31      methacholine airway responsiveness after a 5-h exposure inside a swine containment building.
        April 2002                                7-32        DRAFT-DO NOT QUOTE OR CITE

-------
>
to
o
o
to
                        TABLE 7-7. RESPIRATORY EFFECTS OF AMBIENT BIOAEROSOLS
Species, Gender,
Strain, Age, etc.
Rats, Fischer 344,
8 weeks to
20 months old,
N = 3/group


Humans, healthy;
5 M, 4 F, 24 to 50
years of age




Humans, healthy;
32 M, 32 F, 16 to
50 years of age

Humans, pig
farmers,
82 symptomatic
and
89 asymptomatic
n= 171
Humans, potato
plant workers, low
exposures (37 M),
high exposures
(20 M)

Particle
LPS
(endotoxin)

UF carbon

ozone
LPS
(endotoxin)





Indoor pool
water spray


Dust


Endotoxin


Endotoxin





Exposure
Technique Concentration Particle Size
Inhalation 70 EU 0.72 ij,m
og= 1.6

100 //g/m3 25 nm
og= 1.6
1 ppm
Inhalation 0.5 //g 1-4 ,um
5.0,ug MMAD
50 Mg




Inhalation N/A 0.1-7.5^m



Inhalation 2.63 mg/m3 N/A
og= 1.3

105 ng/m3
og=1.5

Inhalation 21.2EU/m3low N/A
og = 1.6

55.7EU/m3
high
og = 2.1
Exposure
Duration Effect of Particles
12 min Significant interaction of LPS and O3 on inflammatory
responses in young rats. O3 and UF-C interacted with
"priming" by LPS to produce greater PMN response.
6 h LPS has a priming effect on lung inflammatory response
to O3 and UF-C.

30 min Significant decrease in PMN luminol-enhanced
chemiluminescence with 0.5 /^g LPS; increase in blood
CRP and PMNs, and increase in sputum PMNs,
monocytes, and MPO with 5.0 ,ug LPS; increase in
temperature, blood PMNs, blood and urine CRP, sputum
PMNs, monocytes, lymphocytes, TNFa, and ECP with
50 Mg LPS.
N/A Recurring outbreaks of pool-associated granulomatous
pneumonitis (n = 33); case patients had higher cumulative
work hours. Analysis indicated increased levels of
endotoxin in pool air and water.
5 h/day average Large decline in FEV! (73 ml/year) and FVC (55 ml/year)
lifetime exposure associated with long-term average exposure to endotoxin.




8 h Decreased FEV1: FVC, and MMEF over the work shift
that was concentration related; endotoxin effects on lung
function can be expected above 53 EU/m3 (=4.5 ng/m3)
over 8 h.


Reference
Elder et al.
(2000a,b)




Michel et al.
(1997)





Rose et al.
(1998)


Vogelzang et al.
(1998)




Zock et al.
(1998)





-------
 1      This exposure induced significant neutrophilic inflammation in both the nose and the lung.
 2      Although these exposures are massive compared to endotoxin levels in ambient PM in U.S.
 3      cities, these studies serve to illustrate the effects of endotoxin and associated bioaerosol material
 4      in healthy nonsensitized individuals.
 5          Some health effects have been observed after occupational exposure to complex aerosols
 6      containing endotoxin at concentrations relevant to ambient levels. Zock et al. (1998) reported a
 7      decline in FEVj («3%) across a shift in a potato processing plant with up to 56 endotoxin units
 8      (EU)/m3 in the air.  Rose et al. (1998) reported a high incidence (65%) of BAL lymphocytes in
 9      lifeguards working at a swimming pool where endotoxin levels in the air were on the order of
10      28 EU/m3. Although these latter two studies may point towards pulmonary changes at low
11      concentrations of airborne endotoxin, it is not possible to rule out the contribution of other agents
12      in these complex organic aerosols. The contribution of endotoxin to the toxicity of ambient PM
13      has been studied in vitro, and these studies provide preliminary evidence that endotoxin
14      contamination of ambient PM may play a role in the observed in vitro effects (discussed in
15      Section 7.5).
16
17
18      7.3  CARDIOVASCULAR AND SYSTEMIC EFFECTS OF PARTICULATE
19          MATTER IN HUMANS AND LABORATORY ANIMALS: IN VIVO
20          EXPOSURES
21          A growing number of epidemiology studies have demonstrated that increases in cardiac-
22      related deaths are associated with exposure to PM (U.S. Environmental Protection Agency,
23      1996a) and that PM-related cardiac deaths appear to be as great or greater than those attributed to
24      respiratory causes (see Chapter 8). The toxicological consequences of inhaled particles on the
25      cardiovascular system had not been extensively investigated prior to 1996.  Since then (see
26      Table 7-8), Costa and colleagues (e.g., Costa and Dreher, 1997) have demonstrated that
27      intratracheal instillation of high levels of ambient particles can increase or accelerate death in an
28      animal model of cardiorespiratory disease related to monocrotaline administration in rats. These
29      deaths did not occur with all types of ambient particles tested.  Some dusts, such as volcanic ash
30      from Mount Saint Helens, were relatively inert; whereas other ambient dusts, including those
31      from urban sites, were toxic.  These early observations suggested that particle composition plays

        April 2002                               7-34        DRAFT-DO NOT QUOTE OR CITE

-------
TABLE 7-8. CARDIOVASCULAR AND SYSTEMIC EFFECTS OF AMBIENT AND COMBUSTION-RELATED
                               PARTICULATE MATTER
to
o
o
to













^,
OJ
^r<



O
c
"Tl
H
6
o
h-^
-Z
H
O
d
o
3
o
^d
O
i — 1
— 1
w




Species, Gender,
Strain Age, or Body
Weight
Rats, male, F-344;
200-250 g


Rats, male, S-D,
60 days old,
MCT-treated and
healthy, n = 64
Dogs, female
mongrel,
14 to 17 kg


Rats, male, S-D,
60 days old,
MCT-treated,
and healthy

Rats, male, S-D;
60 days old
Humans, healthy
nonsmokers,
18 to 40 years old
Dogs, mongrel,
some with balloon
occluded LAD
coronary artery,
n= 14
Rats


Rats, male, F-344,
MCT-treated

Hamsters, 6-8 mo
old; Bio TO-2
Rats, S-D,
MCT-treated, 250 g


Particle
OTT



ROFA



CAPs




Emission source
PM
Ambient airshed
PM
ROFA
ROFA

CAPs


CAPs




CAPs


CAPs




FOFA



Exposure
Technique
Nose-only
Inhalation


Instillation



Inhalation via
tracheostomy



Instillation




Instillation

Inhalation


Inhalation via
tracheostomy



Nose-only
inhalation

Inhalation




Inhalation



Mass
Concentration
40 mg/m3



0.0,0.25, 1.0,
and 2.5 mg/rat


3-360 Mg/m3




Total mass:
2.5 mg/rat

Total transition
metal: 46 Mg/rat
0.3, 1.7, or
8.3 mg/kg
23.1 to
311.1 Mg/m3

69-828 Mg/m3




110-350 Mg/m3


132-919 Mg/m3




580 ± 110 Mg/m3



Particle
Size
4 to 5 Mm MMAD



1.95 Mm



0.2 to 0.3 Mm




Emission PM:
1.78-4. 17 Mm

Ambient PM:
3. 27-4.09 Mm
1.95 Mm
og = 2.19
0.65 Mm
og = 2.35

0.23 to 0.34 Mm
og = 0.2to2.9



N/A


0.2-1. 2 Mm
og = 0.2-3.9



2.06 Mm MMAD
og= 1.57


Exposure
Duration
4h



Analysis at
96 h


6 h/day for
3 days



Analysis at
24 and 96 h
following
instillation

Analysis at
24 h
2 h, analysis
at 18 h

6 h/day for
3 days



3h


3 h, evaluated
at 3 and 24 h



6 h/day for
3 days


Cardiovascular Effects
Increased plasma levels of endothelin-1.
No acute lung injury; however, lung NO production
decreased and macrophage inflammatory protein-2 from
lung lavage cells increased after exposure.
Dose-related hypothermia and bradycardia in healthy rats,
potentiated by compromised models.


Peripheral blood parameters were related to specific
particle constituents. Factor analysis from paired and
crossover experiments showed that hematologic changes
were not associated with increases in total CAP mass
concentration.
ROFA alone induced some mild arrhythumas;
MCT-ROFA showed enhanced neutrophilic inflammation;
MCT-ROFA animals showed more numerous and
severe arrhythmias including S-T segment inversions
and A-V block.
Increased plasma fibrinogen at 8.3 mg/kg only.

Increased blood fibrinogen.


Decreased time to ST segment elevation and increased
magnitude in compromised dogs. Decreased heart and
respiratory rate and increased lavage fluid neutrophils in
normal dogs.

Small but consistent increase in HR; no pulmonary injury
was found; increased peripheral blood neutrophils and
decreased lymphocytes.
No increase in cardiac arrhythmias; PM associated
increases in HR and blood cell differential counts, and
atrial conduction time of rats were inconsistent. No
adverse cardiac or pulmonary effects in hamsters.

Increased expression of the proinflammatory chemokine
MP-2 in the lung and heart of MCT-treated rats; less in
healthy rats. Significant mortality only in MCT-treated
rats.
Reference
Bouthillier et al.
(1998)


Campen et al.
(2000)


Clarke et al.
(2000a)



Costa and
Dreher (1997)



Gardner et al.
(2000)
Ohio et al.
(2000a)

Godleski et al.
(2000)



Gordon et al.
(1998)

Gordon et al.
(2000)



Killingsworth
etal. (1997)



-------
       TABLE 7-8 (cont'd). CARDIOVASCULAR AND SYSTEMIC EFFECTS OF AMBIENT AND COMBUSTION-RELATED

                                          PARTICULATE MATTER
to
O
O
to















7"1
LtJ
ON



M
\^
C
^Tl
H
6
O
2,
O
H
/— s
Species, Gender,
Strain Age, or Body
Weight
Rats, male WKY and
SH, 12 to 13-week-
old


Rats, male SH and
WKY; 12-13 weeks
old


Dogs, beagles,
10. 5 -year- old,
healthy, n = 4
Rabbits, female,
New Zealand White,
1.8 to 2.4 kg


Rats, Wistar






Rats, male, S-D,
MCT-treated



Particle
ROFA




ROFA from a
precipitator of
an oil-burning
power plant

ROFA


Colloidal carbon




Ottawa ambient
(EHC-93)
(ECH-93L)
Diesel soot
(DPM)
Carbon black
(CB)
ROFA




Exposure
Technique
Nose-only
inhalation



Inhalation;
and
intratracheal
instillation

Oral
inhalation

Instillation




Inhalation
(nose only)





Instillation




Mass Particle
Concentration Size
15 mg/m3 N/A




15 mg/m3 1.5 fj.m
1 and 5 mg/kg og = 1.5



3 mg/m3 2.22 ^m MMAD
og = 2.71

2mLof 1% <1 //m
colloidal carbon
(20 mg)


48 mg/m3 36, 56, 80, 100,
49 mg/m3 and 300 ,um
5 mg/m3

5mg/m3


0.25, 1.0, or 1.95,umMMAD
2.5mgin0.3mL og = 2.19
saline


Exposure
Duration
6 h/day for
3 days



6 h/days, 3
days per week
forl, 2, or 4
weeks

3 h/day for
3 days

Examined for
24 to 192 h
after
instillation

4h






Monitored for
96 h after
instillation of
ROFA
particles
Cardiovascular Effects
Cardiomyopathy and monocytic cell infiltration, along with
increased cytokine expression, was found in left ventricle
of SH rats because of underlying cardiovascular disease.
ECG showed exacerbated ST segment depression caused
by ROFA.
IT exposure increased plasma fibrinogen and decreased
peripheral lymphocytes in both SH and WKY rats. Acute
IH exposure increased plasma fibrinogen in SH rats only;
longer exposure caused pulmonary injury but no changes in
fibrinogen.
No consistent changes in ST segment, the form or
amplitude of the T wave, or arrhythmias; slight bradycardia
during exposure.
Colloidal carbon stimulated the release of BRDU-labeled
PMNs from the bone marrow. The supernatant of alveolar
macrophages treated with colloidal carbon in vitro also
stimulated the release of PMNs from bone marrow, likely
via cytokines.
EHC-93 elevated blood pressure and ET-1 and ET-3 levels
EHC-93 L No effect on blood pressure, transient effect on
ET-1, -2, -3 levels
DPM no effect on blood pressure, but elevated ET-3 levels
CB no effect


Dose-related increases in the incidence and duration of
serious arrhythmic events in normal rats. Incidence and
severity of arrythmias were increased greatly in the MCT
rats. Deaths were seen at each instillation level in MCT
rats only (6/12 died after MCT + ROFA).
Reference
Kodavanti et al.
(2000b)



Kodavanti et al.
(2002)



Muggenburg
et al. (2000)

Terashima et al.
(1997)



Vincent et al.
(2001)





Watkinson et al.
(1998)



O
HH
H
W

-------
             TABLE 7-8 (cont'd).  CARDIOVASCULAR AND SYSTEMIC EFFECTS OF AMBIENT AND COMBUSTION-RELATED
                                                                            PARTICULATE MATTER
 to
 o
 o
 to
Species, Gender,
Strain Age, or Body
Weight
Particle
Exposure
Technique
Mass
Concentration
Particle
Size
Exposure
Duration
Cardiovascular Effects
Reference
          (l)Rats, S-D healthy
          and cold-stressed,
          ozone-treated, and
          MCT-treated

          (2) Rats, S-D, SH
          rats, WKY rats,
          healthy and
          MCT-treated,

          (3) Rats, SH,
          15-mo-old

          (4) Rats, S-D
          MCT-treated
                               ROFA
ROFA
OTT
ROFA
MSH

Fe2(S04)3
VS04
NiSO,
Intratracheal
instillation
                 Inhalation
Intratracheal
instillation
Intratracheal
instillation
                               0.0,0.25,1.0,
                               or 2.5 mg/rat
               15 mg/m3
              2.5 mg
              0.5 mg
              2.5 mg

              105 /^g
              245 ,ug
              262.5 //g
                                                 1.95 urn
                                                 1.95 urn
                                1.95 urn
Monitored for     (1) Healthy rats exposed IT to ROFA demonstrated dose-
96 h after        related hypothermia, bradycardia, and increased
instillation       arrhythmias. Compromised rats demonstrated exaggerated
                hypothermia and cardiac responses to IT ROFA. Mortality
                was seen only in the MCT-treated rats exposed to ROFA by
6 h/day for       IT. (2) Pulmonary hypertensive (MCT-treated S-D) and
3 days           systemically hypertensive (SH) rats exposed to ROFA by
                inhalation demonstrated similar effects, but of diminished
                amplitude. There were no lethalities by the inhalation
Monitored for     route. (3) Older rats exposed IT to OTT showed a
96 h after        pronounced biphasic hypothermia and a severe drop in HR
instillation       accompanied by increased arrhythmias; exposure to ROFA
                caused less pronounced, but similar effects. No cardiac
Monitored for     effects were seen with exposure to MSH. (4) Ni and V
96 h after        showed the greatest toxicity; Fe-exposed rats did not differ
instillation       from controls.
                                                                                                                      Watkinson et al.
                                                                                                                      (2000a,b)
 H
 6
 o
 o
 H
O
o
HH
H
W

-------
 1      an important role in the adverse health effects associated with episodic exposure to ambient PM,
 2      despite the "general particle" effect attributed to the epidemiological associations of ambient PM
 3      exposure and increased mortality in many regions of the United States (i.e., regions with varying
 4      particle composition). Work that examines the role of inherent susceptibility to the adverse
 5      effects of PM in compromised animal models of human pathophysiology provides a potentially
 6      important link to epidemiological observations and is discussed below.
 7           To date,  studies examining the systemic and cardiovascular effects of particles have used a
 8      number of compromised animal models, largely rodent  models.  Two studies in normal or
 9      compromised dogs (Godleski et al., 2000; Muggenburg et al., 2000) also have been published as
10      well as the preliminary results from studies in which human subjects were  exposed to
11      concentrated ambient PM (see Section 7.4.1).  Although the majority of animal studies
12      examining the systemic effects of PM have used metal-laden ROFA as a source particle, a
13      growing number of studies have used collected and stored ambient PM or real-time generated
14      concentrated ambient particles. The following discussion of the  systemic effects of PM first
15      describes the ROFA studies and then compares these findings with the ambient PM studies.
16           Killingsworth and colleagues (1997) used a fuel oil fly ash to examine the adverse effects
17      of a model urban particle in an animal model (monocrotaline, MCT) of cardiorespiratory disease;
18      MCT causes progressive lung injury and vascular inflammation in rats. The lung injury induced
19      with MCT can lead, within two weeks of treatment, to pulmonary hypertension and right heart
20      enlargement, common features of chronic obstructive pulmonary disease in humans.  They
21      observed 42% mortality in MCT rats exposed to approximately 580 //g/m3 fly ash for 6 h/day for
22      3  consecutive days. Deaths did not occur in MCT rats exposed to filtered air or in saline-treated
23      rats exposed to fly ash. The increase in deaths in the MCT/fly ash group was accompanied by an
24      increase in neutrophils in lavage fluid and an increased  immunostaining of MIP-2 in the heart
25      and lungs of the MCT/fly ash animals. Cardiac immunohistochemical analysis indicated
26      increased MIP-2 in cardiac macrophages. The fly ash-induced deaths did not result from a
27      change in pulmonary arterial pressure and the cause of death was not identified.
28           In a similar experimental model, Watkinson et al.  (1998) examined the effects of
29      intratracheally instilled ROFA (0.0, 0.25, 1.0, 2.5 mg in 3 mL saline) on ECG measurements in
30      control and MCT rats. They observed a dose-related increase in  the incidence and duration of
31      serious arrhythmic events in control  animals exposed to ROFA particles, and these effects were

        April 2002                                7-3 8        DRAFT-DO NOT QUOTE OR CITE

-------
 1      clearly exacerbated in the MCT animals. Similar to the results of Killingsworth et al. (1997),
 2      healthy animals treated with ROFA suffered no deaths, but there were 1, 2, and 3 deaths in the
 3      low-, medium-, and high-dose MCT groups, respectively. Thus, ROFA PM was linked to the
 4      conductive and hypoxemic arrhythmias associated with cardiac-related deaths in the MCT
 5      animals.
 6           To examine the biological relevance of intratracheal instillation of ROFA particles,
 7      Kodavanti et al. (1999) exposed MCT rats to ROFA by either instillation (0.83 or 3.33 mg/kg) or
 8      nose-only inhalation (15 mg/m3, 6 h/day for 3 consecutive days). Similar to Watkinson et al.
 9      (1998), intratracheal instillation of ROFA in MCT rats resulted in -50% mortality. Notably, no
10      mortality occurred in MCT rats exposed to ROFA by the inhalation route despite the high
11      exposure concentration (15 mg/m3). In addition, no mortality occurred in healthy rats exposed to
12      ROFA or in MCT rats exposed to clean air. Despite the fact that mortality was not associated
13      with ROFA inhalation exposure of MCT rats, exacerbation of lung lesions and pulmonary
14      inflammatory cytokine gene expression, as well as ECG abnormalities, clearly were evident.
15           Watkinson and colleagues further examined the effect of instilled ROFA in rodents
16      previously exposed to ozone or housed in the cold (Watkinson et al., 2000a,b; Campen et al.,
17      2000).  The effect of ozone-induced pulmonary inflammation (preexposure to 1 ppm ozone for
18      6 h) or housing in the cold (10  °C) on the response to instilled ROFA in rats was similar to that
19      produced with MCT. Bradycardia, arrhythmias, and hypothermic changes were consistently
20      observed in the ozone exposed and hypothermic animals treated with ROFA, although, unlike in
21      the MCT animals, no deaths occurred.  Thus, in rodents with cardiopulmonary disease/stress,
22      instillation of 0.25 mg or more of ROFA can produce systemic changes that may be used to study
23      potential mechanisms of toxicity that are consistent with the epidemiology and panel studies
24      showing cardiopulmonary effects in humans.
25           While studies of instilled residual oil fly ash demonstrated immediate and delayed
26      responses, consisting of bradycardia, hypothermia,  and arrhythmogenesis in conscious,
27      unrestrained rats (Watkinson et al., 1998; Campen et al., 2000), further study of instilled ROFA-
28      associated transition metals showed that vanadium  induced the immediate responses, while
29      nickel was responsible for the delayed effects (Campen et al., 2002a).  Moreover, Ni, when
30      administered concomitantly, potentiated the immediate effects caused by V.


        April 2002                                7-39        DRAFT-DO NOT QUOTE OR CITE

-------
 1           In another study, Campen et al. (2001) examined the responses to these metals in conscious
 2      rats by whole-body inhalation exposure. The authors attempted to ensure valid dosimetric
 3      comparisons with the instillation studies, by using concentrations of V and Ni ranging from 0.3-
 4      2.4 mg/m3.  The concentrations used in this study incorporated estimates of total inhalation dose
 5      derived using different ventilatory parameters.  Heart rate (HR), core temperature (T[CO]), and
 6      electrocardiographic (ECG) data were measured continuously throughout the exposure.  Animals
 7      were exposed to aerosolized Ni, V, or Ni + V for 6 h per day for 4 days, after which serum and
 8      bronchoalveolar lavage samples were taken. While Ni caused delayed bradycardia, hypothermia,
 9      and arrhythmogenesis at concentrations > 1.2 mg/m3, V  failed to induce any significant change in
10      HR or T (CO),  even at the highest concentration. When combined, Ni and V produced
11      observable delayed effects at 0.5 mg/m3 and potentiated responses at 1.3  mg/m3, greater than
12      were produced by the highest concentration of Ni (2.1 mg/m3) alone.  Although these studies
13      were performed at metal concentrations that were orders of magnitude greater than ambient
14      concentrations, the results indicate a possible synergistic relationship between inhaled Ni and V.
15           Watkinson and colleagues (2000a,b) also sought to examine the relative toxicity of
16      different particles on the cardiovascular system of spontaneously hypertensive rats.  They
17      instilled 2.5 mg of representative particles from ambient (Ottawa) or natural (Mount Saint Helens
18      volcanic ash) sources and compared the response to 0.5 mg ROFA. Instilled particles were either
19      mass equivalent dose or adjusted to produce equivalent metal dose. They observed adverse
20      changes in ECG, heart rate, and arrhythmia incidence that were much greater in the Ottawa- and
21      ROFA-treated rats than in the Mount Saint Helens-treated rats.  The cardiovascular changes
22      observed with the Ottawa particles were actually greater than with the ROFA particles. These
23      series of experiments by Watkinson and colleagues clearly demonstrate that instillation of
24      ambient air particles,  albeit at a very high concentration, can produce  cardiovascular effects.
25      They also demonstrate that PM exposures of equal mass dose did not  produce the same
26      cardiovascular  effects, suggesting that PM composition was responsible for the observed effects.
27           Because of concerns regarding the relevance of particles administered by intratracheal
28      instillation, investigators also have examined the cardiovascular effects of ROFA particles using
29      more realistic inhalation exposure protocols.  Kodavanti et al. (2000b) found that exposure to a
30      high concentration of ROFA (15 mg/m3 for 6 h/day for 3 days) produced alterations in the ECG
31      waveform of spontaneously hypertensive (SH) but not normotensive rats. Although the ST

        April 2002                                7-40        DRAFT-DO NOT QUOTE OR CITE

-------
 1      segment area of the ECG was depressed in the SH rats exposed to air, further depressions in the
 2      ST segment were observed at the end of the 6-h exposure to ROFA on Days 1 and 2.  The
 3      enhanced ST segment depression was not observed on the third day of exposure, suggesting that
 4      adaptation to the response had occurred.  Thus, exposure to a very high concentration of ROFA
 5      exacerbated a defect in the electroconductivity pattern of the heart in an animal model of
 6      hypertension.  This ROFA-induced alteration in the ECG waveform was not accompanied by an
 7      enhancement in the monocytic cell infiltration and cardiomyopathy that also develop in SH rats.
 8      Further work is necessary to determine the relevance of this ROFA study to PM at concentrations
 9      relevant to ambient exposures.
10           Godleski and colleagues (2000) have performed a series of experiments examining the
11      cardiopulmonary effects of inhaled concentrated ambient PM on normal mongrel dogs and on
12      dogs with coronary artery occlusion.  Dogs were exposed by inhalation via a tracheostomy tube
13      to concentrated ambient PM for 6 h/day for 3 consecutive days.  The investigators found little
14      biologically-relevant evidence of pulmonary inflammation or injury in normal dogs exposed to
15      PM (daily range of mean concentrations was approximately 100 to 1000 //g/m3). The only
16      statistically significant effect observed was a doubling of the percentage of neutrophils in lung
17      lavage. Despite the absence of major pulmonary effects, a significant increase in heart rate
18      variability (an index of cardiac autonomic activity), a decrease in heart rate, and an increase in T
19      alternans (an index of vulnerability to ventricular fibrillation) were observed. Exposure
20      assessment of particle composition produced no specific components of the particles that were
21      correlated with the day-to-day variability in response.  The significance of these effects is not yet
22      clear because the effects did not occur on all exposure days. For example, the change in heart
23      rate variability was observed on only 10 of the 23 exposure days. Although the heart rate
24      variability change and the increase in T alternans suggest a possible proarrhythmic response to
25      inhaled concentrated ambient PM, the clinical significance of this effect is currently unknown.
26           The most important finding in the experiments of Godleski et al. (2000) was the
27      observation of a potential increase in ischemic stress of the cardiac tissue from repeated exposure
28      to concentrated ambient PM. During coronary occlusion in four dogs exposed to PM, they
29      observed a significantly more rapid development of ST elevation of the ECG waveform.
30      In addition, the peak ST-segment elevation was greater after PM exposure.  Together, these
31      changes suggest that concentrated ambient PM can augment the ischemia associated with

        April 2002                                7-41        DRAFT-DO NOT QUOTE OR CITE

-------
 1      coronary artery occlusion in this dog model. Additional work in more dogs as well as other
 2      species is necessary to determine the significance of these findings to the human response to
 3      ambient PM.
 4           Muggenburg and colleagues (2000) reported that inhalation exposure to high
 5      concentrations of ROFA produces no consistent changes in amplitude of the ST-segment, form
 6      of the T wave, or arrhythmias in dogs.  In their studies, four beagle dogs were exposed to
 7      3 mg/m3 ROFA particles for 3 h/day for 3 consecutive days. They noted a slight but variable
 8      decrease in heart rate, but the changes were not statistically or biologically significant.  The
 9      transition metal content of the ROFA used by Muggenburg was approximately 15% by mass,
10      a value that is on the order of a magnitude higher than that found in ambient urban PM samples.
11      Although the study did not specifically address the effect of metals, it suggests that inhalation of
12      high concentrations of metals may have little effect on the cardiovascular system of a healthy
13      individual.
14           In a series of studies, Gordon, Nadziejko, and colleagues examined the  response of the
15      rodent cardiovascular system to concentrated ambient PM derived from New York City air
16      (Gordon et al., 2000). Particles of 0.2 to 2.5 //m in diameter were concentrated up to 10 times
17      their levels in ambient air (~ 150 to 900 //g/m3) to maximize possible differences in effects
18      between normal and cardiopulmonary-compromised laboratory animals. ECG changes were not
19      detected in normal Fischer 344 rats  or hamsters exposed by inhalation to concentrated ambient
20      PM for 1 to 3 days. Similarly, no deaths or ECG changes were observed in MCT rats  or
21      cardiomyopathic hamsters exposed to PM.  Contrary to the decrease in heart rate observed in
22      dogs exposed to concentrated ambient PM (Godleski  et al., 2000), heart rate was increased in
23      both normal and MCT rats exposed to PM.  The increase was approximately  5% and was not
24      observed on all exposure days. Thus, extrapolation of the heart rate changes in these animal
25      studies to human health effects is difficult, although the increase in heart rate in rats is similar to
26      that observed in some human population studies.
27           Gordon and colleagues (1998) have reported other cardiovascular effects in animals
28      exposed to inhaled CAP.  Increases in peripheral blood platelets  and neutrophils were  observed
29      in control and MCT rats at 3 h, but not 24 h, after exposure to 150 to 400 //g/m3 concentrated
30      ambient PM (CAP).  This neutrophil effect did not appear to be dose related and did not occur on
31      all exposure days, suggesting that day-to-day changes in particle  composition may play an

        April 2002                                7-42        DRAFT-DO NOT QUOTE OR CITE

-------
 1      important role in the systemic effects of inhaled particles. The number of studies reported was
 2      small and; therefore, it is not possible to statistically determine if the day-to-day variability was
 3      truly due to differences in particle composition or even to determine the size of this effect.
 4      Terashima et al. (1997) also examined the effect of particles on circulating neutrophils. They
 5      instilled rabbits with 20 mg colloidal carbon, a relatively inert particle (<1 //m), and observed a
 6      stimulation of the release of 5'-bromo-2'deoxyuridine (BrdU)-labeled PMNs from the bone
 7      marrow at 2 to 3 days after instillation. Because the instilled supernatant from rabbit AMs
 8      treated in vitro with colloidal carbon also stimulated the release of PMNs from the bone marrow,
 9      the authors hypothesized that cytokines released from activated macrophages could be
10      responsible for this systemic effect. The same research group (Tan et al., 2000) looked for
11      increased white blood cell counts as a marker for bone marrow PMN precursor release in humans
12      exposed to high levels of carbon from biomass burning during the 1997 Southeast Asian smoke-
13      haze episodes. They found a significant association between PM10 (1-day lag) and elevated band
14      neutrophil counts expressed as a percentage of total PMNs.  The biological relevance of this
15      latter study to urban PM-induced systemic effects in unclear, however,  because of the high dose
16      of carbon particles.
17           The results of epidemiology studies have suggested that homeostatic changes in the
18      vascular system can occur after episodic exposure to ambient PM.  Studies by Vincent et al.
19      (2001) indicate that urban particles inhaled by laboratory rats can affect blood levels of
20      endothelin and cause a vasopressor response without causing acute lung injury.  Moreover, the
21      potency to influence hemodynamic changes can be modified by removing the polar organic
22      compounds and soluble elements from the particles. In the study described previously
23      (Section 7.2.3), Ohio et al. (2000a) also have shown that inhalation of concentrated PM in
24      healthy nonsmokers causes increased levels of blood fibrinogen. They  exposed 38 volunteers
25      exercising intermittently at moderate levels of exertion for 2 h to either filtered air or particles
26      concentrated from the air in Chapel Hill, NC (23 to 311 //g/m2).  Blood obtained 18 h after
27      exposure contained significantly more fibrinogen than blood obtained before exposure. The
28      observed effects in blood may be associated with the mild pulmonary inflammation also found
29      18 h after exposure to CAP (see Section 7.2.3).
30           Gardner et al. (2000) examined whether the instillation of particles would alter blood
31      coagulability factors in laboratory animals. Sprague-Dawley rats were  instilled with 0.3,1.7, or

        April 2002                                7-43        DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.3 mg/kg of ROFA or 8.3 mg/kg Mount Saint Helens volcanic ash.  They observed an increase
 2      in plasma fibrinogen in healthy rats. Because fibrinogen is a known risk factor for ischemic heart
 3      disease and stroke, the authors suggested that this alteration in the coagulation pathway could
 4      take part in the triggering of cardiovascular events in susceptible individuals. Elevations in
 5      plasma fibrinogen, however, were observed in healthy rats only at the highest treatment dose; and
 6      no other changes in clotting function were noted.  Because the lower treatment doses are known
 7      to cause pulmonary injury and inflammation, albeit to a lower extent, the absence of plasma
 8      fibrinogen changes at these lower doses suggests that only high levels of pulmonary injury are
 9      able to produce an effect in healthy test animals.
10           To establish the temporal relationship between pulmonary injury, increased plasma
11      fibrinogen, and changes in peripheral lymphocytes, Kodavanti et al., (2002) exposed
12      Spontaneously Hypertensive (SH) and Wistar-Kyoto (WKY) rats to ROFA using both
13      intratracheal and inhalation exposure (acute and long-term) scenarios. Increases in plasma
14      fibrinogen and decreases in circulating white blood cells were found during the acute phase
15      responses to ROFA exposure and were temporally associated with acute, but not long-term, lung
16      injury.  A bolus intratracheal instillation of ROFA increased plasma fibrinogen in both SH and
17      WKY rats; whereas the increase was evident only in SH rats after  acute ROFA inhalation.  The
18      increased fibrinogen in SH rats corresponded to an inability to find increased pulmonary
19      glutathione and greater pulmonary injury and inflammation than was found in the WKY rats.
20           In summary, controlled  animal studies have provided initial evidence that only high
21      concentrations of inhaled or instilled particles can have systemic, especially cardiovascular,
22      effects. In the case of MCT rats, these effects can be lethal.  Controlled human exposure studies
23      also have shown ambient levels of inhaled PM can produce some biochemical and cellular
24      changes in the blood.  Although some of these biochemical  changes have been used as clinical
25      "markers" for cardiovascular diseases, the causal relationship between these changes and the
26      potential life-threatening diseases remains to be established. Understanding the pathways by
27      which very small concentrations of inhaled ambient PM can produce systemic, life-threatening
28      changes also is far from clear. Among the hypotheses that have been proposed to account for the
29      nonpulmonary effects of PM are activation of neural reflexes, cytokine effects on heart tissue
30      (Killingsworth et al., 1997), alterations in coagulability (Seaton et al., 1995;  Sjogren, 1997),
31      perturbations in both conductive and hypoxemic arrythmogenic mechanisms (Watkinson et al.,

        April 2002                                 7-44         DRAFT-DO NOT QUOTE OR CITE

-------
 1      1998; Campen et al., 2000), and altered endothelin levels (Vincent et al., 2001).  A great deal of
 2      research using controlled exposures of laboratory animals and human subjects to PM will be
 3      necessary to test mechanistic hypotheses generated to date, as well as those that are likely to be
 4      proposed in the future (see Section 7.5).
 5
 6
 7      7.4 SUSCEPTIBILITY TO THE EFFECTS OF PARTICULATE
 8          MATTER EXPOSURE
 9          Susceptibility of an individual to adverse health effects of PM can vary depending on a
10      variety of host factors such as age, physiological activity profile, genetic predisposition, or
11      preexistent disease.  The potential for preexistent disease to alter adverse response to toxicant
12      exposure is widely acknowledged but poorly understood. Because of inherent variability
13      (necessitating large numbers of subjects) and ethical concerns associated with using diseased
14      subjects in clinical research studies, a solid database on human susceptibilities is lacking. For
15      more control over both host and environmental variables, animal models often are used.
16      However, care must be taken in extrapolation from animal models of human disease to humans.
17      Rodent models of human disease, their use in toxicology, and the criteria for judging their
18      appropriateness as well as their limitations must be considered (Kodavanti et al., 1998b;
19      Kodavanti and Costa, 1999).
20
21      7.4.1  Pulmonary Effects of Particulate Matter in Compromised Hosts
22          Epidemiological studies suggest there may be subsegments of the population that are
23      especially susceptible to effects from inhaled particles (see Chapter 8). The elderly with chronic
24      cardiopulmonary disease, those with pneumonia and possibly other lung infections, and those
25      with asthma (at any age) appear to be at higher risk than healthy people of similar age.
26      Unfortunately, most toxicology studies have used healthy adult animals.  An increasing number
27      of newer studies have examined effects of ambient particles in compromised host models. Costa
28      and Dreher (1997) used a rat model of cardiopulmonary disease to explore the question of
29      susceptibility and the possible mechanisms by which PM effects are potentiated. Rats with
30      advanced monocrotaline (MCT)-induced pulmonary vasculitis/hypertension were given
31      intratracheal instillations of ROFA (0, 0.25, 1.0, and 2.5 mg/rat). The MCT animals had a
        April 2002                               7-45        DRAFT-DO NOT QUOTE OR CITE

-------
 1      marked neutrophilic inflammation.  In the context of this inflammation, ROFA induced a four- to
 2      fivefold increase in BAL PMNs.  There was increased mortality at 96 h that was ROFA-dose
 3      dependent. The results of this study indicate that particles, albeit at a high concentration,
 4      enhanced mortality in MCT animals but not in healthy animals.
 5           As discussed previously, Kodavanti et al. (1999) also studied PM effects in the MCT rat
 6      model of pulmonary disease. Rats treated with 60 mg/kg MCT were exposed to 0, 0.83. or 3.3
 7      mg/kg ROFA by intratracheal instillation and to 15 mg/m3 ROFA by inhalation.  Both methods
 8      of exposure caused inflammatory lung responses; and ROFA exacerbated the lung lesions, as
 9      shown by increased lung edema, inflammatory cells, and alveolar thickening.
10           The manner in which MCT can  alter the response of rats to inhaled particles was examined
11      by Madl and colleagues (1998).  Rats  were exposed to fluorescent colored microspheres (1 //m)
12      2 weeks after treatment with MCT.  In vivo phagocytosis of the microspheres was altered in the
13      MCT rats in comparison with control  animals. Fewer microspheres were phagocytized in vivo
14      by alveolar macrophages, and there was a concomitant increase in free microspheres overlaying
15      the epithelium at airway bifurcations.  The decrease in in vivo  phagocytosis was not accompanied
16      by a similar decrease in vitro.  Macrophage chemotaxis, however, was impaired  significantly in
17      MCT rats compared with control rats. Thus, MCT appeared to impair particle clearance from the
18      lungs via inhibition of macrophage  chemotaxis.
19           The sulfur dioxide (SO2)-induced model of chronic bronchitis has also been used to
20      examine the potential interaction of PM with preexisting lung injury. Clarke and colleagues
21      pretreated Sprague-Dawley rats for 6 weeks with air or 170 ppm SO2 for 5 h/day and 5 days/week
22      (Clarke et al., 1999). Exposure to concentrated ambient air particles for 5 h/day for 3 days at an
23      average concentration of 515 //g/m3 produced significant changes in both cellular and
24      biochemical markers in lavage fluid.  In comparison to control animal values, protein was
25      increased approximately threefold in SO2-pretreated animals exposed to concentrated ambient
26      PM. Lavage fluid neutrophils and lymphocytes were increased significantly in both groups of
27      rats exposed to concentrated ambient PM, with greater increases in both cell types in the
28      SO2-pretreated rats. Thus, exposure to concentrated ambient PM produced adverse changes in
29      the respiratory system, but no deaths,  in both normal rats and in a rat model of chronic bronchitis.
30           Clarke et al. (2000b) next examined the effect of concentrated ambient PM from Boston,
31      MA, in normal rats of different ages.  Unlike the earlier study that used Sprague-Dawley rats,

        April 2002                                7-46        DRAFT-DO NOT QUOTE OR CITE

-------
 1      4- and 20-mo-old Fischer 344 rats were examined after exposure to concentrated ambient PM for
 2      5 h/day for 3 consecutive days. They found that exposure to the daily mean concentrations of 80,
 3      170, and 50 //g/m3 PM, respectively, produced statistically significant increases in total
 4      neutrophil counts (over 10-fold) in lavage fluid of the young, but not the old, rats. Thus, repeated
 5      exposure to relatively low concentrations of ambient PM produced an inflammatory response,
 6      although the actual percent neutrophils in the concentrated ambient PM-exposed young adult rats
 7      was low (approximately 3%). On the other hand, Gordon et al. (2000) found no evidence of
 8      neutrophil influx in the lungs of normal and monocrotaline-treated Fischer 344 rats exposed in
 9      nine separate experiments to concentrated ambient PM from New York, NY, as  high as
10      400 //g/m3 for a 6-h exposure or 192 //g/m3 for three daily 6-h exposures. Similarly, normal and
11      cardiomyopathic hamsters showed no evidence of pulmonary inflammation or injury after a
12      single exposure to the same levels of concentrated ambient PM. Gordon and colleagues did
13      report a statistically significant doubling in  protein concentration in lavage fluid in
14      monocrotaline-treated rats exposed for 6 h to 400 //g/m3 concentrated ambient PM. Because of
15      the disparity in findings in the response of normal Fischer 344 rats to concentrated ambient PM
16      between these two labs, it is important that the reproducibility of these experiments be examined.
17           Kodavanti and colleagues (1998b) also have examined the effect of concentrated ambient
18      PM in normal rats and rats with sulfur dioxide-induced chronic bronchitis. Among the four
19      separate exposures to PM, there was a significant increase in lavage fluid protein in bronchitic
20      rats from only one exposure protocol in which the rats were exposed  to 444 and  843 //g/m3 PM
21      on 2 consecutive days (6 h/day). Neutrophil counts were increased in bronchitic rats exposed to
22      concentrated ambient PM in three of the four exposure protocols, but was decreased in the fourth
23      protocol.  No other changes in normal or bronchitic rats were observed, even in the exposure
24      protocols with higher PM concentrations. Thus, rodent studies have demonstrated that
25      inflammatory changes can be produced in normal and compromised animals exposed to
26      concentrated ambient PM. These findings are important because only a limited number of
27      studies have used real-time inhalation exposures to actual ambient urban PM.
28           Pulmonary function measurements are often less invasive than other means to assess the
29      effects of inhaled air pollutants on the mammalian lung.  After publication of the 1996 PM
30      AQCD, a number of investigators examined the response of rodents and dogs to inhaled ambient
31      particles. In general, these investigators have demonstrated that ambient PM has minimal effects

        April 2002                                7-47        DRAFT-DO NOT QUOTE OR CITE

-------
 1      on pulmonary function tests. Gordon et al. (2000) exposed normal and monocrotaline-treated
 2      rats to filtered air or 181 //g/m3 concentrated ambient PM for 3 h.  For both normal and
 3      monocrotaline-treated rats, no differences in lung volumes or diffusion capacities for carbon
 4      monoxide were observed between the air or PM exposed animals at 3 or 24 h after exposure.
 5      Similarly, in cardiomyopathic hamsters, concentrated ambient PM had no effect on these same
 6      pulmonary function measurements.
 7           Other pulmonary function endpoints have been studied in animals exposed to concentrated
 8      ambient PM. Clarke et al. (1999) observed that tidal volume was increased slightly in both
 9      control rats and rats with sulfur dioxide-induced chronic bronchitis exposed to 206 to 733 //g/m3
10      PM on 3 consecutive days.  No changes in peak expiratory flow, respiratory frequency, or minute
11      volume were observed after exposure to concentrated ambient PM. In the series of dog studies
12      by Godleski et al. (2000) (also see Section 7.3), no signficant changes in pulmonary function
13      were observed in normal mongrel dogs exposed to concentrated ambient PM, although a 20%
14      decrease in respiratory frequency was observed in dogs that underwent coronary artery occlusion
15      and were exposed to PM. Thus, studies using normal and compromised animal models exposed
16      to concentrated ambient PM have found minimal biological effects of ambient PM on pulmonary
17      function.
18           Johnston et al. (1998)  exposed 8-week-old mice (young) and 18-mo-old mice (old) to
19      polytetrafluoroethylene (PTFE) fumes (0, 10, 25, and 50 //g/m3) for 30 min.  Lung lavage
20      endpoints (PMN, protein, LDH, and p-glucuronidase) as well as lung tissue mRNA levels for
21      various cytokines, metallothionein and for Mn superoxide dismutase were measured 6 h
22      following exposure. Protein, lymphocyte, PMN, and TNF-a mRNA levels were increased in
23      older mice when compared  to younger mice. These findings suggest that the inflammatory
24      response to PTFE fumes is altered with age, being greater in the older animals. Although
25      ultrafme PTFE fumes are not a valid surrogate for ambient ultrafme particles (Oberdorster et  al.,
26      1992), this study did provide evidence to support the hypothesis that particle-induced pulmonary
27      inflammation is different between young and old mice.  Further studies on age-related PM effects
28      are described in Section 7.6 (Responses to PM and Gaseous Pollutant Mixtures).
29           Kodavanti et al.  (2000b; 2001) used genetically predisposed spontaneously hypertensive
30      (SH) rats as a model of cardiovascular disease to study PM-related susceptibility.  The SH rats
31      were found to be more susceptible to acute pulmonary injury from intratracheal ROFA exposure

        April 2002                                7-48       DRAFT-DO NOT QUOTE OR CITE

-------
 1     than normotensive control Wistar Kyoto (WKY) rats (Kodavanti et al., 2001).  The primary
 2     metal constituents of ROFA, V and Ni, caused differential species-specific effects. Vanadium,
 3     which was less toxic than Ni in both strains, caused inflammatory responses only in WKY rats;
 4     whereas Ni was injurious to both WKY and SH rats (SH > WKY).  This differential
 5     responsiveness of V and Ni was correlated with their specificity for airway and parenchymal
 6     injury, discussed in another study (Kodavanti et al., 1998b).  When exposed to the same ROFA
 7     by inhalation,  SH rats were more sensitive than WKY rats in regards to vascular leakage
 8     (Kodavanti et  al., 2000b).  The SH rats exhibited a hemorrhagic response to ROFA.  Oxidative
 9     stress was much higher in ROFA exposed SH rats than matching WKY rats. Also, SH rats,
10     unlike WKY rats, showed a compromised ability to increase BALF glutathione in response to
11     ROFA, suggesting a potential link to increased susceptibility. Cardiovascular effects were
12     characterized by ST-segment area depression of the ECG in ROFA-exposed SH but not WKY
13     rats. When the same rats were exposed to ROFA by inhalation (Kodavanti et al., 2002),
14     differences in  effects were found depending on the length of exposure. After acute exposure,
15     increased plasma fibrinogen was associated with lung injury; longer-term, episodic ROFA
16     exposure resulted in progressive protein leakage and inflammation that was significantly worse in
17     SH rats when compared to WKY rats.  These studies demonstrate the potential utility of
18     cardiovascular disease models for the study of PM health effects and show that genetic
19     predisposition to oxidative stress and cardiovascular disease may play a role in sensitivity to
20     increased PM-related cardiopulmonary injury.
21           On the basis of in vitro studies, Sun et al. (2001) predicted that the antioxidant and lipid
22     levels in the lung lining fluid may determine susceptibility to inhaled PM. In a subsequent study
23     from the same laboratory, Norwood et al. (2001) conducted inhalation studies on guinea pigs to
24     test this hypothesis. The guinea pigs were divided on the basis of dietary supplementation or
25     depletion of ascorbic acid (C) and glutathione (GSH) into four groups: (+C+GSH), (+C-GSH),
26     (-C+GSH), and (-C-GSH). All groups were exposed, nose-only, to clean air or 19-25 mg/m3
27     ROFA (< 2.5 jim) for 2 h. Nasal lavage and BAL fluid and cells were examined at 0 h and 24 h
28     postexposure.  Exposure to ROFA increased lung injury in the (-C-GSH) group only, as shown
29     by increased BAL fluid protein, LDH, and PMNs and decreased BAL macrophages, and resulted
30     in lower antioxidant concentrations in BAL fluid than were found with single deficiencies.


       April 2002                                7-49        DRAFT-DO NOT QUOTE OR CITE

-------
 1           In summary, although more of these studies are just emerging and are only now being
 2      replicated or followed more thoroughly to investigate the mechanisms, they do provide evidence
 3      of enhanced susceptibility to inhaled PM in "compromised" hosts.
 4
 5      7.4.2 Genetic Susceptibility to Inhaled Particles and their Constituents
 6           A key question in understanding the adverse health effects of inhaled PM is which
 7      individuals are susceptible to PM. Although factors such as age and health status have been
 8      studied in both epidemiology and toxicology studies, a number of investigators have begun to
 9      examine the importance of genetic susceptibility in the response to inhaled particles because of
10      considerable evidence that genetic factors play a role in the response to inhaled pollutant gases.
11      To accomplish this goal, investigators typically have studied the interstrain response to particles
12      in rodents. The response to ROFA instillation in different strains of rats has been investigated by
13      Kodavanti et al. (1996, 1997a).  In the first study, male Sprague-Dawley (SD) and Fischer-344
14      (F-344) rats were instilled intratracheally with saline or ROFA particles. ROFA instillation
15      produced an increase in lavage fluid neutrophils in both SD  and F-344 rats; whereas a time-
16      dependent increase in eosinophils occurred only in SD rats.  In the  subsequent study (Kodavanti
17      et al., 1997a),  SD, Wistar (WIS), and F-344 rats (60 days old) were exposed to saline or ROFA
18      (8.3 mg/kg) by intratracheal instillation and examined for up to 12  weeks.  Histology indicated
19      focal areas of lung damage showing inflammatory cell infiltration as well as alveolar, airway, and
20      interstitial thickening in all three rat strains during the week following exposure.  Trichrome
21      staining for fibrotic changes indicated a sporadic incidence of focal alveolar fibrosis at 1, 3, and
22      12 weeks in SD rats; whereas WIS and F-344 rats showed only a modest increase in trichrome
23      staining in the septal areas.  One of the isoforms of fibronectin mRNA was upregulated in
24      ROFA-exposed SD and WIS rats, but not in F-344 rats. Thus, in rats there appears to be a
25      genetic based difference in susceptibility to lung injury induced by  instilled ROFA.
26           Differences in the degree of pulmonary inflammation have been described  in rodent strains
27      exposed to airborne pollutants.  To understand the underlying causes, signs of airway
28      inflammation (i.e., airway hyper-responsiveness,  inflammatory cell influx) were  established in
29      responsive (BABL/c) and non-responsive (C57BL/6) mouse strains exposed to ROFA (Veronesi
30      et al., 2000). Neurons taken from the ganglia (i.e., dorsal root ganglia) that innervate the nasal
31      and upper airways were cultured from each mouse strain and exposed to ROFA.  The difference
        April 2002                                7-50        DRAFT-DO NOT QUOTE OR CITE

-------
 1      in inflammatory response noted in these mouse strains in vivo was retained in culture, with
 2      C57BL/6 neurons showing significantly lower signs of biological activation (i.e., increased
 3      intracellular calcium levels) and cytokine (i.e., IL-6, IL-8) release relative to BALB/c mice.
 4      RT-PCR and immunocytochemistry indicated that the BALB/c mouse strain had a significantly
 5      higher number of neuropeptide and acid-sensitive (i.e., NK1, VR1) sensory receptors on their
 6      sensory ganglia relative to the C57BL/6 mice. Such data indicate that genetically-determined
 7      differences in sensory inflammatory receptors can influence the degree of PM-induced airway
 8      inflammation.
 9           Kleeberger and colleagues have examined the role that genetic susceptibility plays in the
10      effect of inhaled acid-coated particles on macrophage function. Nine inbred strains of mice were
11      exposed nose-only to carbon particles coated with acid (10 mg/m3 carbon with 285 //g/m3 sulfate)
12      for 4 h (Ohtsuka et al., 2000a).  Significant inter-strain differences in Fc-receptor-mediated
13      macrophage phagocytosis were observed, with C57BL/6J mice being the most sensitive.
14      Although neutrophil counts were increased more in C3H/HeOuJ and C3H/HeJ strains of mice
15      than in the other strains, the overall magnitude of change was small and not correlated with the
16      changes in macrophage phagocytosis.  In follow-up studies using the same type particle, Ohtsuka
17      et al. (2000a,b) performed a genome-wide scan with an intercross cohort derived from C57BL/6J
18      and C3H/HeJ mice.  Analyses of macrophage dysfunction phenotypes of segregant and
19      nonsegregant populations derived from these two strains indicate that two unlinked genes control
20      susceptibility. They identified a 3-centiMorgan segment on mouse chromosome 17 that contains
21      an acid-coated particle susceptibility locus.  Interestingly, this quantitative trait locus overlaps
22      with those described for ozone-induced inflammation (Kleeberger et al., 1997) and acute lung
23      injury (Prows et al., 1997) and contains several promising candidate genes that may be
24      responsible for the observed genetic susceptibility for macrophage dysfunction in mice exposed
25      to acid-coated particles.
26           Leikauf and colleagues (Leikauf et al., 2000; Wesselkamper et al., 2000; McDowell et al.,
27      2000; Prows and Leikauf, 2001; Leikauf et al., 2001) have identified a genetic susceptibility in
28      mice that is associated with mortality following exposures to high concentrations (from 15 to 150
29      |ig/m3) of a MSO4 aerosol  (0.22 |im MMAD) for up to 96 h.  These studies also have
30      preliminarily identified the chromosomal locations of a small number of genes that may be
31      responsible for this genetic susceptibility. This finding is particularly significant in light of the

        April 2002                               7-51        DRAFT-DO NOT QUOTE OR CITE

-------
 1      toxicology studies demonstrating that bioavailable, first-row transition metals participate in the
 2      acute lung injury following exposure to emission and ambient air particles. Similar genes may be
 3      involved in human responses to particle-associated metals.  However, additional studies will be
 4      required to determine whether the identified metal susceptibility genes are involved in human
 5      responses to ambient levels of particulate-associated metals.
 6           One study has examined the interstrain susceptibility to ambient particles.  C57BL/6J and
 7      C3H/HeJ mice were exposed to 250 //g/m3 concentrated ambient PM2 5 for 6 h and examined at
 8      0 and 24 h after exposure for changes in lavage fluid parameters and cytokine mRNA expression
 9      in lung tissue (Shukla et al., 2000). No interstrain differences in response were observed.
10      Surprisingly, although no indices of pulmonary inflammation or injury were increased over
11      control values in the lavage fluid, increases in cytokine mRNA expression were observed in both
12      murine strains exposed to PM2 5. Although the increase in cytokine mRNA expression was
13      generally small (approximately twofold), the effects on IL-6, TNF-a, TGF-P2, and y-interferon
14      were consistent.
15           Thus, a handful of studies have begun to demonstrate that genetic susceptibility can play a
16      role in the response to inhaled particles. However, the doses of PM administered in these
17      studies, whether by inhalation or instillation, were extremely high when compared to ambient
18      PM levels. Similar strain differences in response to inhaled metal particles have been  observed
19      by other investigators (McKenna et al., 1998; Wesselkamper et al., 2000), although the
20      concentration of metals used in these studies were also more relevant to occupational rather than
21      environmental exposure levels.  It remains to be determined whether genetic susceptibility plays
22      as significant a role in the adverse effects  of ambient PM as does age or health status.
23
24      7.4.3 Effect of Particulate Matter on Allergic  Hosts
25           Relatively little is known about the effects of inhaled particles on humoral (antibody) or
26      cell-mediated immunity. Alterations in the response to a specific antigenic challenge have been
27      observed in animal models at high concentrations of acid sulfate aerosols (above 1,000 //g/m3)
28      (Pinto et al., 1979; Kitabatake et al., 1979; Fujimaki et al., 1992). Several studies have reported
29      an enhanced response to nonspecific bronchoprovocation agents, such  as acetylcholine and
30      histamine, after exposure to inhaled particles. This nonspecific airway hyperresponsiveness,
31      a central feature of asthma, occurs in animals and human subjects exposed to sulfuric acid  under
        April 2002                                7-52         DRAFT-DO NOT QUOTE OR CITE

-------
 1      controlled conditions (Gearhart and Schlesinger, 1986; Utell et al., 1983). Although, its
 2      relevance to specific allergic responses in the airways of atopic individuals is unclear, it
 3      demonstrates that the airways of asthmatics may become sensitized to either specific or
 4      nonspecific triggers that could result in increases in asthma severity and asthma-related hospital
 5      admissions  (Peters et al., 1997; Jacobs et al., 1997; Lipsett et al., 1997). Combustion particles
 6      also may serve as carrier particles for allergens (Knox et al., 1997).
 7           A number of in vivo and in vitro studies have demonstrated that DPM can alter the immune
 8      response to  challenge with specific antigens and suggest that DPM may act as an adjuvant.
 9      These studies have shown that treatment with DPM enhances the secretion of antigen-specific
10      IgE in mice (Takano et al., 1997) and in the nasal cavity of human subjects (Diaz-Sanchez et al.,
11      1996, 1997; Ohtoshi et al., 1998). Because IgE levels play a major role in allergic asthma
12      (Wheatley and Platts-Mills, 1996), upregulation of its production could lead to an increased
13      response to  inhaled antigen in particle-exposed individuals.
14           Van Zijverden et al.  (2000a,b) used mouse models to assess the potency of particles to
15      adjuvate an immune response to a protein antigen. All particles exert an adjuvant effect on the
16      immune response to co-administered antigen, apparently  stimulated by the particle core rather
17      than the attached chemical factors. Different particles, however, stimulate distinct types of
18      immune responses.  In  one model (Van Zijverden et al., 2001), BALB/c mice were intranasally
19      treated with a mixture of antigen (model antigen TNP-Ovalbumin, TNP-OVA) and particles on
20      three consecutive days.  On day 10 after sensitization mice were challenged with the antigen
21      TNP-OVA  alone and five days later the immune response was assessed. DPM, as well as carbon
22      black particles  (CB), were capable of adjuvating the immune response to TNP-OVA as
23      evidenced by an increase of TNP-specific antibody (IgGl and IgE) secreting B cells antibodies in
24      the lung-draining lymph nodes. Increased antigen-specific IgGl, IgG2a, and IgE isotypes were
25      measured in the serum, indicating that the response resulted in systemic sensitization.
26      Importantly, an increase of eosinophils in the bronchio-alveolar lavage was observed with CB.
27      Companion studies with the intranasal exposure  model showed that the adjuvant effect of
28      particles (CB) was even more pronounced when the particles were given during both the
29      sensitization and challenge phases; whereas administration during the challenge phase caused
30      only marginal changes  on the immune response.  These data show that particulate matter can
31      increase both the sensitization  and challenge responses to a protein antigen, and the  immune

        April 2002                                 7-53         DRAFT-DO NOT QUOTE OR CITE

-------
 1      stimulating activity of particles appears to be a time-dependent process, suggesting that an
 2      inflammatory microenvironment, such as may be created by the particles, is crucial for enhancing
 3      sensitization by particles.
 4           Only a small number of studies have examined the mechanisms underlying the
 5      enhancement of allergic asthma by ambient urban particles.  Ohtoshi et al. (1998) reported that a
 6      coarse size-fraction of resuspended ambient PM, collected in Tokyo, induced the production of
 7      granulocyte macrophage colony stimulating factor (GMCSF), an upregulator of dendritic cell
 8      maturation and lymphocyte function, in human airway epithelial cells in vitro.  In addition to
 9      increased GMCSF, epithelial cell supernatants contained increased IL-8 levels when incubated
10      with DPM, a principal component of ambient particles collected in Tokyo. Although the sizes of
11      the two types of particles used in this study were not comparable, the results suggest that ambient
12      PM, or at least the DPM component of ambient PM, may be able to upregulate the immune
13      response to inhaled antigen through GMCSF production.  Similarly, Takano et al. (1998) has
14      reported airway inflammation, airway hyperresponsiveness, and increased GMGSF and IL-5 in
15      mice exposed to diesel exhaust.
16           In a study by Walters et al. (2001),  PM10 was found to induce airway hyperresponsiveness,
17      suggesting that PM exposure may be an important factor in increases in asthma prevalence.
18      Naive mice were exposed to a single dose (0.5 mg/ mouse) of ambient PM, coal fly ash, or diesel
19      PM.  Exposure to PM10 induced increases in airway responsiveness and BAL cellularity; whereas
20      diesel PM induced significant increases in BAL cellularity, but not airway responsiveness.
21      On the other hand, coal fly ash exposure did not elicit significant changes in either of these
22      parameters. Ambient PM-induced airway hyperresponsiveness was sustained over 7 days. The
23      increase in airway responsiveness was preceded by increases in BAL eosinophils; whereas a
24      decline in airway responsiveness was associated with increases in macrophages. Thus, ambient
25      PM can induce asthma-like parameters in naive mice.
26           In an examination  of the effect of concentrated ambient PM on airway responsiveness in
27      mice, Goldsmith and colleagues (1999) exposed control and ovalbumin-sensitized mice to an
28      average concentration of 787 //g/m3 PM for 6 h/day for 3 days.  Although ovalbumin
29      sensitization itself produced an increase in the nonspecific airway responsiveness to inhaled
30      methylcholine, concentrated ambient PM did not change the response to methylcholine in
31      ovalbumin-sensitized or control mice.  For comparison, these investigators examined the effect

        April 2002                                7-54        DRAFT-DO NOT QUOTE OR CITE

-------
 1      of inhalation of an aerosol of the active soluble fraction of ROFA on control and ovalbumin-
 2      sensitized mice and demonstrated that ROFA could produce nonspecific airway
 3      hyperresponsiveness to methylcholine in both control and ovalbumin-sensitized mice. Similar
 4      increases in airway responsiveness have been observed after exposure to ROFA in normal and
 5      ovalbumin-sensitized rodents (Gavett et al., 1997, 1999; Hamada et al., 1999, 2000).
 6           Gavett et al. (1999) have investigated the effects of ROFA (intratracheal instillation) in
 7      ovalbumin (OVA) sensitized and challenged mice. Instillation of 3 mg/kg (approximately 60 //g)
 8      ROFA induced inflammatory and physiological responses in the OVA mice that were related to
 9      increases in Th2 cytokines (IL-4, IL-5).  Compared to OVA sensitization alone, ROFA induced
10      greater than additive increases in eosinophil numbers and in airway responsiveness to
11      methylcholine.
12           Hamada et al. (1999, 2000) have examined the effect of a ROFA leachate aerosol in a
13      neonatal mouse model of allergic asthma.  In the first study, neonatal  mice sensitized by
14      intraperitoneal (ip) injection with OVA developed airway hyperresponsiveness, eosinophilia, and
15      elevated serum anti-ovalbumin IgE after a challenge with inhaled OVA. Exposure to the ROFA
16      leachate aerosol had no marked effect on the airway responsiveness to inhaled methacholine in
17      nonsensitized  mice, but did enhance the airway hyperresponsiveness to methylcholine produced
18      in OVA-sensitized mice.  No other interactive effects of ROFA exposure with OVA were
19      observed. In a subsequent study, Hamada et al. clearly demonstrated  that, whereas inhaled OVA
20      alone was not  sufficient to sensitize mice to a subsequent inhaled OVA challenge, pretreatment
21      with a ROFA leachate aerosol prior to the initial exposure to aerosolized OVA resulted in an
22      allergic response to the inhaled OVA challenge.  Thus, exposure to a ROFA leachate aerosol can
23      alter the immune response to inhaled OVA both at the sensitization stage at an early age  and at
24      the challenge stage.
25           Lambert et al. (1999) also examined the effect of ROFA on a rodent model of pulmonary
26      allergy.  Rats were instilled intratracheally with 200  or 1,000 //g ROFA 3 days prior to
27      sensitization with house dust mite (HDM) antigen. HDM sensitization after 1000 //g ROFA
28      produced increased eosinophils, LDH, BAL protein, and IL-10 relative to HDM alone. The
29      immediate bronchoconstrictive and associated antigen-specific IgE response to a subsequent
30      antigen challenge was increased in the ROFA-treated group in comparison with the control
31      group. Together, these studies  suggest the components of ROFA can augment the immune

        April 2002                                7-55        DRAFT-DO NOT QUOTE OR CITE

-------
 1      response to antigen.  Evidence that metals are responsible for the ROFA-enhancement of an
 2      allergic sensitization was demonstrated by Lambert et al. (2000). In this follow-up study, Brown
 3      Norway rats were instilled with 1 mg ROFA or the three main metal components of ROFA (iron,
 4      vanadium, or nickel) prior to sensitization with instilled house dust mite.  The three individual
 5      metals were found to augment different aspects of the immune response to house dust mite.
 6      Nickel and vanadium produced an enhanced immune response to the antigen as seen by higher
 7      house dust mite-specific IgE serum levels after an antigen challenge at 14 days after sensitization.
 8      Nickel and vanadium also produced an increase in the lymphocyte proliferative response to
 9      antigen in vitro. In addition, the antigen-induced bronchoconstrictive response was greater only
10      in nickel-treated rats. Thus, instillation of metals at concentrations equivalent to those present in
11      the ROFA leachate mimicked the response to ROFA, suggesting that the metal components of
12      ROFA are responsible for the increased allergic sensitization observed in ROFA-treated animals.
13           Although these studies demonstrate that inhalation or instillation of ROFA augments the
14      immune response in  allergic hosts, the applicability of these findings to ambient PM is an
15      important consideration.  Goldsmith et al. (1999) have compared the effect of inhalation of
16      concentrated ambient PM for 6 h/day for 3 days versus the effect of a single exposure to a ROFA
17      leachate aerosol on the airway responsiveness to methylcholine in OVA-sensitized mice.
18      Exposure to ROFA leachate aerosols significantly enhanced the airway hyperresponsiveness in
19      OVA-sensitized mice; whereas, exposure to concentrated ambient PM (average concentration of
20      787 //g/m3) had no effect on airway responsiveness in six separate experiments. Thus, the effect
21      of the ROFA leachate aerosols on the induction of airway hyperresponsiveness in allergic mice
22      was significantly different than that of a high concentration of concentrated ambient PM.
23      Although airway responsiveness was examined at only one post-exposure time point,  these
24      findings do suggest that a great deal of caution should be used in interpreting the results of
25      studies using ROFA  particles or leachates in the attempt to investigate the biologic plausibility of
26      the adverse health effects of PM.
27           Several other studies have examined in greater detail the contribution of the particle
28      component and the organic fraction of DPM to allergic  asthma.  Tsien et al. (1997) treated
29      transformed IgE-producing human B lymphocytes in vitro with the  organic extract of DPM.  The
30      organic phase extraction had no effect on cytokine production but did increase IgE production.
31      In these in vitro experiments, DPM appeared to be acting on cells already committed to IgE

        April 2002                                7-56        DRAFT-DO NOT QUOTE OR CITE

-------
 1      production, thus suggesting a mechanism by which the organic fraction of combustion particles
 2      can directly affect B cells and influence human allergic asthma.
 3           Cultured epithelial cells from atopic asthmatics show a greater response to DPM exposure
 4      when compared with cells from nonatopic nonasthmatics.  IL-8, GM-CSF, and soluble ICAM-1
 5      increased in response to DPM at a concentration of 10 //g/mL DPM (Bayram et al., 1998a,b).
 6      This study suggests that particles could modulate airway disease through their actions on airway
 7      epithelial cells. This study also suggests that bronchial epithelial cells from asthmatics are
 8      different from those of nonasthmatics in regard to their mediator release in response to DPM.
 9           Sagai and colleagues (1996) repeatedly instilled mice with DPM for up to 16 weeks and
10      found increased numbers of eosinophils, goblet cell hyperplasia, and nonspecific airway
11      hyperresponsiveness, changes which are central features of chronic asthma (National Institutes of
12      Health, 1997).  Takano et al. (1997) extended this line of research and examined the effect of
13      repeated instillation of DPM on the antibody response to antigen (OVA) in mice.  They observed
14      that antigen-specific IgE and IgG levels were significantly greater in mice repeatedly instilled
15      with both DPM and OVA. Because this upregulation in antigen-specific immunoglobulin
16      production was not accompanied by an increase in inflammatory cells or cytokines in lavage
17      fluid, it would suggest that, in vivo, DPM may act directly on immune system cells, as described
18      in the work by Tsien et al. (1997). Animal studies have confirmed that the adjuvant activity of
19      DPM also applies to the sensitization of Brown-Norway rats to timothy grass pollen (Steerenberg
20      etal., 1999).
21           Diaz-Sanchez and colleagues (1996) have continued to study the mechanism of DPM-
22      induced upregulation of allergic response in the nasal cavity of human subjects. In one study, a
23      200 //L aerosol bolus containing 0.15 mg of DPM was delivered into each naris of subjects with
24      or without seasonal allergies.  In addition to increases in IgE in nasal lavage fluid (NAL),  they
25      found an enhanced production of IL-4, IL-6, and IL-13, cytokines known to be B cell
26      proliferation factors. The levels of several other cytokines also were increased, suggesting a
27      general inflammatory response to a nasal challenge with DPM. In a following study, these
28      investigators delivered ragweed antigen, alone or in combination with DPM, on two occasions, to
29      human subjects with both allergic rhinitis and positive skin tests to ragweed (Diaz-Sanchez et al.
30      1997).  They found that the combined challenge with ragweed antigen and DPM produced
31      significantly greater antigen-specific IgE and IgG4 in NAL. A peak response was seen at 96 h

        April 2002                                7-57        DRAFT-DO NOT QUOTE OR CITE

-------
 1      postexposure. The combined treatment also induced expression of IL-4, IL-5, IL-10, and IL-13,
 2      with a concomitant decrease in expression of Thl-type cytokines.  Although the treatments were
 3      not randomized (antigen alone was given first to each subject), the investigators reported that
 4      pilot work showed no interactive effect of repeated antigen challenge on cellular and biochemical
 5      markers in NAL.  DPM also resulted in the nasal influx of eosinophils, granulocytes, monocytes,
 6      and lymphocytes, as well as the production of various inflammatory mediators.  The combined
 7      DPM plus ragweed exposure did not increase the rhinitis symptoms beyond those of ragweed
 8      alone. Thus, diesel exhaust (particles and gases) can produce an enhanced response to antigenic
 9      material in the nasal cavity.
10           Extrapolation of these findings of enhanced allergic response in the nose to the human lung
11      would suggest that ambient combustion particles containing DPM may have significant effects
12      on allergic asthma. A study by Nordenhall et al. (2001) has addressed the effects of diesel PM on
13      airway hyperresponsiveness, lung function and airway inflammation in a group of atopic
14      asthmatics with stable disease. All were hyperresponsive to methacholine. Each subject was
15      exposed to DPM (300 //g/m3) and air for 1 h on two separate occasions. Lung function was
16      measured before and immediately after the exposures.  Sputum induction was performed 6 h, and
17      methacholine inhalation test 24 h, after each exposure. Exposure to DE was associated with a
18      significant increase in the degree of hyperresponsiveness, as compared to after air, a significant
19      increase in airway resistance and in sputum levels of interleukin (IL)-6 (p=0.048). No changes
20      were detected in sputum levels of methyl-histamine, eosinophil cationic protein,
21      myeloperoxidase, and IL-8.
22           These studies provide biological  plausibility for the exacerbation of allergic asthma
23      associated with episodic exposure to PM. Although DPM may make up only a fraction of the
24      mass of urban PM, because of their small size, DPM may represent a significant fraction of the
25      ultrafme particle mode in urban air, especially in cities and countries that rely heavily on diesel-
26      powered vehicles.  It must be noted that the potential contribution of DPM to the rising
27      prevalence in asthma is complicated by the fact that DPM levels have been decreasing over the
28      last decade (CALEPA report). The reported decrease in DPM levels is a result of the increased
29      combustion efficiency of diesel engines. This improvement in diesel  engine design also has
30      brought about a significant decrease in the particle size of diesel emissions.  Thus, the balance


        April 2002                                 7-58       DRAFT-DO NOT QUOTE OR CITE

-------
 1      between a decrease in diesel emissions (in terms of mass) versus the production of a smaller and
 2      potentially more toxic particle size needs further exploration.
 3
 4      7.4.4 Resistance to Infectious Disease
 5           The development of an infectious disease requires both the presence of the appropriate
 6      pathogen, as well as host susceptibility to the pathogen. There are numerous specific and
 7      nonspecific host defenses against microbes, and the ability of inhaled particles to modify
 8      resistance to bacterial infection could result from a decreased ability to clear or kill microbes.
 9      Rodent infectivity models frequently have been used to examine the effect of inhaled particles on
10      host defense and infectivity. Mice or rats are challenged with a bacterial or viral load either
11      before or after exposure to the particles (or gas) of interest; mortality rate, survival time, or
12      bacterial clearance are then examined.  A number of studies that have used the infectivity model
13      to assess the effect of inhaled PM were discussed previously (U.S. Environmental Protection
14      Agency, 1982, 1989, 1996a). In general, acute exposure to sulfuric  acid aerosols at
15      concentrations up to 5,000 //g/m3 were not very effective in enhancing mortality in a bacterially
16      mediated murine model.  In rabbits, however, sulfuric acid aerosols altered anti-microbial
17      defenses after exposure for 2 h/day for 4 days to 750 //g/m3 (Zelikoff et al., 1994). Acute or
18      short-term repeated exposures to high concentrations of relatively inert particles have produced
19      conflicting results.  Carbon black (10,000 //g/m3) was found to have no effect on susceptibility to
20      bacterial infection (Jakab, 1993); whereas a very high concentration (20,000 //g/m3) of TiO2
21      decreased the clearance of microbes and the bacterial response of lymphocytes isolated from
22      mediastinal lymph nodes (Gilmour et al., 1989a,b). In addition, exposure to DPM (2 mg/m3,
23      7h/d, 5d/wk for 3 and 6 mo) has been shown to enhance the susceptibility of mice to the lethal
24      effects of some, but not all, microbial agents (Hahon et al., 1985). Thus, the pulmonary response
25      to microbial agents has been shown to be altered at relatively high particle concentrations in
26      animal models.  Moreover, these effects appear to be highly dependent on the microbial
27      challenge and the test animal studied. Pritchard et al. (1996) observed in rats exposed to particles
28      with a high concentration of metals (e.g., ROFA), that the increased mortality rate after
29      streptococcus infection was associated with the amount of metal in the PM.
30           Despite the reported association between ambient PM and deaths caused by pneumonia
31      (Schwartz, 1994), there are few recent  studies that have examined the mechanisms that may be
        April 2002                                 7-59        DRAFT-DO NOT QUOTE OR CITE

-------
 1      responsible for the effect of PM on infectivity.  In one study, Cohen and colleagues (1997)
 2      examined the effect of inhaled vanadium (V) on immunocompetence. Healthy rats were
 3      repeatedly exposed to 2 mg/m3 V, as ammonium metavanadate, and then instilled with
 4      polyinosinic-polycytidilic acid (poly I:C), a double-stranded polyribonucleotide that acts as a
 5      potent immunomodulator.  Induction of increases in lavage fluid protein and neutrophils was
 6      greater in animals preexposed to V.  Similarly, IL-6 and interferon-gamma were increased in
 7      V-exposed animals. Alveolar macrophage function, as determined by zymosan-stimulated
 8      superoxide anion production and by phagocytosis of latex particles, was depressed to a greater
 9      degree after poly I:C instillation in V-exposed rats as compared to filtered air-exposed rats.
10      These findings provide evidence that inhaled V, a trace metal found in combustion particles and
11      shown to be toxic in vivo in studies using instilled or inhaled ROFA  (Dreher et al., 1997;
12      Kodavanti et al., 1997b, 1999), has the potential to inhibit the pulmonary response to microbial
13      agents.  It must be taken into consideration that these effects were found at very high exposure
14      concentrations of V, and as with many studies, care must be taken in extrapolating the results to
15      the ambient exposure of healthy individuals or those with preexisting cardiopulmonary disease to
16      trace concentrations (approximately 3 orders of magnitude lower concentration) of metals in
17      ambient PM.
18
19
20      7.5  PARTICULATE MATTER TOXICITY AND PATHOPHYSIOLOGY:
21           IN VITRO EXPOSURES
22      7.5.1 Introduction
23           Toxicological studies play an integral role in determining the biological plausibility for the
24      health effects associated with ambient PM exposure.  At the time of completion of the previous
25      PM AQCD (U.S. Environmental Protection Agency,  1996a)  very little was known about the
26      potential mechanisms that could explain the morbidity and mortality observed in populations
27      exposed to PM.  One of the difficulties in trying to sort out possible mechanisms is the nature of
28      particles themselves. Ambient PM has diverse physicochemical properties (Table 7-9) ranging
29      from the physical characteristics of the particle to the chemical components in or on the surface
30      of the particle. Any one of these properties could change at any time in the ambient exposure
31      atmosphere, making it hard to replicate the actual properties in a controlled experiment. As a
        April 2002                               7-60        DRAFT-DO NOT  QUOTE OR CITE

-------
           TABLE 7-9.  PHYSICOCHEMICAL PROPERTIES OF PARTICIPATE MATTER
                 Physical Characteristics
            Chemical Components
          particle mass (size, shape, density)
          particle number
          surface area
          surface chemistry
          surface charge
          acidity
 • elemental and organic carbon
 • volatile organics
 • metals (Fe, Cd, Co, Cu, Mn, Ni, Pb, Ti, V, Zn)
 • biologicals (e.g., pollen, microbes)
 • sulfates
 • nitrates
 * pesticides	
 1      result, controlled exposure studies as yet have not been able to unequivocally determine the
 2      particle properties and the specific mechanisms by which ambient PM may affect biological
 3      systems.
 4           Despite these underlying difficulties, a larger number of toxicological studies have become
 5      available since 1996 to help explain how ambient particles may exert toxic effects on the
 6      cardiovascular and respiratory systems. The following section discusses the more recently
 7      published studies that provide an approach toward identifying potential  mechanisms by which
 8      PM mediates health effects.  The remaining sections discuss potential mechanisms in relation to
 9      PM characteristics based on these available data.
10
11      7.5.2 Experimental Exposure Data
12           In vitro exposure is a useful technique to provide information on potential hazardous PM
13      constituents and mechanisms of PM injury, especially when only limited quantities of the test
14      material are available.  In addition, in vitro exposure allows the  examination of the response to
15      particles in only one or two cell types. Respiratory epithelial cells that line the airway lumen,
16      constitute the initial targets of airborne pollutants.  These cells have been featured in numerous
17      studies involving airborne pollutants and show inflammatory responses  similar to that of human
18      primary epithelial cultures. Limitations of in vitro studies include difficulty in extrapolating
19      dose-response relationships and from in vitro to in vivo biological response and mechanistic
20      extrapolations. In addition to alterations in physiochemcial characteristics of PM because of the
21      collection and resuspension processes, these exposure conditions do not simulate the air-cell
        April 2002
7-61
DRAFT-DO NOT QUOTE OR CITE

-------
 1      interface that actually exists within the lungs, and, thus, the exact dosage delivered to target cells
 2      in vivo is not known. Furthermore, unless an in vitro exposure system that is capable of
 3      delivering particles uniformly to monolayers of airway epithelial cells cultured in an air-liquid
 4      interface system is used (Chen et al.,  1993), the conventional incubation system alters the
 5      microenvironment surrounding the cells and may alter the mechanisms of cellular injury induced
 6      by these agents.
 7           Even with these limitations, in vitro studies do provide an approach to identify potential
 8      cellular and molecular mechanisms by which PM mediates health effects. These mechanisms
 9      can then be evaluated in vivo.  In vitro studies are summarized in Table 7-10.
10
11      7.5.2.1 Ambient Particles
12           Several studies have exposed airway epithelial cells, alveolar macrophages, or blood
13      monocytes and erythrocytes to aqueous extracts of ambient PM to investigate cellular processes
14      such as oxidant generation and cytokine production that may contribute to the pathophysiological
15      response seen in vivo.  Among the ambient PM being examined were samples collected from
16      Boston, MA, (Goldsmith et al., 1998); North Provo, UT (Ohio et al., 1999a,b); St. Louis, MO
17      (SRM 1648, Dong et al., 1996; Becker and Soukup, 1998); Washington, DC (SRM 1649, Becker
18      and Soukup, 1998); Ottawa, Canada (EHC-93, Becker and Soukup, 1998); Dusseldorf and
19      Duisburg, Germany (Hitzfeld et al., 1997), Mexico City (Bonner et al., 1998), Terni, Italy
20      (Fabiani et al., 1997); and Rome, Italy (Diociaiuti et al., 2001).  In any in vitro studies, however,
21      there is a potential for contamination of ambient PM by biologic material during collection on
22      filters.  Endotoxin contamination, in particular, can occur at any time in the manufacture of the
23      filter media or during handling of the filter samples before, during, and after the particle
24      collection process.  This potential inadvertent contamination of filter samples can make
25      extrapolation of the study results difficult,  although careful handling, characterization, and
26      controls can eliminate these concerns.
27           Because soluble metals of ambient surrogates like ROFA have been associated with
28      biological effect and toxicity, several studies have investigated whether the soluble components
29      of ambient PM may have the same biological activities. Extracts of ambient PM samples
30      collected from North Provo, UT, (during 1981 and 1982) were used to test whether the soluble
31      components or ionizable metals, which accounted for approximately 0.1% of the mass, are

        April 2002                                 7-62        DRAFT-DO NOT QUOTE OR CITE

-------
>
to
o
o
to
       TABLE 7-10. IN VITRO EFFECTS OF PARTICIPATE MATTER AND PARTICIPATE MATTER CONSTITUENTS
Species, Cell type,
etc.
Human bronchial
epithelial cells,
asthmatic (ASTH)
nonasthmatic (NONA)



Human bronchial
epithelial cells
(smokers)

Human and
rat AM







Human AM and
blood monocytes






Rat AM







NHBE cells



Particle or Exposure
Constituent Technique
DPM In vitro






DPM In vitro



Four Urban air In vitro exposure,
particles: 2 x 105 cells
ROFA exposed for 2 h
DPM
Volcanic ash
Silica



Urban air In vitro
particles;
St. Louis SRM
1648;
Washington,
DC, SRM 1649;
Ottawa, Canada,
EHC-93
PM10 In vitro
Mexico City
1993; volcanic
ash (MSHA)




ROFA In vitro



Concentration Particle Size Exposure Duration
10-100 //g/mL 0.4 //m 2, 4, 6, 24 h






10-100 Mg/mL 0.4 i/m 24 h



Urban and DPM: Urban particles: 2 h for cytotoxicity,
12,27,111,333, 0.3-0.4 Mm 16-18 h for cytokine
or 1000 //g/niL DPM: 0.3 /an assay;
SiO2 and TiO2: ROFA: 0.5 /an chemiluminescence at
4, 12, 35, or Volcanic ash: 30 minutes
167//g/mL 1.8//m
Fe2O3: 1:1, 3:1; Silica- 05-10 /an
10:1 particles/cell TiO2: <5 /an
ratio Latex: 3.8 /an
33 or 100 Mg/mL 0.2 to 0.7 /an 3, 6, or 18-20 h







1-100 //g/mL 
-------
 H
 6
 o
 o
 H
O
o
HH
H
W
                           TABLE 7-10 (cont'd).  IN VITRO EFFECTS OF PARTICIPATE MATTER AND PARTICIPATE
                                                                          MATTER CONSTITUENTS
to
o
o
to
Species, Cell type,
etc.
Human
erythrocytes;
RAW 264.7 cells
Particle or
Constituent
PM10.2.5; PM25 from
Rome, Italy
Exposure
Technique
In vitro
Concentration
50 ± 45 Mg/m3
31±24,ug/m3
19 ± 20 |/g/m3
Particle Size
PM10
PM25
PM10.2.5
Exposure
Duration
Ih
24 h
Effect of Particles
Oxidative stress on cell membranes is related to PM
surface per volume unit of suspension; small particles
are more effective at decreasing viability and increasing
markers of inflammation.
Reference
Diociaiuti
etal. (2001)
          Supercoiled
          DNA
PM10 from
Edinburgh, Scotland
          Rat AM
                               UAP
                               DPM
In vitro     996.2± 181.8           PM10                8h          PM10 caused damage to DNA; mediated by hydroxyl
           l/g/filter in                                              radicals (inhibited by mannitol) and iron (inhibited by
           100 ,uL                                                 DEF). Clear supernatant has all of the suspension
                                                                  activity.  Free radical activity is derived either from a
                                                                  fraction that is not centrifugeable on a bench centrifuge
                                                                  or that the radical generating system is released into
                                                                  solution.

In vitro     50 to            DPM: 1.1 - 1.3 ^m   2 h exposure;         Dose dependent increase in TNF-a,  IL-6, CINC, MIP-2
           200 ,ug/mL       UAP: St Louis,       supernatant          gene expression by urban particles but not with DPM;
                           between 1974 and    collected 18 h        cytokine production were not related to ROS; cytokine
                           1976 in a baghouse,   postexposure         production can be inhibited by polymyxin B; LPS was
                           sieved through                           detected on UAP but not DPM; endotoxin is
                           200-mesh (125 ,um)                       responsible for the cytokine gene  expression induced by
                                                                  UAP in AM..
Donaldson
etal. (1997)
                                                                                                                                        Dong et al.
                                                                                                                                        (1996)
          Primary cultures
          ofRTE
                                ROFA
                                                   In vitro
                                                 1.95 ,um MMAD      Analysis at 6 and     Particle induced epithelial-cell detachment and lytic
                                                                     24 h                cell injury; alterations in the permeability of the
                                                                                        cultured RTE cell layer; increase in LDH, G-6-PDH,
                                                                                        gluathione reductase, glutathione S-transferase;
                                                                                        mechanism of ROFA-induced RTE cytotoxicity and
                                                                                        pulmonary cellular inflammation involves
                                                                                        the development of an oxidative burden.
Dye et al.
(1997)
Primary cultures
of RTE
Peripheral blood
monocytes
BEAS-2B
ROFA; metal
solutions
Organic extract of
TSP, Italy
Provo PM10 extract
In vitro 5, 10, or
20 ,ug/m2
In vitro 42.5 ^g
extract/m3
(acetone)
In vitro 125, 250,
1 .95 ij,m MMAD Analysis at 6 and
24 h
N/A, collected from 2 h
high-volume
sampler (60 m3/h)
PM10 2 and 24 h
Over 24 h ROFA, V, or Ni + V, but not Fe or Ni,
increased epithelial permeability, decreased cellular
glutathione, cell detachment, and lytic cell injury;
treatment with DMTU inhibited expression of MIP-2
and IL-6 genes.
Superoxide anion generation was inhibited at
a particulate concentration of 0. 17 mg/mL when
stimulated with PMA; 50% increase in LDH;
disintegration of plasma membrane.
Dose-dependent increase in IL-6 and IL-8 produced by
particles collected while the steel mill was in operation;
particles collected during plant closure had the lowest
concentrations of soluble Fe, Cu, And Zn
Dye et al.
(1999)
Fabiani et al.
(1997)
Frampton et al.
(1999)

-------
 H
 6
 o

 o
 H
O
o
HH
H
W
Species, Cell
type, etc.
Rat AM
NHBE
BEAS-2B


BEAS-2B
respiratory
epithelial cells
BEAS-2B
0X174 RF1
DNA
Hamster AM
Hamster AM
AMs from
female CD rats
Exposure
Particle or Constituent Technique Concentration
ROFA, iron sulfate, In vitro 0.01-1.0 mg/mL
nickel sulfate, vanadyl (0.7 x 106
sulfate cells/mL)
Latex particles with
metal complexed on
the surface
ROFA In vitro 5-200 ^g/mL


ROFA In vitro 100 ,ug/mL
Provo In vitro 500 ,ug/mL
TSP soluble and
insoluble extract
PM10 from Edinburgh, In vitro 3.7 or 7.5 i/g/mL
Scotland
ROFA or CAPs In vitro 0, 25, 50, 100, or
200 //g/mL
CAPs, ROFA, and their In vitro 0-200 mg/mL
water-soluble and
particulate fractions
Vanadyl chloride In vitro 10-1000 //m
sodium metavanadate metavanadate
Exposure
Particle Size Duration Effect of Particles
3.6^mMMAD Up to 400 min Increase chemiluminescence, inhibited by DEF and
hydroxyl radical scavengers; solutions of metal
sulfates and metal-complexed latex particles
similarly elevated chemiluminescence in a dose-
and time-dependent manner.
3.6,um 2 and 24 h mRNA for ferritin did not change; ferritin protein
increase; mRNA for transferrin receptor decreased,
mRNA for lactoferrin increased; transferrin
decreased whereas lactoferrin increased;
deferoxamine alone increased lactoferrin mRNA.
N/A ~ 1 h Lactoferrin binding with PM metal occurred
within 5 min. V and Fe (m), but not Ni, increased
the concentration of lactoferrin receptor.
TSP 24 h Water soluble fraction caused greater release of
IL-than insoluble fraction. The effect was
blocked by deferoxamine and presumably
because of metals (Fe, Cu, Zn, Pb).
PM10 8 h Significant free radical activity on degrading
supercoiled DNA; mainly because of hydroxyl
radicals (inhibited by mannitol); Fe involvement
(DEF-B conferred protection); more Fe3+ was
released compared to Fe2+, especially at pH 4.6
than at 7.2.
CAPs: 30 min Dose-dependent increase in AM oxidant stress with
0.1-2.5 //m incubation, both ROFA and CAP. Increase in particle uptake;
(from Harvard analysis Mac-type SR mediate a substantial proportion of
concentrator) immediately AM binding; particle-associated components (e.g.,
TiO2: 1 fjm following transition metals) are likely to mediate intracellular
oxidant stress and proinflammatory activation.
CAPs = 0. 125 i^m 30 min ROFA and CAPs (water soluble components)
ROFA =1.0 fjm caused increases in DCFH oxidation; CAPs
samples and components showed substantial day-
to-day variability in their oxidant effects; ROFA
increased MIP-2 and TNF-a production in AM
and can be inhibitable by NAC.
N/A 30 min Metavanadate caused increased production of
ROS. The LOEL was 50 ^M.
Reference
Ohio et al.
(1997a)
Ohio et al.
(1998c)


Ohio et al.
(1999b)
Ohio et al.
(1999a)
Gilmour et al.
(1996)
Goldsmith
etal. (1997)
Goldsmith
etal. (1998)
Grabowski
etal. (1999)

-------
TABLE 7-10 (cont'd). IN VITRO EFFECTS OF PARTICIPATE MATTER AND PARTICIPATE
                          MATTER CONSTITUENTS
to
o
o
to
















-;J
ON
ON



O
c
•T1
H
6
o
l>
-^H
n
s.**
H


O
w


o
H H
H
W



Species, Cell type,
etc.
Human PMN









Human AM





Primary GPTE
cells




Human Bronchial
Epithelial
(BEAS-2B) cells









Rat AM


Human lung
mucoepidermoid
carcinoma cell line,
NCI-H292
BEAS-2B, airway
epithelial cells
Particle or
Constituent
Aqueous and
organic extracts of
TSP in Dusseldorf
and Duisburg,
Germany





UAP
(#1648, 1649)
Volcanic ash
ROFA


ROFA
DOFA
STL
WDC
OT
MSH
TSP collected in
Provo










ROFA, 10 samples
with differing
metal composition
ROFA



ROFA

Exposure
Technique Concentration
In vitro 0.42-0.78 mg
dust/mL








In vitro 0, 25, 100, or
200 Mg/mL




In vitro 6-25, 12.5, 25,
and 50 Mg/cm3




In vitro TSP filter
samples
(36.5 mg/mL)
agitated in
deionized H2O2
for 96 h,
centrifuged at
1200g for 30 min,
lyophylized and
resuspended in
deionized H2O2 or
saline
In vitro 0 or 50 Mg/mL


In vitro 30 Mg/ml



In vitro 0, 0.5, or 2.0 mg
in 10 mL
Particle Size
Collected by high
volume sampler, 90%
<5 Mm, 50% 
-------
to
o
o
to
 H
 6
 o
 o
 H
O
o
HH
H
W
                         TABLE 7-10 (cont'd). IN VITRO EFFECTS OF PARTICIPATE MATTER AND PARTICIPATE
                                                                     MATTER CONSTITUENTS
Species, Cell type,
etc.
Male (Wistar) rat
lung macrophages
Human blood
monocytes and
neutrophils (PMN)
Human airway
epithelium-derived
cell lines BEAS-2B
(S6-subclone)
Human airway
epithelium-derived
cell line BEAS 2B
Human airway
epithelium-derived
cell line BEAS
Human airway
epithelium-derived
Particle or
Constituent
Urban dust SRM
1649, TiO2,
quartz
Ambient air
particles, carbon
black, oil fly ash,
coal fly ash
ROFA

ROFA
ROFA
Synthetic ROFA
(soluble Ni, Fe,
and V)
Particle
components As,
Exposure
Technique Concentration Particle Size
In vitro 0-100 //g in 1 mL N/A
In vitro 100 //g in N/A
0.2 mL
In vitro 0, 6, 12, 25, or 1.96 fj.m
50 ,ug/mL

In vitro 2, 20, or 60 //g/cm2 1 .96 ^m
In vitro ROFA: 0-200 ROFA: 1. 96 ^m
Mg/mL Synthetic ROFA: N/A
Synthetic ROFA (soluble)
(lOO^g/mL):
Ni, 64 i/M
Fe, 63 ,uM
V, 370 mM
In vitro 500 ^M of As, F, N/A (soluble)
Cr (III), Cu, V, Zn
Exposure
Duration
18h
40 min.

1 and 24 h

24-h
exposure
Up to 24 h
20 min and
6 and 24 h
Effect of Particles
Cytotoxicity ranking was quartz > SRM 1649 > TiO2,
based on cellular ATP decrease and LDH, acid
phosphatase, and p-glucuronidase release.
ROS generation, measured by LCL increased in PMN,
was correlated with Si, Fe, Mn, Ti, and Co content but
not V, Cr, Ni, and Cu. Deferoxamine, a metal ion-
chelator, and did not affect LCL in PMN, suggesting that
metal ions are not related to the induction of LCL.
Activation of IL-6 gene by NF-KB activation and
binding to specific sequences in promoter of IL-6 gene;
inhibition of NF-KB activation by DEF and NAC;
increase in PGE2, IL-6, TNF, and IL-8; activation
NF-B may be a critical first step in the inflammatory
cascade following exposure to ROFA particles.
Epithelial cells exposed to ROFA for 24 h secreted
substantially increased amounts of the PHS products
prostaglandins E2 and F2a; ROFA-induced increase in
prostaglandin synthesis was correlated with a marked
increase in PHS activity.
Tyrosine phosphatase activity, which was known to
be inhibited by vanadium ions, was markedly diminished
after ROFA treatment; ROFA exposure induces
vanadium ion-mediated inhibition of tyrosine
phosphatase activity, leading to accumulation of protein
phosphotyrosines in BEAS cells.
Noncytotoxic concentrations of As, V, and Zn induced
a rapid phosphorylation of MAPK in BEAS cells;
Reference
Nadeau et al.
(1996)
Prahalad et al.
(1999)
Quay etal. (1998)

Samet et al.
(1996)
Samet et al.
(1997)
Samet et al.
(1998)
         cell lines BEAS-2B  Cr, Cu, Fe, Ni, V,
                           and Zn
         A549
         OX174RFIDNA
                           Urban particles:
                           SRM 1648,
                           St. Louis
                           SRM 1649,
                           Washington, DC
In vitro      Img/mLforFe     SRM 1648:
            mobilization assay     50% < 10 ,um
                             SRM 1649:
                               30% < 10//m
Up to 25 h
activity assays confirmed marked activation of ERK,
JNK, and P38 in BEAS cells exposed to As, V, and
Zn. Cr and Cu exposure resulted in a relatively small
activation of MAPK, whereas Fe and Ni did not activate
MAPK under these conditions; the transcription factors
c-Jun and ATF-2, substrates of JNK and P38,
respectively, were markedly phosphorylated in BEAS
cells treated with As, Cr, Cu, V, and Zn; acute exposure
to As, V, or Zn that activated MAPK was sufficient to
induce a subsequent increase in IL-8 protein expression
in BEAS cells.

Single-strand breaks in DNA were induced by PM
only in the presence of ascorbate, and correlated with
amount of Fe that can be mobilized; ferritin in A549
cells was increased with treatment of PM suggesting
mobilization of Fe in the cultured cells.
Smith and Aust
(1997)

-------
          TABLE 7-10 (cont'd). IN VITRO EFFECTS OF PARTICIPATE MATTER AND PARTICIPATE
                                      MATTER CONSTITUENTS
to
o
o
to














^1
ON
oo



o
§>
-n
H
6
O

H
O
c|
o
3
o
^
o
H
W
Species, Cell type,
etc.*
Human AMs


Human AMs



Rat (Wistar) AM
RAM cells
(a rat AM cell line)

A549


A549






RLE-6TN cells
(type II like cell
line)


Rat, Long Evans
epithelial cells

BEAS-2B human
bronchial epithelial
cells
NHBE
BEAS-2B


Tell tunes- RTF = R
Particle or
Constituent*
Provo PM10
extract

Chapel Hill PM
extract; both H20
soluble(s) and
insoluble(is)
TiO2



ROFA, a-quartz,
TiO2

TiO2, Fe2O3,
CAP, and the
fibrogenic
particle a-quartz



PM2 5, Burlington,
VT;'
Fine/ultrafme
TiO2

CFA
PFA
a-quartz.
ROFA
Birmingham, AL.
188mg/gofVO
Provo PM10
extract


attrar.liHal Hnitlinlial r.H
Exposure
Technique Concentration Particle Size
In vitro 500 Mg PM10


In vitro 100 Mg/mL PM25
PM10.2.5


In vitro 20, 50, or N/A
80 Mg/mL


In vitro 1 mg/mL N/A


In vitro TiO2 [40 Mg/mL], N/A
Fe2O3 [100
Mg/mL], a-quartz
[200 Mg/mL], or
CAP [40 Mg/mL]


In vitro 1, 2.5, 5, or PM25: 39 nm
10 Mg/mL Fine TiO2: 159 nm
UF TiO2: 37 nm


2.6 Mm
17.7 Mm
2.5 Mm
In vitro 100 Mg/mL N/A


In vitro 50, 100, PM10
200 Mg/mL


!!«• ("rPTF = ("riiinHa mo trar.liRal Hnitlinlial r.nlls
Exposure
Duration Effect of Particles
24 h AM phagocytosis of (FITC)-labeled Saccharomyces
cerevisiae inhibited 30% by particles collected
before steel mill closure.
24 h Increased cytokine production (IL-6, TNFa,
MCP-1); is PM10 > , PM10 > is PM25; , PM25 was
inactive; endotoxin was partially responsible.

4 h Opsonization of TiO2 with surfactant components
resulted in a modest increase in AM uptake
compared with that of unopsonized TiO2; surfactant
components increase AM phagocytosis of particles.
60 min Exposure of A549 cells to ROFA, a-quartz, but not
TiO2, caused increased IL-8 production in TNF-a
primed cells in a concentration-dependent manner.
24 h TiO2 > Fe2O3 > a-quartz > CAP in particle binding;
binding of particle was found to be calcium-
dependent for TiO2 and Fe2O3, while a-quartz
binding was calcium-independent; scavenger
receptor, mediate particulate binding; a-quartz, but
not TiO2 or CAP, caused a dose-dependent
production of IL-8.
24 and 48 h Increases in c-Jun kinase activity, levels of
exposure phosphorylated c-Jun immunoreactive protein, and
transcriptional activation of activator protein-
1 -dependent gene expression; elevation in number
of cells incorporating 5 '-bromodeoxyuridine.
3 h CFA produced highest level of hydroxyl radicals;
iron content is more important than quartz content.

2-6 h ROFA caused increased intracellular Ca++, IL-6,
IL-and TNF-a through activation of capsicin-
and pH-sensitive receptors.
24 h Dose-dependent increase in expression of IL-8
produced by particles collected when the steel mill
was in operation.



Reference
Soukup et al.
(2000)

Soukup and
Becker (2001)


Stringer and
Kobzik
(1996)

Stringer and
Kobzik (1998)

Stringer et al.
(1996)





Timblin et al.
(1998)



Van Maanen
etal. (1999)

Veronesi et al.
(1999b)

Wu etal. (2001)




"•Particles: See Table 7-1

-------
 1      responsible for the biological activity of the extracted PM components. The oxidant generation
 2      (thiobarbituric acid reactive products), release of IL-8 from BEAS-2B cells, and PMN influx in
 3      rats exposed to these samples correlated with sulfate content and the ionizable concentrations of
 4      metals in these PM extracts (Ohio et al., 1999a,b). In addition, these  extracts stimulated IL-6 and
 5      IL-8 production as well as increased IL-8 mRNA and enhanced expression of intercellular
 6      adhesion molecule-1 (ICAM-1) in BEAS-2B cells (Kennedy et al., 1998).  Cytokine secretion
 7      was preceded by activation of nuclear factor kappa B (NF-KB) and was reduced by treatment
 8      with superoxide dismutase (SOD), Deferoxamine (DBF), or N-acetylcysteine. The addition of
 9      similar quantities of Cu+2 as found in the Provo extract replicated the  biological effects observed
10      with particles alone. When normal constituents of airway lining fluid (mucin or ceruloplasmin)
11      were added to BEAS cells, particulate-induced secretion of IL-8 was  modified. Mucin reduced
12      IL-8 secretion; whereas ceruloplasmin significantly increased IL-8 secretion and activation of
13      NF-KB.  The authors suggest that copper ions may cause some of the  biologic effects of inhaled
14      PM in the Provo region and may provide an explanation for the sensitivity of asthmatics to Provo
15      PM seen in epidemiologic studies.
16          Frampton et al. (1999) examined the effects of the same ambient PM samples collected
17      from Utah Valley in the late 1980s (see Section 7.2.1). Aqueous extracts of the filters were
18      analyzed for metal and oxidant production and added to cultures of human respiratory epithelial
19      cells (BEAS-2B) for 2 or 24 h. Particles collected in 1987, when the  steel mill was closed had
20      the lowest concentrations of soluble iron, copper, and zinc and showed the least oxidant
21      generation. Ambient PM collected before and after plant closing induced expression of IL-6 and
22      IL-8 in a dose-response relationship (125, 250, and 500 //g/mL). Ambient PM collected after
23      reopening of the steel mill also caused cytotoxicity, as demonstrated by microscopy and LDH
24      release at the highest concentration used (500 //g/mL).
25           Soukup et al.  (2000) used similar ambient PM extracts as Frampton et al. (1999) to
26      examine effects on human alveolar macrophages.  The phagocytic activity and oxidative response
27      of AMs  was measured after segmental instillation of aqueous extracts from the Utah Valley or
28      after overnight in vitro cell culture.  Ambient PM collected before closure of the steel mill
29      inhibited AM phagocytosis of (FITC)-labeled Saccharomyces cerevisiae by 30%; no significant
30      effect on phagocytosis was seen with the other two extracts. Furthermore, although extracts of
31      ambient PM collected before and after plant closure inhibited oxidant activity of AMs when

        April 2002                                 7-69        DRAFT-DO NOT QUOTE OR CITE

-------
 1      incubated overnight in cell culture, only the former particles caused an immediate oxidative
 2      response in AMs.  Host defense effects were attributed to apoptosis which was most evident in
 3      particles collected before plant closure. Interpretation of loss of these effects by chelation
 4      removal of the metals was complicated by the observed differences in apoptosis despite similar
 5      metal contents of ambient PM collected during the steel mill operation.
 6           Wu et al (2001) investigated the intracellular signaling mechanisms for the pulmonary
 7      responses to Utah Valley PM extracts.  Human primary airway epithelial cells were exposed to
 8      aqueous extracts of PM collected from the year before, during, and after the steel mill closure in
 9      Utah Valley.  Transfection with kinase-deficient extracellular signal-regulated kinase (ERK)
10      constructs partially blocked the PM-induced interleukin (IL)-8 promoter reporter activity. The
11      mitogen-activated protein kinase/ERK kinase (MEK) activity inhibitor PD-98059 significantly
12      abolished IL-8 released in response to the PM, as did the epidermal growth factor (EOF) receptor
13      kinase inhibitor AG-1478. Western blotting showed that the PM-induced phosphorylation of
14      EOF receptor tyrosine, MEK1/2, and ERK1/2 could be ablated with AG-1478 or PD-98059.  The
15      results indicate that the potency of Utah Valley PM collected during plant closure was lower than
16      that collected while the steel mill was in operation and imply that Utah Valley PM can induce IL-
17      8 expression partially through the activation of the EGF receptor signaling.
18           There are regional as well as daily variations in the composition of ambient PM and, hence,
19      its biological activities. For example, concentrated ambient PM (CAP, from Boston urban air)
20      has substantial day-to-day variability in its composition and oxidant effects (Goldsmith et al.,
21      1998).  Similar to Utah PM, the water-soluble component of Boston CAPs significantly
22      increased AM oxidant production and inflammatory cytokine (MIP2 and TNFa) production over
23      negative control values. These effects can be blocked by metal  chelators or antioxidants. The
24      regional difference in biological  activity of ambient PM has been shown by Becker and Soukup
25      (1998). The  oxidant generation, phagocytosis, as well as the expressions of receptors important
26      for phagocytosis in human alveolar macrophage and blood monocyte  were reduced significantly
27      by PM exposure.
28           Becker and Soukup (1998) and others (Dong et al., 1996, Becker et al., 1996) have
29      suggested that the biological activity of the ambient PM may result from the presence of
30      endotoxin on the particles rather than metal-associated oxidant generation. Using the same urban
31      particles (SRM 1648), cytokine production (TNF-a, IL-1,11-6, CINC, and MIP-2) was increased

        April 2002                                7-70        DRAFT-DO NOT QUOTE OR CITE

-------
 1      in macrophages following treatment with 50 to 200 //g/mL of urban PM (Dong et al., 1996).  The
 2      urban particle-induced TNF-a secretion was abrogated completely by treatment with polymyxin
 3      B, an antibiotic that blocks LPS-associated activities, but not with antioxidants.
 4           The involvement of endotoxin, at least partially, in PM induced biological effects was
 5      supported more recently by Bonner et al. (1998) and Soukup and Becker (2001).  Urban PM10
 6      collected from north, south, and central regions of Mexico City was used with SD rat AM to
 7      examine PM effects on platelet-derived growth factor (PDGF) receptors on lung myofibroblasts
 8      (Bonner et al., 1998).  Mexico City PM10 (but not volcanic ash) stimulated secretion of
 9      upregulatory factors for the PDGF a receptor, possibly via IL-1 p. In the presence of an
10      endotoxin-neutralizing protein, the Mexico City PM10 effect on PDGF was blocked partially,
11      suggesting that LPS was responsible partially for the effect of the PM10 on macrophages.
12      In addition, both LPS  and vanadium (both present in the PM10) acted directly on lung
13      myofibroblasts. However, the V levels in Mexico City PM10 were probably not high enough to
14      exert an independent effect.  The authors concluded that PM10 exposure could lead to airway
15      remodeling by enhancing myofibroblast replication and chemotaxis.
16           Soukup and Becker (2001) collected fresh PM25 and PM10_2 5 from the ambient air of
17      Chapel Hill, NC, and compared the activity of these two particle size fractions. Both water
18      soluble and insoluble components were assessed for cytokine production, inhibition of
19      phagocytosis, and induction of apoptosis. The most potent fraction was the insoluble PM10_2 5.
20      Endotoxin was responsible for much of the cytokine production, while inhibition of phagocytosis
21      was induced by other moieties in the coarse material.  None of the activities were inhibited by the
22      metal chelator deferoxamine.
23           The effects of water soluble as well as organic components (extracted in dichloromethane)
24      of ambient PM were investigated by exposing human PMN to PM extracts (Hitzfeld et al., 1997).
25      PM was collected with high-volume samplers in two German cities, Dusseldorf and Duisburg;
26      these sites have high traffic and high industrial emissions, respectively.  Organic, but not
27      aqueous, extracts of PM alone significantly stimulated the production and release of ROS in
28      resting human PMN.  The effects of the PM extracts were inhibited by SOD, catalase, and
29      sodium azide (NaN3).  Similarly, the organic fraction (extractable by acetone) of ambient PM
30      from Terni, Italy, had been shown to produce cytotoxicity, superoxide release in response to
31      PMA and zymosan in peripheral monocytes (Fabiani et al., 1997).

        April 2002                                7-71        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Diociaiuti et al. (2001) compared the in vitro toxicity of coarse (PM10_25) and fine (PM25)
 2      particulate matter, collected in an urban area of Rome. The in vitro toxicity assays used included
 3      human red blood cell hemolysis, cell viability, and nitric oxide (NO) release in the RAW 264.7
 4      macrophage cell line.  There was a dose-dependent hemolysis in human erythrocytes when they
 5      were incubated with fine and coarse particles.  The hemolytic potential was greater for the fine
 6      particles than for the coarse particles in equal mass concentration.  However, when data were
 7      expressed in terms of PM surface area per volume of suspension, the hemolytic activity of the
 8      fine fraction was equal to the coarse fraction. This result suggested that the oxidative stress
 9      induced by PM on the cell membranes could be due mainly to the interaction between the particle
10      surfaces and the  cell membranes. Although RAW 264.7 cells challenged with fine and coarse
11      particles showed decreased viability and an increased release of NO, a key inflammatory
12      mediator, both effects were not dose-dependent in the tested concentration range. The fine
13      particles were the most effective in inducing these effects when the data were expressed as mass
14      concentration or as surface area per unit volume.  The authors concluded that these differences in
15      biological activity were due to the different physicochemical  natures of the particles.
16
17      7.5.2.2 Comparison of Ambient and Combustion-Related Surrogate Particles
18           In vitro toxicology studies utilizing alveolar macrophages as target cells (Imrich et al.,
19      2000; Long et al., 2001; Ning et al., 2000; Mukae et al., 2000, 2001; van Eeden et al., 2001) have
20      found that urban air particles are much more potent for inducing cellular responses than
21      individual  surrogate combustion particles such as diesel and ROFA. Similar to the results
22      described above in Section 7.5.2.1, these studies also show that when cytokine responses are
23      measured, LPS/endotoxin is  found to be responsible for most of the activity.  Metals, on the  other
24      hand, do not seem to affect cytokine production, as confirmed by studies showing that ROFA
25      does not induce macrophage cytokine production. These results are important because LPS is an
26      important component associated with both coarse and fine mode particles (Menetrez et al., 2001).
27      In fact, in one study (Long et al., 2001),  cytokine responses in the alveolar macrophages were
28      correlated with LPS content  and more LPS was found associated with indoor PM2 5 than outdoor
29      PM25.
30           Imrich et al., (2000) found that when mice alveolar macrophages were stimulated with
31      CAPs (PM2 5), the resulting TNF responses could be inhibited by the use of an endotoxin

        April 2002                                7-72        DRAFT-DO NOT QUOTE OR CITE

-------
 1      neutralizing agent [e.g., polymyxin-B (PB)].  Because the MIP-2 response (IL-8) was only partly
 2      inhibited by PB; however, the authors concluded that endotoxin primed cells to respond to other
 3      particle components. In a related study (Ning et al., 2000), the use of PB showed that particle-
 4      absorbed endotoxin in CAPs suspensions caused activation of normal (control) AMs, while other
 5      (nonendotoxin) components were predominantly responsible for the enhanced cytokine release
 6      observed by primed AMs incubated with CAPs. The non-LPS component was not identified in
 7      this study, however, the AM biological response did not correlate with any of a panel of elements
 8      quantified within the insoluble CAPs samples (e.g., Al, Cd, Cr, Cu, Fe, Mg, Mn, Ni, S, Ti, V).
 9           Van Eeden et al. (2001) compared ROFA, the atmospheric dust sample EHC-93, and
10      different size latex particles for cytokine induction on human alveolar macrophages. The
11      EHC-93 particles produced greater than 8-fold induction of various cytokines, including IL-1,
12      TNF, GMCSF; the other particles induced these cytokines approximately 2-fold. Using the same
13      EHC-93 particles, Mukae et al. (2000, 2001) found that inhalation exposure stimulated bone
14      marrow band cell-granulocyte precursor production.  They also found that the magnitude of the
15      response was correlated with the amount of phagocytosis of the particles by alveolar
16      macrophages. These results may indicate that macrophages produce factors which stimulate
17      bone marrow, including IL-6 and GMCSF. In fact, alveolar macrophages exposed in vitro to
18      these particles released cytokines; and when the supernatant of PM-stimulated macrophages  was
19      instilled into rabbits, the bone marrow was stimulated.
20           In a series of studies using the same ROFA samples, several in vitro experiments have
21      investigated the biochemical and molecular mechanisms involved in ROFA induced cellular
22      injury. Prostaglandin metabolism in cultured human airway epithelial cells (BEAS-2B and
23      NHBE) exposed to ROFA was investigated by Samet et al. (1996).  Epithelial cells exposed  to
24      ROFA for 24 h secreted substantially increased amounts of prostaglandins E2 and F2 a. The
25      ROFA-induced increase in prostaglandin synthesis was correlated with a marked increase in
26      activity of the PHS-2 form of prostaglandin H synthase as well as mRNA coded for this enzyme.
27      In contrast, expression of the PHS1 form of the enzyme was not affected by ROFA treatment of
28      airway epithelial cells. These investigators further demonstrated that noncytotoxic levels of
29      ROFA induced a significant dose- and time-dependent increase in protein tyrosine phosphate, an
30      important index of signal transduction activation leading to a broad spectrum of cellular
31      responses.  ROFA-induced increases in protein phosphotyrosines were associated with its soluble

        April 2002                                7-73       DRAFT-DO NOT  QUOTE OR CITE

-------
 1      fraction and were mimicked by V-containing solutions but not iron or nickel solutions (Samet
 2      etal., 1997).
 3          ROFA also stimulates respiratory cells to secrete inflammatory cytokines such as IL-6,
 4      IL-8, and TNF. Normal human bronchial epithelial (NHBE) cells exposed to ROFA produced
 5      significant amounts of IL-8, IL-6, and TNF, as well as mRNAs coding for these cytokines (Carter
 6      etal., 1997). Increases in cytokine production were dose-dependent.  The cytokine production
 7      was inhibited by the addition of metal chelator, DBF, or the free radical scavenger
 8      dimethylthiourea (DMTU).  Similar to the data of Samet et al. (1997), V but not Fe or Ni
 9      compounds  were responsible for these effects.  Cytotoxicity, decreased cellular glutathione levels
10      in primary cultures of rat tracheal epithelial  (RTE) cells exposed to suspensions of ROFA
11      indicated that respiratory cells exposed to ROFA were under oxidative stress. Treatment with
12      buthionine sulfoxamine (an inhibitor of y-glutamyl cysteine synthetase) augmented ROFA-
13      induced cytotoxicity; whereas treatment with DMTU inhibited ROFA-induced cytoxicity further
14      suggested that ROFA-induced cell injury may be mediated by hydroxyl-radical-like reactive
15      oxygen species (ROS) (Dye et al., 1997).  Using BEAS-2B cells, a time- and dose-dependent
16      increase in IL-6 mRNA induced by ROFA was shown to be preceded by the activation of nuclear
17      proteins, for example, nuclear factor-KB (NF-icB) (Quay et al., 1998). Taken together, ROFA
18      exposure increases oxidative stress, perturbs protein tyrosine phosphate homeostasis, activates
19      NF-KB, and up-regulates inflammatory cytokine and prostaglandin synthesis and secretion to
20      produce lung injury.
21          Stringer and Kobzik (1998) observed that "primed" lung epithelial cells exhibited enhanced
22      cytokine responses to PM.  Compared to normal cells, exposure of tumor necrosis factor (TNF)-
23      a-primed A549 cells to ROFA or a -quartz  caused increased IL-8 production in a concentration-
24      dependent manner for particle concentrations ranging from 0-200 //g/mL. Addition of the
25      antioxidant N-acetylcysteine (NAC) (1.0 mM) decreased ROFA and a -quartz-mediated IL-8
26      production by approximately 50% in both normal and TNF-a-primed A549 cells. Exposure  of
27      A549 cells to ROFA caused an increase in oxidant levels that could be inhibited by NAC. These
28      data suggest that (1) lung epithelial cells primed by inflammatory mediators show increased
29      cytokine production after exposure to PM and (2) oxidant stress is an important mechanism for
30      this response.


        April 2002                               7-74       DRAFT-DO NOT QUOTE OR CITE

-------
 1           In summary, exposure of lung epithelial cells to ambient PM or ROFA leads to increased
 2      production of cytokines and the effects may be mediated, at least in part, through production of
 3      ROS.  Day-to-day variations in the components of PM, such as soluble transition metals, which
 4      may be critical to eliciting the response, are suggested. The involvement of organic components
 5      in ambient PM also was suggested in some studies.
 6
 7      7.5.3 Potential Cellular and Molecular Mechanisms
 8      7.5.3.1 Reactive Oxygen Species
 9           Ambient particulate matter contains transition metals, such as iron (most abundant),
10      copper, nickel, vanadium, and cobalt. These metals are capable of catalyzing the one-electron
11      reductions of molecular oxygen necessary to generate reactive oxygen species (ROS).  These
12      reactions can be demonstrated by the iron-catalyzed Haber-Weiss reactions that follow.
13
14                          Reductant11 + Fe(III) -> Reductantn+1 + Fe(II)                     (1)
15                                   Fe(II) + O2 -» Fe(III) + O2                              (2)
16                                  HO2+O2+H+^ O2+H2O2                             (3)
17                      Fe(II) + H2O2 -»  Fe(III)+*OH + HO"(FentonReaction)                 (4)
18
19      Iron will continue to participate in the redox cycle in the above reactions as long as there is
20      sufficient O2 or H2O2 and reductants.
21           Soluble metals from inhaled PM dissolved into the fluid lining of the airway lumen can
22      react directly with biological molecules (acting as a reductant in the above reactions) to produce
23      ROS.  For example, ascorbic acid in the human lung epithelial lining fluid can react with Fe(in)
24      from inhaled PM to cause single strand breaks in supercoiled plasmid DNA, cj)Xl74 RFI (Smith
25      and Aust, 1997).  The DNA damage caused by a PM10 suspension can be inhibited by mannitol,
26      an hydroxyl radical scavenger, further confirming the involvement of free radicals in these
27      reactions (Gilmour et al., 1996; Donaldson et al., 1997; Li et al., 1997). Because the clear
28      supernatant of the centrifuged PM10 suspension contained  all of the suspension activity, the free
29      radical activity is derived either from a fraction that is not centrifugable (10 min at 13,000 rpm

        April 2002                                7-75        DRAFT-DO NOT QUOTE OR CITE

-------
 1      on a bench centrifuge) or the radical generating system is released into solution (Gilmour et al.,
 2      1996; Donaldson et al., 1997; Li et al., 1997).
 3          In addition to measuring the interactions of ROS and biomolecules directly, the role of
 4      ROS in PM-induced lung injury also can be assessed by measuring the electron spin resonance
 5      (ESR) spectrum of radical adducts or fluorescent intensity of dichlorofluorescin (DCFH), an
 6      intracellular dye that fluoresces on oxidation by ROS. Alternatively, ROS can be inhibited using
 7      free radical scavengers, such as dimethylthiourea (DMTU); antioxidants, such as glutathione or
 8      N-acetylcysteine (NAC); or antioxidant enzymes, such as superoxide dismutase (SOD).  The
 9      diminished response to PM after treatment with these antioxidants may indicate the involvement
10      of ROS; however,  some antioxidants (e.g., thiol-containing) can interact with metal ions directly.
11          As described earlier, Kadiiska et al. (1997) used the ESR spectra of 4-POBN [a-(4-pyridyl
12      l-oxide)-N-tert-butylnitrone] adducts to measure ROS in rats instilled with ROFA and
13      demonstrated the association between ROS production within the lung and soluble metals in
14      ROFA. Using DMTU to inhibit ROS production, Dye et al. (1997) had shown that systemic
15      administration of DMTU impeded development of the cellular inflammatory response to ROFA,
16      but did not ameliorate biochemical alterations in BAL fluid. Goldsmith et al. (1998), as
17      described earlier, showed that ROFA and CAPs caused increases in ROS production in AMs.
18      The water-soluble  component of both CAPs and ROFA significantly increased AM oxidant
19      production over negative control values. In addition, increased PM-induced cytokine production
20      was inhibited by NAC.  Li et al. (1996, 1997) instilled rats with PM10 particles (collected  on
21      filters from an Edinburgh, Scotland, monitoring station). Six hours after intratracheal instillation
22      of PM10, they observed a decrease in glutathione (GSH) levels in the BAL fluid. Although this
23      study does not describe the composition of the PM10, the authors suggest that changes in GSH, an
24      important lung antioxidant, support the contention that the free radical activity of PM10 is
25      responsible for its biological activity in vivo.
26          In addition to ROS generated directly by PM, resident or newly recruited AMs or PMNs
27      also are capable of producing these reactive species on stimulation.  The ROS produced during
28      the oxidative burst can be measured using a chemiluminescence (CL) assay. With this assay,
29      AM CL signals in vitro have been shown to be greatest with ROFA containing primarily soluble
30      V and were less with ROFA containing Ni plus V (Kodavanti  et al., 1998a). As described
31      earlier, exposures to Dusseldorf and Duisburg PM increased the resting ROS production in

        April 2002                                7-76        DRAFT-DO NOT QUOTE OR CITE

-------
 1      PMNs, which could be inhibited by SOD, catalase, and sodium azide (Hitzfeld et al., 1997).
 2      Stringer and Kobzik (1998) showed that addition of NAC (1.0 mM) decreased ROFA-mediated
 3      IL-8 production by approximately 50% in normal and TNF-a-primed A549 cells. In addition,
 4      exposures of A549 cells to ROFA caused a substantial (and NAC inhibitable) increase in oxidant
 5      levels as measured by DCFH oxidation. In human AMs, Becker et al. (1996) found a CL
 6      response for ROFA, but not urban air particles (Ottawa and Dusseldorf) or volcanic ash.
 7           Metal compounds of PM are the most probable species capable of catalyzing ROS
 8      generation on exposure to PM.  To determine elemental content and solubility in relation to their
 9      ability to generate ROS, PMN or monocytes were exposed to a wide range of ambient air
10      particles from divergent sources (one natural dust, two types of oil fly ash, two types of coal fly
11      ash, five different ambient air samples, and one carbon black sample) (Prahalad et al., 1999), and
12      CL production was measured over a 20-min period postexposure.  Percent of sample mass
13      accounted for by XRF detectable elements was 1.2% (carbon black); 22 to 29% (natural dust and
14      ambient air particles); 13 to 22% (oil fly ash particles); and 28 to 49% (coal fly ash particles).
15      The maj or proportion of elements in most of these particles were aluminosilicates and insoluble
16      iron, except oil derived fly ash particles in which soluble vanadium and nickel were in highest
17      concentration, consistent with particle acidity as measured in the supernatants. All particles
18      induced CL response in cells, except carbon black. The CL response of PMNs in general
19      increased with all washed particles, with oil fly ash and one urban air particle showing statistical
20      differences between deionized water washed and unwashed particles. These CL activities were
21      significantly correlated with the insoluble Si, Fe, Mn, Ti, and Co content of the particles.
22      No relationship was found between CL and soluble transition metals such as V, Cr, Ni, and Cu.
23      Pretreatment of the particles with a metal ion chelator, deferoxamine, did not affect CL activities.
24      Particle sulfate content and acidity of the particle suspension did not correlate with CL activity.
25           Soluble metals can be mobilized into the epithelial cells or AMs to produce ROS
26      intracellularly. Size fractionated coal fly ash particles (2.5, 2.5 to 10, and <10 //m) of bituminous
27      b (Utah coal), c (Illinois coal), and lignite (Dakota coal) were used to compare the amount of iron
28      mobilization in A549 cells and by citrate (1 mM) in cell-free suspensions (Smith et al.,  1998).
29      Iron was mobilized by citrate from all three size fractions of all three coal types.  More iron, in
30      Fe(ni) form, was mobilized by citrate from the <2.5-//m fraction than from the >2.5-//m
31      fractions. In addition, the amount of iron mobilized was dependent on the type of coal used to

        April 2002                                 7-77         DRAFT-DO NOT QUOTE OR CITE

-------
 1      generate the fly ash (Utah coal > Illinois coal = Dakota coal) but not related to the total amount
 2      of iron present in the particles. Ferritin (an iron storage protein) levels in A549 cells increased by
 3      as much as 12-fold in cells treated with coal fly ash (Utah coal > Illinois coal > Dakota coal).
 4      More ferritin was induced in cells treated with the <2.5-//m fraction than with the >2.5-//m
 5      fractions. Mossbauer spectroscopy of a fly ash sample showed that the bioavailable iron was
 6      assocated with the glassy aluminosilicate fraction of the particles (Ball et al., 2000).  As with the
 7      bioavailability of iron, there was an inverse correlation between the production of IL-8 and fly
 8      ash particle size with the Utah coal fly ash being the most potent.
 9          Using ROFA and colloidal iron oxide, Ohio et al. (1997b; 1998a,b,c; 1999c; 2000b) have
10      shown that exposures to these particles disrupted iron homeostasis and induced the production of
11      ROS in vivo and in vitro. Treatment of animals or cells with metal-chelating agents  such as DEF
12      with an associated decrease in response has been used to infer the involvement of metal in PM-
13      induced lung injury. Metal chelation by DEF (1 mM) caused significant inhibition of particulate-
14      induced AM oxidant production, as measured using DCFH (Goldsmith et al., 1998).  DEF
15      treatment also reduced NF-KB activation and cytokine secretion in a human bronchial epithelial
16      cell  line (BEAS-2B cells) exposed to Provo PM (Kennedy et al., 1998).  However, treatment of
17      ROFA suspension with DEF was not effective in blocking teachable metal induced acute lung
18      injury  (Dreher et al., 1997). Dreher et al. (1997) indicated that DEF could chelate Fe(HI) and
19      V(n), but not Ni(n), suggesting  that metal interactions played a significant role in ROFA-induced
20      lung injury.
21          Other than Fe, several V compounds have been shown to increase mRNA levels for
22      selected cytokines in BAL cells  and also to induce pulmonary inflammation (Pierce et al., 1996).
23      NaVO3 and VOSO4, highly soluble forms of V, tended to induce pulmonary inflammation and
24      inflammatory cytokine mRNA expression more rapidly and more intensely than the less soluble
25      form, V2O5, in rats.  Neutrophil  influx was greatest following exposure to VOSO4 and lowest
26      following exposure to V2O5.  However, metal components of fly ash have not been shown to
27      consistently increase ROS production from bovine AM treated with combustion particles
28      (Schluter et al., 1995). For example, As(ni), Ni(n), and Ce(in), which are major components of
29      fly ash, had been shown to inhibit the secretion of superoxide anions (O2") and hydrogen
30      peroxide. In the same study, O2" were lowered by Mn(II) and Fe(n); whereas V(IV) increased O2"
31      and  H2O2.  In contrast, Fe(in) increase O2" productions, demonstrating that the oxidation state of

        April 2002                                7-78        DRAFT-DO NOT QUOTE OR CITE

-------
 1      metal may influence its oxidant generating properties. Other components of fly ash, such as
 2      Cd(II), Cr(HI), and V(V), had no effects on ROS.
 3           It is likely that a combination of several metals rather than a single metal in PM is
 4      responsible for the PM induced cellular response. For example, V and Ni+V but not Fe or Ni
 5      alone (in saline with the final pH at 3.0) resulted in increased epithelial permeability, decreased
 6      cellular glutathione, cell detachment, and lytic cell injury in rat tracheal epithelial cells exposed
 7      to soluble salts of these metals at equivalent concentrations found in ROFA (Dye et al., 1999).
 8      Treatment of V-exposed cells with buthionine sulfoximine further increased cytotoxicity.
 9      Conversely, treatment with radical scavenger dimethyl thiourea inhibited the effects in a dose-
10      dependent manner.  These results  suggest that soluble metal or combinations of several metals in
11      ROFA may be responsible for these effects.
12           Similar to combustion particles such as ROFA, the biological response to exposure to
13      ambient PM also may be influenced by the metal content of the particles. Human subjects were
14      instilled with 500 //g (in 20 mL sterile saline) of Utah Valley dust (UVD1, 2, 3, collected during
15      3 successive years)  on the left segmental bronchus and on the right side with sterile saline as
16      control. Twenty-four-hours post-instillation, a second bronchoscopy was performed and
17      phagocytic cells were obtained on both sides of the segmental bronchus. AM from subjects
18      instilled with UVD, obtained by bronchoaveolar lavage 24 h post-instillation, were incubated
19      with fluoresceinated yeast (Saccharomyces cerevisiae) to assess their phagocytic ability.
20      Although the same proportion of AMs were  exposed to UVD phagocytized yeast, AMs exposed
21      to UVD1, which were collected while a local steel mill was open, took up significantly less
22      particles than AMs exposed to other extracts (UVD2 when the steel mill was closed and UVD3
23      when the plant reopened). AMs exposed to UVD1 also exhibited a small decrease in oxidant
24      activity (using dihydrorhodamine-123, DHR). AMs from healthy volunteers were incubated in
25      vitro with the various UVD extracts to assess whether similar effects on human AMs function
26      could be observed to those seen following in vivo exposure.  The percentage of AMs that
27      engulfed yeast particles was significantly decreased by exposure to UVD1 at 100 //g/mL, but not
28      at 25 //g/mL. However, the amount of particles engulfed was the same following exposure to all
29      three UVD extracts. AMs also demonstrated increased oxidant stress (using chemiluminescence)
30      after in vitro exposure to UVD1, and this effect was not abolished with pretreatment of the
31      extract with the metal chelator deferoxamine.  As with the AMs exposed to UVD in vivo, AM

        April 2002                                 7-79        DRAFT-DO NOT QUOTE OR CITE

-------
 1      exposed to UVD in vitro had a decreased oxidant activity (DHR assay). UVD1 contains 61 times
 2      and 2 times the amount of Zn compared to UVD 2 and UVD3, respectively; whereas UVD3
 3      contained 5 times more Fe than UVD1. Ni and V were present only in trace amounts.  Using
 4      similarly extracted samples, Frampton et al. (1999) exposed BEAS-2B cells for 2 and 24 h.
 5      Similar results were observed for oxidant generation in these cells (i.e., UVD 2, which contains
 6      the lowest concentrations of soluble iron, copper, and zinc, produced the least response). Only
 7      UVD 3 produced cytotoxicity at a dose of 500 //g/mL.  UVD 1 and 3, but not 2, induced
 8      expression of IL-6 and 8 in a dose-dependent fashion.  Taken together, these data showed that the
 9      biological response to ambient particle extracts is heavily dependent on the source and; hence,
10      the chemical composition of PM.
11
12      7.5.3.2  Intracellular Signaling Mechanisms
13           In has been shown that the intracellular redox state of the cell modulates the activity of
14      several transcription factors, including NF-KB, a critical step in the induction of a variety of
15      proinflammatory cytokine and adhesion-molecule genes.  NF-KB is a heterodimeric protein
16      complex that in most cells resides in an inactive state in the cell cytoplasm by binding to
17      inhibitory kappa B alpha (IicBa). On appropriate stimulation by cytokines or ROS, IicBa is
18      phosphorylated and subsequently degraded by proteolysis. The dissociation of IicBa from NF-KB
19      allows the latter to translocate into the nucleus and bind to appropriate sites in the DNA to
20      initiate transcription of various genes. Two studies in vitro have shown the involvement of
21      NF-KB in particulate-induced cytokine and intercellular adhesion molecule-1 (ICAM-1)
22      production in human airway epithelial cells (BEAS-2B) (Quay et al., 1998; Kennedy et al.,
23      1998).  Cytokine secretion was preceded by activation of NF-KB and was reduced by treatment
24      with antioxidants or metal chelators.  These results suggest that metal-induced oxidative stress
25      may play a significant role in the initiation phase of the inflammatory cascade following
26      particulate exposure.
27           A second well-characterized human transcription factor, AP-1, also responds to the
28      intracellular ROS concentration.  AP-1 exists in two forms, either in a homodimer of c-jun
29      protein or a heterodimer consisting of c-jun and c-fos.  Small amounts of AP-1 already exist in
30      the cytoplasm in an inactive form, mainly as phosphorylated c-jun homodimer.  Many different
31      oxidative stress-inducing stimuli, such as UV light and IL-1,  can activate AP-1. Exposure of rat

        April 2002                                 7-80        DRAFT-DO NOT QUOTE OR CITE

-------
 1      lung epithelial cells to ambient PM in vitro resulted in increases in c-jun kinase activity, levels of
 2      phosphorylated c-jun immunoreactive protein, and transcriptional activation of AP-1-dependent
 3      gene expression (Timblin et al., 1998). This study demonstrated that interaction of ambient
 4      particles with lung epithelial cells initiates a cell signaling cascade related to aberrant cell
 5      proliferation.
 6           Early response gene transactivation has been linked to the development of apoptosis, a
 7      unique type of programmed cell injury and a potential mechanism to account for PM-induced
 8      changes in cellular response. Apoptosis of human AMs exposed to ROFA (25 //g/mL) or urban
 9      PM was observed by Holian et al. (1998).  In addition, both ROFA and urban PM upregulated the
10      expression of the RFD1+ AM phenotype; whereas only ROFA decreased the RFDl+7+ phenotype.
11      It has been suggested that an increase in the AM phenotype ratio of RFDl+/RFDl+7+ may be
12      related to disease progression in patients with inflammatory diseases.  These data showed that
13      ROFA and urban PM can induce apoptosis of human AMs and increase the ratio of AM
14      phenotypes toward a higher immune active state and may contribute to or exacerbate lung
15      inflammation.
16           Somatosensory neurons located in the dorsal root ganglia (DRG), innervate the upper
17      thoracic region of the airways  and extend their terminals under and between the epithelial lining
18      of the lumen.  Given this anatomical proximity, the sensory fibers and their tracheal epithelial
19      targets are the first resident cells to encounter inhaled pollutants, such as PM.  The differential
20      response of these cell types to  PM derived from various sources (i.e., industrial, residential,
21      volcanic) was examined with biophysical and immunological endpoints (Veronesi et al.,  2002a).
22      Although the majority of PM tested stimulated IL-6 release in both BEAS-2B epithelial cells and
23      DRG neurons in a receptor-mediated fashion, the degree of these responses was markedly higher
24      in sensory neurons. Epithelial cells are damaged or denuded in many common health disorders
25      (e.g., asthma,  viral infections), allowing PM particles to directly encounter the sensory terminals
26      and their acid sensitive receptors.  This differential sensitivity of target cells to PM suggests that
27      non-genetic factors (i.e., cell-cell interactions) may also affect the inflammatory response to PM
28      in individuals whose epithelial lining is damaged.
29           Another intracellular signaling pathway that can lead to diverse cellular responses such as
30      cell growth, differentiation,  proliferation, apoptosis, and stress responses to environmental
31      stimuli, is the phosphorylation-dependent, mitogen-activated protein kinase (MAPK).

        April 2002                                 7-81         DRAFT-DO NOT QUOTE OR CITE

-------
 1     Noncytotoxic levels of ROFA have been shown to induce significant dose- and time-dependent
 2     increases in protein tyrosine phosphate levels in BEAS cells (Samet et al., 1997). In a
 3     subsequent study, the effects of As, Cr, Cu, Fe, Ni, V, and Zn on the MAPK, extracellular
 4     receptor kinase (ERK), c-jun N-terminal kinase (INK), and P38 in BEAS cells were investigated
 5     (Samet et al., 1998). Noncytotoxic concentrations of As, V, and Zn induced a rapid
 6     phosphorylation of MAPK in BEAS cells.  Activity assays confirmed marked activation of ERK,
 7     INK, and P38 in BEAS cells exposed to As, V, and Zn. Cr and Cu exposure resulted in a
 8     relatively small activation of MAPK; whereas Fe and Ni did not activate MAPK. Similarly, the
 9     transcription factors c-Jun and ATF-2, substrates of INK and P38, respectively, were markedly
10     phosphorylated in BEAS cells treated with As, Cr, Cu, V, and Zn. The same acute exposure to
11     As, V, or Zn that activated MAPK was sufficient to induce a subsequent increase in IL-8 protein
12     expression in BEAS cells.  These data suggest that MAPK may mediate metal-induced
13     expression of inflammatory proteins in human bronchial epithelial cells.  The ability of ROFA to
14     induce activation of MAPKs in vivo was demonstrated by Silbajoris et al. (2000) (see Table 7-3).
15     In addition, Gercken et al. (1996) showed that the ROS production induced by PM was markedly
16     decreased by the inhibition of protein kinase C as well as phospholipase A2.
17           The major cellular response downstream of ROS and the  cell signaling pathways described
18     above is the production of inflammatory cytokines or other reactive mediators. In an effort to
19     determine the contribution of cyclooxygenase to the pulmonary responses to ROFA exposure
20     in vivo, Samet et al. (2000) intratracheally  instilled Sprague-Dawley rats with ROFA (200 or
21     500 ^g in 0.5 mL saline).  These animals were pretreated ip with 1 mg/kg NS398, a specific
22     prostaglandin H synthase 2 (COX2) inhibitor, 30 min prior to intratracheal exposure. At 12 h
23     after intratracheal instillations, ip injections (1 mL of NS398 in 20% ethanol in saline) were
24     repeated. ROFA treatment induced a marked increase in the level of PGE2 recovered in the BAL
25     fluid, which was effectively decreased by pretreating the animals with the COX2 inhibitor.
26     Immunohistochemical analyses of rat airway showed concomitant expression of COX2 in the
27     proximal airway epithelium of rats treated with soluble fraction of ROFA. This study further
28     showed that, although COX2 products participated in ROFA induced lung inflammation, the
29     COX metabolites  are not involved in IL-6 expression nor the influx of PMN influx into the
30     airway. However, the rationale for the use of intraperitoneal  challenge was not elaborated.


       April 2002                                7-82       DRAFT-DO NOT QUOTE OR CITE

-------
 1           The production of cytokines and mediators also has been shown to depend on the type of
 2     PM used in the experiments. A549 cells (a human airway epithelial cell line) were exposed to
 3     several PM, carbon black (CB, Elftex-12, Cabot Corp.), diesel soot (ND from NIST, LD
 4     produced from General Motors LH 6.2 V8 engine at light duty cycle), ROFA (from the heat
 5     exchange section of the Boston Edison), OAA (Ottowa ambient air PM, EHC-93), SiO2, and
 6     Ni3S2 at Img/cm2 (Seagrave and Nikula, 2000). Results indicated that (1) SiO2 and Ni3S2 caused
 7     dose dependent acute toxicity and apototic changes; (2) ROFA and ND were significant only at
 8     the highest concentrations; (3) SiO2 and Ni3S2 increased IL-8 (three and eight times over the
 9     control, respectively) at low concentrations but suppressed IL-8 at high concentrations; (4) OAA
10     and ROFA also induced IL-8 but lower than SiO2 and Ni3S2; and (5) both diesel soots suppressed
11     IL-8 production. The order of potency in alkaline phasphatase production is OAA > LD =
12     ND > ROFA » SiO2 = Ni3S2. These results demonstrated that the type of particle used has a
13     strong influence on the biological response.
14           Expression of MIP-2 and IL-6 genes was significantly upregulated as early as 6 h
15     post-ROFA-exposure in rat tracheal epithelial cells; whereas gene expression of iNOS was
16     maximally increased 24 h postexposure.  V but not Ni appeared to be mediating the effects  of
17     ROFA on gene expression.  Treatment with dimethylthiourea  inhibited both ROFA and V
18     induced gene expression in a dose-dependent manner (Dye et  al., 1999).
19           It appears that many biological  responses are produced by PM whether it is composed of a
20     single component or a complex mixture.  A technical approach is to use the newly developed
21     gene array to monitor the expressions of many mediator genes, which regulate complex and
22     coordinated cellular events involved  in tissue injury and repair, in a single assay.  Using an  array
23     consisting of 84  rat genes representing inflammatory and anti-inflammatory cytokines, growth
24     factors, adhesion molecules, stress proteins, transcription factors, and antioxidant enzymes,
25     Nadadur et al. (2000) and Nadadur and Kodavanti (2002) measured the pulmonary expressions of
26     these genes in rats intratracheally instilled with ROFA (3.3 mg/kg), NiSO4 (1.3 //mol/kg), and
27     VSO4 (2.2 //mol/kg).  Their data revealed a twofold induction of IL-6 and TEVIP-1 at 24 h post-
28     ROFA or Ni exposure.  The expression of cellular fibronectin (cFn-EIIIA), ICAM-1, IL-lb, and
29     iNOS gene also were increased 24 h  post-ROFA, V, or Ni exposure.  This study demonstrated
30     that gene array may provide a tool for screening the expression profile of tissue specific markers
31     following exposure to PM.

       April 2002                               7-83         DRAFT-DO NOT QUOTE OR CITE

-------
 1           To investigate the interaction between respiratory cells and PM, Kobzik (1995) showed that
 2      scavenger receptors are responsible for AM binding of unopsonized PM and that different
 3      mechanisms mediate binding of carbonaceous dusts such as DPM. In addition, surfactant
 4      components can increase AM phagocytosis of environmental particulates in vitro, but only
 5      slightly relative to the already avid AM uptake of unopsonized particles (Stringer and Kobzik,
 6      1996). Respiratory tract epithelial cells are also capable of binding with PM to secrete cytokine
 7      IL-8. Using a respiratory epithelial cell line (A549), Stringer et al. (1996) found that binding of
 8      particles to  epithelial cells was calcium-dependent for TiO2 and Fe2O3, while a-quartz binding
 9      was not calcium dependent. In addition, as observed in AMs, PM binding by A549 cells also
10      was mediated by scavenger receptors, albeit those distinct from the heparin-insensitive
11      acetylated-LDL receptor. Furthermore, a-quartz, but not TiO2 or CAPs, caused a dose-dependent
12      production of IL-8 (range 1 to 6 ng/mL), demonstrating a particle-specific spectrum of epithelial
13      cell cytokine (IL-8) response.
14
15      7.5.3.3 Other Potential Cellular and Molecular  Mechanisms
16           A potential mechanism involving in the alteration of surface tension may be related to
17      changes in the expression of matrix metalloproteinases (MMPs), such as pulmonary matrilysin
18      and gelatinase A and B, and tissue inhibitor of metalloproteinase (TEVIP) (Su et al., 2000a,b).
19      Sprague-Dawley rats exposed to ROFA by intratracheal injection  (2.5 mg/rat) had increased
20      mRNA levels of matrilysin, gelatinase A, and TEVIP-1. Gelatinase B, not expressed in control
21      animals, was increased significantly from 6 to 24 h following ROFA exposure. Alveolar
22      macrophages, epithelial cells, and inflammatory cells were major cellular sources for the
23      pulmonary MMP expression. The expression of Gelatinase B in rats exposed to the same dose of
24      ambient PM (<1.7 //m and  1.7 to 3.7 //m) collected from Washington, DC, was significantly
25      increased as compared to saline control; whereas the expression of TIMP-2 was suppressed.
26      Ambient PM between 3.7 and 20 //m also increased the Gelatinase B expression. Increases in
27      MMPs, which degrade most of the extracellular matrix, suggest that ROFA and ambient PM can
28      similarly increase the total pool of proteolytic activity to the lung and contribute in the
29      pathogenesis of PM-induced lung injury.
30           The role of sensory nerve receptors in the initiation of PM inflammation has been described
31      in a series of recent studies. Neuropeptide and acid-sensitive sensory irritant (i.e., capsaicin,

        April 2002                                 7-84        DRAFT-DO NOT QUOTE OR CITE

-------
 1      VR1) receptors were first identified on human bronchial epithelial cells (i.e., BEAS-2B).
 2      To address whether PM could initiate airway inflammation through these acid sensitive sensory
 3      receptors, BEAS-2B cells were exposed to ROFA and responded with an immediate increase in
 4      [Ca+2]; followed by a concentration-dependent release of inflammatory cytokine (i.e., IL-6, IL-8,
 5      TNFa) and their transcripts (Veronesi et al., 1999a). To test the relevance of neuropeptide or
 6      capsaicin VR1 receptors to these changes, BEAS-2B cells were pretreated with neuropeptide
 7      receptor antagonists or capsazepine (CPZ), the antagonist for the capsaicin (i.e., VR1) receptor.
 8      The neuropeptide receptor antagonists reduced ROFA-stimulated cytokine release by 25%-50%.
 9      However, pretreatment of cells with CPZ inhibited the immediate increases in [Ca+2];, diminished
10      transcript (i.e., IL-6, IL-8, TNFa) levels and reduced IL-6 cytokine release to control levels
11      (Veronesi et al., 1999b). The above studies suggested that ROFA inflammation was mediated by
12      acid sensitive VR1 receptors located on  the sensory nerve fibers that innervate the airway and on
13      epithelial target cells.
14           Colloidal particles (like ROFA and other PM) carry an inherently negative surface charge
15      (i.e., zeta potential) that attracts protons  from their vaporous milieu. These protons form a
16      neutralizing, positive ionic cloud around the individual particle (Hunter, 1981).  Since VR1
17      irritant receptors respond to acidity (i.e., protonic charge), experiments were designed to
18      determine if the surface charge carried by ROFA and other PM particles could biologically
19      activate cells and stimulate inflammatory cytokine release.  The mobility of ROFA particles was
20      measured in an electrically charged field (i.e., micro-electrophoresis) microscopically and their
21      zeta potential  calculated. Next, synthetic polymer microspheres (SPM) (i.e., polymethacrylic
22      acid nitrophenylacrylate microspheres) were prepared with attached carboxyl groups to  yield
23      SPM particles of the same size and with zeta potentials similar to ROFA (-29 + 0.9 mV)
24      particles. These SPM acted as ROFA surrogates with respect to their size and surface charge, but
25      lacked all other contaminants that were thought to be responsible for its toxicity (e.g., transition
26      metals, sulfates, volatile organics and  biologicals). Similar concentrations of SPM and  ROFA
27      particles were used to test BEAS-2B cells and mouse dorsal root ganglia (DRG) sensory neurons,
28      both targets of inhaled PM. Equivalent degrees of biological activation (i.e., increase in
29      intracellular calcium, [Ca+2];, IL-6 release) occurred in both cell types in response to either ROFA
30      or SPM and both responses could be reduced by antagonists to VR1 receptors or acid-sensitive
31      pathways.  Neutrally charged SPM (i.e.,  zeta potential of 0 mV), however, failed to stimulated

        April 2002                                 7-85        DRAFT-DO NOT QUOTE OR CITE

-------
 1      increases in [Ca+2]; or IL-6 release (Oortgiesen et al., 2000). To expand on these data, a larger set
 2      of PM was obtained from urban (St. Louis, Ottawa), residential (wood stove), volcanic (Mt. St.
 3      Helen), and industrial (oil fly ash, coal fly ash) sources. Each PM sample was described
 4      physicochemically (i.e., size and number of visible particles, acidity, zeta potential) and used to
 5      test BEAS-2B epithelial cells. The resulting biological effect (i.e., increases in [Ca+2];, IL-6
 6      release) was related to their physicochemical descriptions. When examined by linear regression
 7      analysis, the only measured physicochemical property that correlated with increases in [Ca+2]; and
 8      IL-6 release was the zeta potential of the visible particles  (r2 > 0.97) (Veronesi et al., 2002b).
 9      Together, these studies have demonstrated a neurogenic basis for PM inflammation by which the
10      proton cloud associated with negatively-charged colloidal PM particles can activate acid-
11      sensitive VR1 receptors found on human airway epithelial cells and sensory terminals. This
12      activation results in an immediate influx of calcium and the release of inflammatory
13      neuropeptides and cytokines which proceed to initiate and sustain inflammatory events in the
14      airways through the pathophysiology of neurogenic inflammation (Veronesi and Oortgiesen,
15      2001).
16
17      7.5.4  Specific Particle Size and Surface Area Effects
18           Most particles used in laboratory animal toxicology and occupational studies are greater
19      than 0.1 //m in size.  However, the enormous number and huge surface area of the ultrafine
20      particles demonstrate the importance of considering the size of the particle in assessing response.
21      Ultrafine particles with a diameter of 20 nm when inhaled at the same mass concentration have a
22      number concentration that is approximately 6 orders of magnitude higher than for a 2.5-//m
23      diameter particle; particle surface area is also greatly increased (Table 7-11).
24           Many studies summarized in U.S. Environmental Protection Agency (1996a), as well as in
25      this document, suggest that the surface of particles or substances that are released from the
26      surface (e.g., transition metals) interact with the biological system, and that surface-associated
27      free radicals or free radical-generating systems may be responsible for toxicity. Thus, if ultrafine
28      particles were to cause toxicity by a transition metal-mediated mechanism, for example, then the
29      relatively large surface area for a given mass of ultrafine particles would mean high
30      concentrations of transition metals being available to cause oxidative stress to cells.
31
        April 2002                                7-86        DRAFT-DO NOT QUOTE OR CITE

-------
               TABLE 7-11.  NUMBERS AND SURFACE AREAS OF MONODISPERSE
                PARTICLES OF UNIT DENSITY OF DIFFERENT SIZES AT A MASS
                                   CONCENTRATION OF 10
Particle Diameter
Cwm)





Source:
0.02
0.1
0.5
1.0
2.5
Oberdorster (1996).
Particle Number
(per cm3 air)
2,400,000
19,100
153
19
1.2

Particle Surface Area
(,wm2 per cm3 air)
3,016
600
120
60
24

 1           Two groups have examined the toxic differences between fine and ultrafme particles, with
 2      the general finding that the ultrafme particles show a significantly greater response at similar
 3      mass doses (Oberdorster et al., 1992; Li et al., 1996,  1997, 1999). However, only a few studies
 4      have investigated the ability of ultrafme particles to generate a greater oxidative stress when
 5      compared to fine particles of the same material. Studies by Gilmour et al. (1996) have shown
 6      that at equal mass, ultrafme TiO2 caused more plasmid DNA strand breaks than fine TiO2.  This
 7      effect could be inhibited with mannitol. Osier and Oberdorster (1997) compared the response of
 8      rats (F344) exposed by intratracheal inhalation to "fine" (approximately 250 nm) and "ultrafme"
 9      (approximately 21 nm) TiO2 particles with rats exposed to similar doses by intratracheal
10      instillation. Animals receiving particles through inhalation showed a smaller pulmonary
11      response, measured by BAL parameters, in both severity and persistence, when compared with
12      those animals receiving particles through instillation.  These results demonstrate a difference in
13      pulmonary response to an inhaled versus an instilled dose, which may result from differences in
14      dose rate, particle distribution,  or altered clearance between the two methods.  Consistent with
15      these in vivo  studies, Finkelstein et al. (1997) has shown that exposing primary cultures of rat
16      Type n cells to 10 //g/mL ultrafme TiO2 (20  nm) causes increased TNF  and IL-1 release
17      throughout the entire 48-h incubation period. In contrast, fine TiO2 (200 nm) had no effect.
18      In addition, ultrafme polystyrene carboxylate-modified microspheres (UFP, fluorospheres,
19      molecular probes 44 ± 5 nm) have been shown to induce a significant enhancement of both
20      substance P and histamine release after administration of capsaicin (10"4 M), to stimulate C-fiber,
21      and carbachol (10"4 M), a cholinergic agonist in rabbit intratracheally instilled with UFP

        April 2002                                7-87        DRAFT-DO NOT QUOTE OR CITE

-------
 1      (Nemmar et al., 1999). A significant increase in histamine release also was recorded in the
 2      UFP-instilled group following the administration of both Substance P (10"6 M) plus thiorpan
 3      (10"5 M) and compound 48/80 (C48/80, 10"3 M) to stimulate mast cells. BAL analysis showed an
 4      influx of PMN, an increase in total protein concentration, and an increase in lung wet weight/dry
 5      weight ratio. Electron microscopy showed that both epithelial and endothelial injuries were
 6      observed.  The pretreatment of rabbits in vivo with a mixture of either SR 140333 and SR 48368,
 7      a tachykinin NKl and NK2 receptor antagonist, or a mixture of terfenadine and cimetidine,
 8      a histamine Hx and H2 receptor antagonist, prevented UFP-induced PMN influx and increased
 9      protein and lung WW/DW ratio.
10           Given the assumption that the chemical composition of ultrafme particles is the same as
11      larger particles, it is believed that ultrafme particles cause greater cellular injury because of the
12      relatively large surface area for a given mass.  However, in a study that compared the response to
13      carbon black particles of two different sizes, Li et al. (1999) demonstrated that in the instillation
14      model, a localized dose of particle  over a certain level causes the particle mass to dominate the
15      response, rather than the surface area. Ultrafme carbon black (ufCB, Printex 90), 14 nm in
16      diameter, and fine carbon  black (CB, Huber 990), 260 nm in diameter, were instilled
17      intratracheally in rats and  BAL profile at 6 h was assessed. At mass of 125 //g or below, ufCB
18      generated a greater response (increase LDH, epithelial permeability, decrease in GSH, TNF, and
19      NO production) than fine  CB at various times postexposure. However, higher dose of CB caused
20      more PMN influx than the ufCB. In contrast to the effect of CB, which showed dose-related
21      increasing inflammatory response,  ufCB at the highest dose caused less of a neutrophil influx
22      than at the lower dose, confirming  earlier work reported by Oberdorster et al. (1992). Moreover,
23      when the PMN influx was expressed as a function of surface area, CB produced greater response
24      than ufCB at all doses used in this  study.  Although particle insterstitialization with a consequent
25      change in the chemotatic gradient for PMN was offered as an explanation, these results need
26      further scrutiny.
27           Oberdorster et al. (2000) recently completed a series of studies in rats and mice using
28      ultrafme particles of various chemical compositions (PTFE, TiO2, C, Fe, Fe2O3, Pt, V, and V2O5).
29      In old rats sensitized with  endotoxin and exposed to ozone plus ultrafme carbon particles, they
30      found a ninefold greater release of reactive oxygen species in old rats than in similarly treated


        April 2002                                7-88        DRAFT-DO NOT QUOTE OR CITE

-------
 1      young rats. Exposure to ultrafine PM alone in sensitized old rats also caused an inflammatory
 2      response.
 3           Although the exact mechanism of ultrafme-induced lung injury remains unclear, it is likely
 4      that ultrafine particles, because of their small size, can easily penetrate the airway epithelium and
 5      cause cellular damage.  Using electron microscopy to examine rat tracheal explants treated with
 6      fine (0.12 //m) and ultrafine (0.021 //m) TiO2 particles for 3 or 7 days, Churg et al. (1998) found
 7      both size particles in the epithelium at both time points, but in the subepithelial tissues, they were
 8      found only at Day 7. The volume proportion (the volume of TiO2 over the entire volume of
 9      epithelium or subepithelium area) of both fine and ultrafine particles in the epithelium increased
10      from 3 to 7 days. It was greater for ultrafine at 3 days but was greater for fine at 7 days. The
11      volume proportion of particles in the subepithelium at day 7 was equal for both particles, but the
12      ratio of epithelial to subepithelial volume proportion was 2:1 for fine and 1:1 for ultrafine.
13      Ultrafine particles persisted in the tissue as relatively large  aggregates; whereas the size of fine
14      particle aggregates became smaller over time.  Ultrafine particles appeared to enter the
15      epithelium faster and, once in the epithelium, a greater proportion of them were translocated to
16      the subepithelial space compared to fine particles. However, the authors assumed that the
17      volume proportion is representative of particle number and the number of particles reaching the
18      interstitial space is directly proportional to the number applied (i.e., there is no preferential
19      transport from lumen to interstitium by size). These data are in contrast to the results of
20      instillation or inhalation of fine and ultrafine TIO2 particles reported earlier (Ferin et al., 1990,
21      1992).  However, the explant and intratracheal instillation test systems differ in many aspects
22      making direct comparisons difficult. Limitations of the explant test system include traumatizing
23      the explanted tissue, introducing potential artifacts through the use of liquid suspension for
24      exposure, the absence of inflammatory cells, and possible overloading of the explants with dust.
25           Only two studies examined the influence of specific surface area on biological activity
26      (Lison et al., 1997; Oettinger et al., 1999). The biological responses to various MnO2 dusts with
27      different specific surface area (0.16, 0.5, 17, and 62 m2/g) were compared in vitro and  in vivo
28      (Lison et al., 1997). In both systems, the results show that the amplitude of the response is
29      dependent on the total surface area that is in contact with the biological system, indicating that
30      surface chemistry phenomena are involved in the biological reactivity.  Freshly ground particles
31      with a specific surface area of 5  m2/g also were examined in vitro.  These particles exhibited an

        April 2002                                7-89        DRAFT-DO NOT QUOTE OR CITE

-------
 1      enhanced cytotoxic activity that was almost equivalent to that of particles with a specific surface
 2      area of 62 m2/g, indicating that undefined reactive sites produced at the particle surface by
 3      mechanical cleavage also may contribute to the toxicity of insoluble particles.  In another study,
 4      two types of carbon black particles, Printex 90 (P90, Degussa, Germany, formed by controlled
 5      combustion, consists of defined granules with specific surface area of 300 m2/g and particle size
 6      of 14 nm) and FR 101 (Degussa, Germany, with specific surface area of 20 m2/g and particle size
 7      of <95 nm, has a  coarse structure, and the ability to adsorb polycyclic and other carbons) were
 8      used in the study  (Oettinger et al., 1999). Exposure of AMs to 100 //g/106 cells of FR 101 and
 9      P90 resulted in a  1.4- and 2.1-fold increase in ROS release.  These exposures also caused a
10      fourfold up-regulation of NF-KB gene expression. These studies indicated that PM of single
11      component with larger surface area produce greater biological response than similar particles
12      with smaller surface area.  By exposing bovine AMs to metal oxide coated silica particles,
13      Schluter et al. (1995) showed that most of the metal coatings (As, Ce, Fe, Mn,  Ni, Pb, and V) had
14      no effect on ROS production by these cells. However, coating with CuO markedly lowered the
15      O2" and H2O2, whereas V(IV)  increases both ROI. This study demonstrated that, in addition to
16      specific area, chemical composition of the particle surface also influence its cellular response.
17      Thus, ultrafine particles have  the potential to  significantly contribute to the adverse effects of
18      PM. These studies, however, have overlooked the portion of ambient ultrafine particles that are
19      not solid in form. Droplets (e.g., sulfuric acid droplets) and organic based ultrafine particles do
20      exist in the ambient environment, but their role in the adverse effects of ultrafine particles has
21      been ignored. Moreover, the  ability of these droplet ultrafine particles to spread, disperse, or
22      dissolve after contact with liquid surface layers must be considered.
23
24      7.5.5 Pathophysiological Mechanisms for the Effects of Low Concentrations
25            of Particulate Air Pollution
26           The pathophysiological  mechanisms involved in PM-associated cardiovascular and
27      respiratory health effects still  are not  elucidated fully, but progress has been made since the  1996
28      PM AQCD (U. S. Environmental Protection Agency, 1996a) was prepared. This section
29      summarizes several current hypotheses and reviews the toxicological evidence for these potential
30      pathophysiological mechanisms.
31

        April 2002                                7-90        DRAFT-DO NOT QUOTE OR CITE

-------
 1      7.5.5.1 Direct Pulmonary Effects
 2           When the 1996 PM AQCD (U. S. Environmental Protection Agency, 1996a) was written,
 3      the lung was thought to be the primary organ that was affected by particulate air pollution.
 4      Although the lung still is a primary organ affected by PM inhalation, there is growing
 5      toxicological and epidemiological evidence that the cardiovascular system is affected,  as well,
 6      and may be a co-primary organ system related to certain health endpoints such as mortality.
 7      Nonetheless, understanding how particulate air pollution causes or exacerbates respiratory
 8      disease remains an important goal. There is some toxicological evidence for the following three
 9      mechanisms for direct pulmonary effects.
10
11      Particulate Air Pollution Causes Lung Injury and Inflammation
12           In the last few years, numerous studies have shown that instilled and inhaled ROFA, a
13      product of fossil fuel combustion, can cause substantial lung injury and inflammation.  The toxic
14      effects of ROFA are largely caused by its high content of soluble metals, and some of the
15      pulmonary effects of ROFA can be reproduced by equivalent exposures to soluble metal salts.
16      In contrast, controlled exposures of animals to sulfuric acid aerosols, acid-coated carbon, and
17      sulfate salts cause little lung injury or inflammation, even at high concentrations.  Inhalation of
18      concentrated ambient PM (which contains only small amounts of metals) by laboratory animals
19      at concentrations in the range of 100 to  1000 //g/m3 have been shown in some (but not all)
20      studies to cause mild pulmonary injury and inflammation.  Rats with SO2-induced bronchitis and
21      monocrotaline-treated rats have been reported to have a greater inflammatory response to
22      concentrated ambient PM than normal rats.  These studies suggest that exacerbation of
23      respiratory disease by ambient PM may be caused in part by lung injury and inflammation.
24
25      Particulate Air Pollution Causes Increased Susceptibility to Respiratory Infections
26           At this time there are no newly published studies on the effects of inhaled concentrated
27      ambient PM on host susceptibility to infectious agents. Ohtsuka et al. (2000a,b) have shown that
28      in vivo exposure of mice to acid-coated carbon particles at a mass concentration of 10,000 //g/m3
29      carbon black causes decreased phagocytic activity of alveolar macrophages, even in the absence
30      of lung injury.
31

        April 2002                                 7-91         DRAFT-DO NOT QUOTE OR CITE

-------
 1     Particulate Air Pollution Increases Airway Reactivity and Exacerbates Asthma
 2           The strongest evidence supporting this hypothesis is from studies on diesel particulate
 3     matter (DPM).  DPM has been shown to increase production of antigen-specific IgE in mice and
 4     humans (summarized in Section 7.2.1.2).  In vitro studies have suggested that the organic
 5     fraction of DPM is involved in the increased IgE production. ROFA leachate also has been
 6     shown to enhance antigen-specific airway reactivity in mice (Goldsmith et al., 1999), indicating
 7     that soluble metals can also enhance an allergic response. However, in this same study, exposure
 8     of mice to concentrated ambient PM did not affect antigen-specific airway reactivity. It is
 9     premature to conclude from this one experiment that concentrated ambient PM does not
10     exacerbate allergic airways disease because the chemical composition of the PM (as indicated by
11     studies with DPM and ROFA) may be more important than the mass concentration.
12
13     7.5.5.2 Systemic Effects Secondary to Lung Injury
14           When the 1996 PM AQCD was written, it was thought that cardiovascular-related
15     morbidity and mortality most likely would be secondary to impairment of oxygenation  or some
16     other consequence of lung injury and inflammation. Newly available toxicologic studies provide
17     some additional evidence regarding such possibilities.
18
19     Lung Injury from Inhaled Particulate Matter Causes Impairment of Oxygenation and
20     Increased Work of Breathing That Adversely Affects the Heart
71
22           Instillation of ROFA has been shown to cause a 50% mortality rate in monocrotaline-
23     treated rats (Watkinson et al., 2000a,b). Although blood oxygen levels were not measured in this
24     study, there were ECG abnormalities consistent with severe hypoxemia in about half of the rats
25     that subsequently died. Given the severe inflammatory effects of instilled ROFA and the fact
26     that monocrotaline-treated rats have increased lung permeability as well as pulmonary
27     hypertension, it is plausible that instilled ROFA can cause severe hypoxemia leading to death in
28     this rat model. Results from studies in which animals (normal and compromised) were exposed
29     to concentrated ambient PM (at concentrations many times higher than would be encountered in
30     the United States) indicate that ambient PM is unlikely to cause severe disturbances in
31     oxygenation or pulmonary function. However, even a modest decrease in oxygenation can have
32     serious consequences in individuals with ischemic heart disease.  Kleinman et al.  (1998) has

       April 2002                                7-92       DRAFT-DO NOT QUOTE OR CITE

-------
 1      shown that a reduction in arterial blood saturation from 98 to 94% by either mild hypoxia or by
 2      exposure to 100 ppm CO significantly reduced the time to onset of angina in exercising
 3      volunteers.  Thus, information is needed on the effects of PM on arterial blood gases and
 4      pulmonary function to fully address the above hypothesis.
 5
 6      Lung Inflammation and Cytokine Production Cause Adverse Systemic Hemodynamic Effects
 1          It has been suggested that systemic effects of particulate air pollution may result from
 8      activation of cytokine production in the lung (Li et al., 1997). In support of this idea,
 9      monocrotaline-treated rats exposed to inhaled ROFA (15,000 //g/m3, 6 h/day for 3 days) showed
10      increased pulmonary cytokine gene expression, bradycardia, hypothermia, and increased
11      arrhythmias (Watkinson et al., 2000a,b). However, spontaneously hypertensive rats had a similar
12      cardiovascular response to inhaled ROFA  (except that they also developed ST segment
13      depression) with no increase in pulmonary cytokine gene expression.  Studies in dogs exposed to
14      concentrated ambient PM showed minimal pulmonary inflammation and no positive staining for
15      IL-8, IL-1,  or TNF in airway biopsies. However, there was a significant decrease in the time of
16      onset of ischemic ECG changes following coronary artery occlusion in PM-exposed dogs
17      compared to controls (Godleski  et al., 2000).  Thus, the link between changes in the production
18      of cytokines in the lung and cardiovascular function is not clear-cut, and basic information on the
19      effects of mild pulmonary injury on cardiovascular function is needed to understand the
20      mechanisms by which inhaled PM affects the heart.
21
22      Lung Inflammation from Inhaled Particulate Matter Causes Increased Blood Coagulability
23      That Increases the Risk of Heart Attacks and Strokes
24
25          There is abundant evidence linking risk of heart attacks and strokes to small prothrombotic
26      changes in the blood coagulation system.  However, the published toxicological evidence that
27      moderate lung inflammation causes increased blood coagulability is inconsistent.  Ohio et al.
28      (2000a) have shown that inhalation of concentrated ambient PM in healthy nonsmokers causes
29      increased levels of blood fibrinogen. Gardner et al. (2000) have shown that a high dose
30      (8,300 //g/kg) of instilled ROFA in rats causes increased levels of fibrinogen, but no effect was
31      seen at lower doses. Exposure of dogs to concentrated ambient PM had no effect on fibrinogen
32      levels (Godleski et al., 2000).  The coagulation system is as multifaceted and complex as the

        April 2002                                7-93        DRAFT-DO NOT QUOTE OR CITE

-------
 1      immune system, and there are many other sensitive and clinically significant parameters that
 2      should be examined in addition to fibrinogen. Thus, it is premature to draw any conclusions on
 3      the relationship between PM and blood coagulation.
 4
 5      Interaction of Particulate Matter with the Lung Affects Hematopoiesis
 6           Terashima et al. (1997) found that instillation of fine carbon particles (20,000 //g/rabbit)
 7      stimulated release of PMNs from the bone marrow. In further support of this hypothesis,  Gordon
 8      and colleagues reported that the percentage of PMNs in the peripheral blood increased in  rats
 9      exposed to ambient PM in some but not all exposures. On the other hand, Godleski et al. (2000)
10      found no changes in peripheral blood counts  of dogs exposed to concentrated ambient PM.
11      Thus, direct evidence that PM ambient concentrations can affect hematopoiesis remains to be
12      demonstrated.
13
14      7.5.5.3  Direct Effects on the Heart
15           Changes in heart rate, heart rate variability, and conductance associated with ambient PM
16      exposure have been reported in animal studies (Godleski et al., 2000; Gordon et al., 2000;
17      Watkinson et al., 2000a,b; Campen et  al., 2000), in several human panel studies (described in
18      Chapter 8), and in a reanalysis of data  from the MONICA study (Peters et al., 1997).  Some of
19      these studies included endpoints related to respiratory effects but few significant adverse
20      respiratory changes were detected.  This raises the possibility that ambient PM may have effects
21      on the heart that are independent of adverse changes in the lung. There is certainly precedent for
22      this idea. For example, tobacco smoke (which is a mixture of combustion-generated gases and
23      PM) causes cardiovascular disease by  mechanisms that are independent of its effect on the lung.
24      Two types of hypothesized direct effects of PM on the heart are noted below.
25
26      Inhaled Particulate Matter Affects the Heart by Uptake of Particles into the Circulation
27      or Release of a Soluble Substances into the  Circulation,
29           Drugs can be rapidly and efficiently delivered to the systemic circulation by inhalation.
30      This implies that the pulmonary vasculature absorbs inhaled materials, including charged
31      substances such as small proteins and peptides. Cigarettes are a widely used method for
32      delivering nicotine to the blood stream. It is  likely that soluble materials absorbed onto airborne

        April 2002                                7-94        DRAFT-DO NOT QUOTE OR CITE

-------
 1      particles find their way into the blood stream, but it is not clear whether the particles themselves
 2      enter the blood. It is anticipated that more information will be available on this important
 3      question in the next few years.
 4
 5      Inhaled Particulate Matter Affects Autonomic Control of the Heart and
 6      Cardiovascular System
 8          There is growing evidence for this idea as described above. This raises the question of how
 9      inhaled particles could affect the autonomic nervous system. Activation of neural receptors in
10      the lung is a logical area to investigate.  Studies in conscious rats have shown that inhalation of
11      wood smoke causes marked changes in sympathetic and parasympathetic input to the
12      cardiovascular system that are mediated by neural reflexes (Nakamura and Hayashida, 1992).
13      Although research on airway neural receptors and neural-mediated reflexes is a well established
14      discipline, the cardiovascular effects of stimulating airway receptors continue to receive less
15      attention than the  pulmonary effects. Previous studies of airway reflex-mediated cardiac effects
16      usually employed very high doses of chemical irritants, and the results may not be applicable to
17      air pollutants.  There is a need for basic physiological studies to examine effects on
18      cardiovascular system when airway and  alveolar neural receptors are stimulated in a manner
19      relevant to air pollutants.
20
21
22      7.6 RESPONSES TO PARTICULATE MATTER AND GASEOUS
23          POLLUTANT MIXTURES
24          Ambient PM itself is a mixture of particles of varying size and composition. The following
25      discussion examines effects of mixtures of ambient PM, or PM surrogates, with gaseous
26      pollutants. Ambient PM co-exists in indoor and outdoor air with a number of co-pollutant gases,
27      including ozone, sulfur dioxide, oxides of nitrogen, and carbon monoxide.  Toxicological
28      interactions between PM and gaseous co-pollutants may be antagonistic, additive, or synergistic
29      (Mauderly, 1993). The presence and nature of any interaction appears to depend on the chemcial
30      composition, size, concentration and ratios of pollutants in the mixture, exposure duration, and
31      the endpoint being examined. It may be difficult to predict a priori from the presence of certain
        April 2002                                7-95        DRAFT-DO NOT QUOTE OR CITE

-------
 1      pollutants whether any interaction will occur and, if there is interaction, whether it will be
 2      synergistic, additive, or antagonistic (Table 7-12).
 3           Mechanisms responsible for the various forms of interaction are speculative. In terms of
 4      potential health effects, the greatest hazard from pollutant interaction is the possibility of synergy
 5      between particles and gases, especially if effects occur at concentrations at which no effects
 6      occur when individual constituents are inhaled. Various physical and chemical mechanisms may
 7      underlie synergism. For example, physical adsorption or absorption of some material on a
 8      particle could result in transport to more sensitive sites,  or sites where this material would not
 9      normally be deposited in toxic amounts. This physical process may explain the interaction  found
10      in studies of mixtures of carbon black and formaldehyde or of carbon black and acrolein (Jakab,
11      1992,1993).
12           Chemical interactions between PM and gases can occur on particle surfaces, thus, forming
13      secondary products that may be more active lexicologically than the primary materials and that
14      can then be carried to a sensitive site. The hypothesis of such chemical interactions has been
15      examined in the gas and particle exposure studies by Amdur and colleagues (Amdur and Chen,
16      1989; Chen et al., 1992) and Jakab and colleagues (Jakab and Hemenway, 1993; Jakab et al.,
17      1996).  These investigators have suggested that synergism occurs as secondary chemical species
18      are produced, especially under conditions of increased temperature and relative humidity.
19           Another potential mechanism of gas-particle interaction may involve a pollutant-induced
20      change in the local microenvironment of the lung,  enhancing the effects of the co-pollutant.
21      For example,  Last et al. (1984)  suggested that the observed synergism between ozone (O3) and
22      acid sulfates in rats was due to a decrease in the local microenvironmental pH of the lung
23      following deposition of acid, enhancing the effects of O3 by producing a change in the reactivity
24      or residence time of reactants, such as radicals, involved in O3-induced tissue injury. Likewise,
25      Pinkerton et al (1989) showed increased retention of the mass and number of asbestos fibers in
26      rats exposed to O3, suggesting an increase in lung fiber burden due to exposure to this gaseous
27      pollutant.
28           As noted in U.S. Environmental Protection Agency (1996a), the toxicology database for
29      mixtures containing PM other than acid sulfates was and is still quite sparse. Vincent et al.
30      (1997) exposed rats to 0.8 ppm O3 in combination with 5 or 50 mg/m3 of resuspended urban
31      particles for 4 h. Although PM alone caused no change in cell proliferation (3H-thymidine

        April 2002                                 7-96        DRAFT-DO NOT QUOTE OR CITE

-------
                    TABLE 7-12. RESPIRATORY AND CARDIOVASCULAR EFFECTS OF MIXTURES
to
o
o
to
H

6
o


o
H

O
o
HH
H
W
Species, Gender,
Strain Age, or
Body Weight
Rats, Fischer
NNia, male,
22 to 24 mo old




Rats




Humans; healthy
15 M, 10 F,
34.9+10 years of
age
Humans; healthy
children




Humans; 59
healthy childern
in Mexico City;
19 controls in
Gulf port town
Humans; 15
healthy children
in Mexico City;
1 1 children in
Veracruz; 4-15
years of age
Humans; 83
healthy children
in Mexico City;
24 children in
Isla Mujeres;
6- 12 years of
age
Gases and PM
Carbon,
ammonium
bisulfate,
and O3



O3 and Ottawa
urban dust



CAPs



Ambient gases
and particles




Ambient gases
and particles



Ambient gases
and particles




Ambient gases
and particles





Particle
Exposure Technique Mass Concentration Size
Inhalation 50 ^g/m3 carbon + 0.4/^mMMAD
70 fj.g/m1 ammonium og = 2.0
bisulfate + 0.2 ppm
O3 or 100 Mg/m3
carbon +140 i/g/m3
ammonium bisulfate
+ 0.2 ppm O3
Inhalation 40,000 ,ug/m3 and 4.5 ,um
0.8 ppm O3 MM AD



Inhalation ISO^g/m3 PM25
120 ppb O3


Natural 24 h
exposure in
Southwest
Metropolitan
Mexico City
(SWMMC)
Natural 24 h
exposure in
SWMMC compared
to low pollution
Gulf of Mexico
Natural 24 h
exposure in
SWMMC compared
to low pollution
Gulf Coast

Natural 24 h
exposure in
SWMMC compared
to low pollution
Caribbean


Exposure Cardiopulmonary Effects of Inhaled
Duration PM and Gases
4 h/day, No changes in protein concentration in lavage
3 days/week for fluid or in prolyl 4-hydroxylase activity in
4 weeks blood. Slight, but statistically significant
decreases in plasma fibronectin in animals
exposed to the combined atmospheres
compared to animals exposed to O3 alone.

Single 4-h Co-exposure to particles potentiated O3-induced
exposure followed septal cellurity. Enhanced septal thickening
by 20 h clean air associated with elevated production of
macrophage inflammatory protein-2 and
endothelin 1 by lung lavage cells.
2 h Acute brachial artery vasoconstriction as
determined by vascular ultrasonography
performed before and 10 min after exposure.

Radiological evidence of lung hyperinflation
from chest X-rays.




Increased upper and lower respiratory
symptoms; bilateral symmetric mild lung
hyperinflation from chest X-rays.


Nasal biopsies revealed increased basal,
ciliated, goblet, and squamous metaplastic
and intermediate cells; cellular abnormalities
and possible dyskinesia were noted.


Nasal biopsies revealed p53 accumulation by
immunochemistry; increased upper and lower
respiratory symptoms.




Reference
Bolarin et al.
(1997)





Bouthillier et al.
(1998)



Brook et al. (2002)



Calderon-Garciduenas
et al. (2000a)




Calderon-Garciduenas
et al. (2000b)



Calderon-Garciduenas
etal. (2001a)




Calderon-Garciduenas
etal. (2001b)






-------
 
-------
                            TABLE 7-12 (cont'd).  RESPIRATORY AND  CARDIOVASCULAR EFFECTS OF MIXTURES
13.
to
o
o
to
Species, Gender,
Strain Age, or
Body Weight
Rats
Humans,
children, healthy
and asthmatic
Gases and PM
H2SO4 and O3
H2S04,
SO2, and O3
Exposure Technique
Inhalation,
whole body
Inhalation
Mass Concentration
20 to 150 Mg/m3
H2SO4and0.12or
0.2 ppm O3
60 to 140 Mg/m3
H2SO4, 0.1 ppm
SO2, and 0.1 ppm O3
Particle
Size
0.4 to 0.8 Mm
0.6 Mm H2SO4
Exposure
Duration
Intermittent
(12 h/day) or
continuous
exposure for up
to 90 days
Single 4-h
exposure with
intermittent
Cardiopulmonary Effects of Inhaled
PM and Gases
No interactive effect of H2SO4 and O3 on
biochemical and morphometric endpoints.
A positive association between acid concentration
and symptoms, but not spirometry, in asthmatic
children. No changes in healthy children.
Reference
Last and
Pinkerton (1997)
Linn et al.
(1997)

-------
 1      labeling), co-exposure to either concentration of resuspended PM with O3 greatly potentiated the
 2      proliferative effects of exposure to O3 alone.  These interactive changes occurred in epithelial
 3      cells of the terminal bronchioles and the alveolar ducts.  These findings using resuspended dusts,
 4      although at high concentrations, are consistent with studies demonstrating interaction between
 5      sulfuric acid (H2SO4) aerosols and O3. Kimmel and colleagues (1997) examined the effect of
 6      acute co-exposure to O3 and fine or ultrafine H2SO4 aerosols on rat lung morphology.  They
 7      determined morphometrically that alveolar septal volume was increased in animals co-exposed to
 8      O3 and ultrafine, but not fine, H2SO4. Interestingly, cell  labeling, an index of proliferative cell
 9      changes, was increased only in animals co-exposed to fine H2SO4 and O3, as compared to animals
10      exposed to O3 alone.  Importantly, Last and Pinkerton (1997) extended their previous work and
11      found that subchronic exposure to acid aerosols (20 to 150 //g/m3 H2SO4) had no interactive
12      effect on the biochemical and morphometric changes produced by either intermittent or
13      continuous O3 exposure (0.12 to 0.2 ppm). Thus, the interactive effects of O3 and acid aerosol
14      co-exposure in the lung disappeared during the long-term exposure.  Sindhu et al. (1998)
15      observed an increase in rat lung putrescine levels after repeated, combined exposures to  O3 and a
16      nitric acid vapor.
17           Kleinman et al.  (1999) examined the effects of O3  plus fine, H2SO4-coated, carbon particles
18      (MMAD = 0.26 //m) for 1 or 5 days.  They found the inflammatory response with the O3-particle
19      mixture was greater after 5 days (4 h/day) than after Day 1. This contrasted with O3 exposure
20      alone (0.4 ppm), which caused marked inflammation on acute exposure, but no inflammation
21      after 5 consecutive days of exposure.
22           Kleinman et al.  (2000) examined the effects of a mixture of elemental carbon particles, O3,
23      and ammonium bisulfate on rat lung collagen content and macrophage activity.  Decreases in
24      lung collagen, and increases in macrophage respiratory burst and phagocytosis were observed
25      relative to other pollutant combinations. Mautz et al.  (2001) used a similar mixture (i.e.,
26      elemental carbon particles, O3, ammonium bisulfate, but with NO2 also) and exposure regimen as
27      Kleinman (2000). There were decreases in pulmonary macrophage Fc-receptor binding  and
28      phagocytosis and increases in acid phosphatase staining. Bronchoalveolar epithelial permeability
29      cell proliferation were increased.  Altered breathing-patterns were also observed, with some
30      adaptations  occurring.


        April 2002                                7-100        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Studies have examined interactions between carbon particles and gaseous co-pollutants.
 2      Jakab et al. (1996) challenged mice with a single 4-h exposure to a high concentration of carbon
 3      particles (10 mg/m3) in the presence of SO2 at low and high relative humidities. Macrophage
 4      phagocytosis was depressed significantly only in mice exposed to the combined pollutants under
 5      high relative humidity conditions.  This study suggests that fine carbon particles can serve as an
 6      effective carrier for acidic sulfates where chemical conversion of adsorbed SO2 to acid sulfate
 7      species occurred. Interestingly, the depression in macrophage function was present as late as
 8      7 days postexposure.  Bolarin et al. (1997) exposed rats to only 50 or 100 //g/m3 carbon particles
 9      in combination with ammonium bisulfate and O3. Despite 4 weeks of exposure, they observed
10      no changes in protein concentration in lavage fluid or blood prolyl 4-hydroxylase, an enzyme
11      involved in collagen metabolism.  Slight decreases in plasma fibronectin were present in animals
12      exposed to the combined pollutants versus O3 alone. Thus as, previously noted, the potential for
13      adverse effects in the lungs of animals challenged with a combined exposure to particles and
14      gaseous pollutants is dependent on numerous factors, including the gaseous co-pollutant,
15      concentration, and time.
16           In a complex series of exposures, Oberdorster and colleagues examined the interaction of
17      ultrafine carbon particles (100 //g/m3) and O3 (1 ppm) in young and old Fischer 344 rats that were
18      pretreated with aerosolized endotoxin (Elder et al., 2000a,b).  In old rats, exposure to carbon and
19      O3 produced an interaction that resulted in a greater influx in neutrophils than that produced by
20      either agent alone. This interaction was not seen in young rats. Oxidant release from lavage
21      fluid cells was also assessed and the combination of endotoxin, carbon particles, and O3
22      produced an increase in oxidant release in old rats. This combination produced the opposite
23      response in the cells recovered from the lungs of the young rats,  indicating that the lungs of the
24      aged animals underwent greater oxidative stress in response to this complex pollutant mix of
25      particles, O3, and a biogenic agent.
26           Wagner et al. (2001)  examined the synergistic effect of co-exposure to O3 and endotoxin on
27      the transition and respiratory epithelium of rats that also was  mediated, in part, by neutrophils.
28      Fisher 344 rats (10 to 12 week old) exposed to 0.5 ppm O3, 8 h per day, for 3 days, developed
29      mucous cell metaplasia in the nasal transitional epithelium, an area normally devoid of mucous
30      cells; whereas, intratracheal instillation of endotoxin (20 //g)  caused mucous cell metaplasia
31      rapidly in the respiratory epithelium of the conducting airways. A synergistic increase of

        April 2002                               7-101        DRAFT-DO NOT QUOTE OR CITE

-------
 1      intraepithelial mucosubstances and morphological evidence of mucous cell metaplasia were
 2      found in rat maxilloturbinates upon exposure to both ozone and endotoxin, compared to each
 3      pollutant alone.
 4           The effects of O3 modifying the biological potency of PM (diesel PM and carbon black)
 5      was examined by Madden et al. (2000). Reaction of NIST Standard Reference Material # 2975
 6      diesel PM with 0.1 ppm O3 for 48 hr increased the potency (compared to unexposed or
 7      air-exposed diesel PM) to induce neutrophil influx, total protein, and LDH in lung lavage fluid in
 8      response to intratracheal instillation. Exposure of the diesel PM to high, non-ambient O3
 9      concentration (1.0 ppm) attenuated the increased potency, suggestion destruction of the bioactive
10      reaction products.  Unlike the diesel particles, carbon black particles exposed to 0.1 ppm O3 did
11      not exhibit an increase in biological potency, which suggested that the reaction of organic
12      components of the diesel PM with O3 were responsible for the increased potency. Reaction of
13      particle components with O3 was ascertained by chemical determination of specific classes of
14      organic compounds.
15           The interaction of PM and O3 was further examined in a murine model of ovalbumin
16      (OVA)-induced asthma. Kobzik et al. (2001) investigated whether coexposure to inhaled,
17      concentrated PM from Boston, MA and to O3 could exacerbate asthma-like symptoms. On days
18      7 and 14 of life, half of the BALB/c mice used in this study were sensitized by ip injection of
19      OVA and then exposed to OVA aerosol on three successive days to create the asthma phenotype.
20      The other half received the ip OVA, but were exposed to a phosphate-buffered saline aerosol
21      (controls). The mice were further subdivided (n >6I/group) and exposed for 5 h to CAPs,
22      ranging from 63 to 1,569 //g/m3, 0.3 ppm O3, CAPs + O3, or to filtered air. Pulmonary resistance
23      and airway responsiveness to an aerosolized MCh challenge were measured after exposures. A
24      small, statistically significant increase in pulmonary resistance and airway responsiveness,
25      respectively, was found in both normal and asthmatic mice immediately after exposure to CAPs
26      alone and to CAPs + O3, but not to  O3 alone or to filtered air. By 24 h after exposure, the
27      responses returned to baseline levels.  There were no significant increases in airway
28      inflammation after any of the pollutant exposures.  In this well-designed study of a small-animal
29      model of asthma, O3 and CAPs did not appear to be synergistic. In further analysis of the data
30      using specific elemental groupings  of the CAPs, the acutely increased pulmonary resistance was
31      found to be associated withe the AISi fraction of PM.  Thus, some components of concentrated

        April 2002                                7-102       DRAFT-DO NOT QUOTE OR CITE

-------
 1      PM2 5 may affect airway caliber in sensitized animals, but the results are difficult to extrapolate to
 2      people with asthma.
 3          Linn and colleagues (1997) examined the effect of a single exposure to 60 to 140 //g/m3
 4      H2SO4, 0.1 ppm SO2, and 0.1 ppm O3 in healthy and asthmatic children. The children performed
 5      intermittent exercise during the 4-h exposure to increase the inhaled dose of the pollutants.  An
 6      overall effect on the combined group of healthy and asthmatic children was not observed. A
 7      positive association between acid concentration and symptoms was seen, however, in the
 8      subgroup of asthmatic children. The combined pollutant exposure had no effect on spirometry in
 9      asthmatic children, and no changes in symptoms or spirometry were observed in healthy children.
10      Thus, the effect of combined exposure to PM and gaseous co-pollutants appeared to have less
11      effect on asthmatic children exposed under controlled laboratory conditions in comparison with
12      field studies of children attending summer camp (Thurston et al., 1997). However, prior
13      exposure to H2SO4 aerosol may enhance the  subsequent response to O3 exposure (Linn et al.,
14      1994; Frampton et al., 1995); and the timing and sequence of the exposures may be important.
15          Six unique animal studies have examined the adverse cardiopulmonary effects of complex
16      mixtures in  urban and rural environments of Italy (Gulisano et al., 1997),  Spain (Lorz and Lopez,
17      1997), and Mexico (Vanda et al., 1998;  Calderon-Garciduefias et al., 2001c,d; Moss et al., 2001).
18      Five of these studies, identified in Table 7-12, have taken advantage of the differences in
19      pollutant mixtures of urban and rural environments to report primarily morphological changes in
20      the nasopharynx and lower respiratory tract (Gulisano et al., 1997; Lorz and Lopez, 1997;
21      Calderon-Garciduefias et al., 2001c) and in the heart (Calderon-Garciduefias et al., 2001d) of
22      lambs, pigeons, and dogs, respectively, after natural, continuous exposures to ambient pollution.
23      Each study has provided evidence that animals living in urban air pollutants have greater
24      pulmonary and cardiac changes than would occur in a rural and presumably cleaner,
25      environment.  The study by Moss et al. (2001) examined the nasal and lung tissue of rats exposed
26      (23 h/day) to Mexico City air for up to 7 weeks and compared them to controls similarly exposed
27      to filtered air.  No inflammatory or epithelial lesions were found using quantitative
28      morphological techniques; however, the concentrations of pollutants were low (see Table 7-12).
29      Extrapolation  of these results to humans is restricted, however, by uncontrolled exposure
30      conditions, small sample sizes, and other unknown exposure and nutritional factors in the studies
31      in mammals and birds, and the negative studies in rodents. They also bring up the issue of which

        April 2002                                7-103        DRAFT-DO NOT QUOTE OR CITE

-------
 1      species of "sentinel" animals is more useful for predicting urban pollutant effects in humans.
 2      Thus, in these field studies, it is difficult to assign a specific role to PM (or to any other
 3      component of the mixture) in the significant cardiopulmonary effects reported.
 4           Similar morphological changes (Calderon-Garciduefias et al., 2000a; 2001a,b) and chest
 5      X-ray evidence of mild lung hyperinflation (Calderon-Garciduefias et al., 2000b) have been
 6      reported in children residing in urban and rural areas of Mexico City.  The ambient air in urban
 7      areas, particularly in Southwest Metropolitan Mexico City (SWMMC), is a complex mixture of
 8      particles and gases, including high concentrations of O3 and aldehydes that previously have been
 9      shown to  cause airway inflammation and epithelial lesions in humans  (e.g., Calderon-
10      Garciduefias et al., 1992, 1994, 1996) and laboratory animals (Morgan et al., 1986; Heck et al.,
11      1990; Harkema et al., 1994, 1997a,b).  The described effects demonstrate a persistent, ongoing
12      upper and lower airway inflammatory process and chest X-ray abnormalities in children residing
13      predominantly in SWMMC. Again, extrapolation of these results to urban populations of the
14      United  States is difficult because of the unique complex of urban air in Mexico City,
15      uncontrolled exposure conditions, and  other unknown exposure and nutritional factors.
16           Only one controlled study has examined the effect of a combined inhalation exposure to
17      CAPs and O3 in human subjects. In a randomized, double-blind crossover study, Brook et al.
18      (2002)  exposed 25 healthy male and female subjects, 34.9 ±10 (SD) years of age, to filtered
19      ambient air containing 1.6 //g/m3 PM2 5 and 9 ppb O3 (control) or to unfiltered air containing
20      150 //g/m3 CAPs and 120 ppb O3 while at rest for 2 h. Blood pressure was measured and high-
21      resolution brachial artery ultrasonography (BAUS) was performed prior to and 10 min after
22      exposure. The BAUS technique was used to measure brachial artery diameter (BAD),
23      endothelium-dependent flow-mediated dilation  (FMD), and endothelial-independent
24      nitroglycerine-mediated dilation (NMD). Although no changes in  blood pressure or endothelial-
25      dependent or independent dilatation were observed, a small  (2.6%) but statistically significant
26      (p = 0.007) decrease in BAD was observed in CAPs plus O3 exposures (-0.09 mm) when
27      compared to filtered air exposures (+0.01 mm).  Pre-exposure BAD showed no significant day-
28      to-day variation (0.03 mm), and no significant exposure differences were found for other gaseous
29      pollutants (CO, NOX, SO2) in the ambient air. This finding suggests that combined exposure to a
30      mixture of CAPs and O3 produces vasoconstriction, potentially via autonomic reflexes or as a
31      result of an increase in circulating endothelin, as has been described in rats exposed to urban PM

        April 2002                                7-104       DRAFT-DO NOT QUOTE OR CITE

-------
 1      (Vincent et al., 2001). It is not known, however, whether this effect is caused by CAPS or O3
 2      alone, or if vasoactive responses would be found at lower PM25 and O3 concentrations typically
 3      found in most urban locations in North America.
 4           The effects of gaseous pollutants on PM-mediated responses also have been examined by in
 5      vitro studies, though to a limited extent.  Churg et al. (1996) demonstrated increased uptake of
 6      asbestos or TiO2 into rat tracheal explant cultures in response to 10 min O3 (up to 1.0 ppm) pre-
 7      exposure.  These data suggest that low concentrations O3 may increase the penetration of some
 8      types of PM into epithelial cells. Additionally, Madden et al. (2000) demonstrated a greater
 9      potency for ozonized diesel PM to induce prostaglandin E2 production from human epithelial cell
10      cultures, suggesting that O3 can modify the biological activity of PM derived from diesel exhaust.
11           No effect of NO2 exposure on PM-induced interleukin-8 production by A549 epithelial cell
12      line was found (Dick et al., 2001). The PM10 used in this study was collected from gas stoves.
13
14
15      7.7  SUMMARY
16      7.7.1 Biological Plausibility
17           Toxicological studies can play an integral role in answering the following two key
18      questions regarding biological plausibility of PM health effects.
19        (1) What component (or components) of ambient PM cause health effects?
20        (2) Are the statistical associations between PM and health effects biologically plausible?
21      This summary focuses on the progress that toxicological studies have made towards answering
22      these questions.
23
24      7.7.1.1 Link Between Specific Particulate Matter Components and Health Effects
25           Key to the validity of the biological plausibility is the need to understand the linkage
26      between the components of airborne PM responsible for the adverse effects and the individuals at
27      risk. The plausibility of the association between PM and increases in morbidity and mortality has
28      been questioned because the adverse cardiopulmonary effects have been observed at very low
29      PM concentrations, often below the current NAAQS for PM10.  To date, toxicology studies on
30      PM have provided only very limited evidence for specific PM components being responsible for

        April 2002                                7-105       DRAFT-DO NOT QUOTE OR CITE

-------
 1      observed cardiopulmonary effects of ambient PM. Studies have shown that some components of
 2      particles are more toxic than others. For example, high concentrations of ROFA and associated
 3      soluble metals have produced clinically significant effects (including death) in compromised
 4      animals. The relevance of these findings to understanding the adverse effects of PM components
 5      is tempered, however, by the large difference between metal concentrations delivered to the test
 6      animals and metal concentrations present in the ambient urban environment. Such comparisons
 7      must be applied to the interpretation of all studies that examine the individual components of
 8      ambient urban PM. A summary of potential  contributions of individual physical/chemical factors
 9      of particles to cardiopulmonary effects is given below.
10
11      Acid Aerosols
12           There is relatively little new information on the effects of acid aerosols, and the conclusions
13      of the 1996 PM AQCD are unchanged.  It was previously concluded that acid aerosols cause
14      little or no change in pulmonary function in healthy subjects, but asthmatics may develop small
15      changes in pulmonary function. This conclusion is supported by the recent study of Linn and
16      colleagues (1997) in which children (26 children with allergy or asthma and 15 healthy children)
17      were exposed to sulfuric acid aerosol (100 //g/m3) for 4 h.  There were no significant effects on
18      symptoms or pulmonary function when data from the entire group was analyzed, but the allergy
19      group had a significant increase in symptoms after the acid aerosol exposure.
20           Although pulmonary effects of acid aerosols have been the subject of extensive research in
21      past decades, the cardiovascular effects of acid aerosols have received little attention. Zhang
22      et al. (1997) reported that inhalation of acetic acid fumes caused reflex-mediated increases in
23      blood pressure in normal and spontaneously hypertensive rats. Thus, acid components  should
24      not be ruled out as possible mediators of PM health effects.  In particular, the cardiovascular
25      effects of acid aerosols at realistic concentrations need further investigation.
26
27      Metals
28           The previous PM AQCD (U.S. Environmental Protection Agency, 1996a) mainly relied on
29      data related to occupational exposures to evaluate the potential toxicity of metals in particulate
30      air pollution.  Since that time, in vivo and in  vitro studies using ROFA or soluble transition
31      metals have contributed substantial new information on the health effects of particle-associated

        April 2002                               7-106        DRAFT-DO NOT QUOTE OR CITE

-------
 1      soluble metals. Although there are some uncertainties about differential effects of one transition
 2      metal versus another, water soluble metals leached from ROFA have been shown consistently
 3      (albeit at high concentrations) to cause cell injury and inflammatory changes in vitro and in vivo.
 4          Even though it is clear that combustion particles that have a high content of soluble metals
 5      can cause lung injury and even death in compromised animals, it has not been established that the
 6      small quantities of metals associated with ambient PM are sufficient to cause health effects.
 7      Moreover, it cannot be assumed that metals are the primary toxic component of ambient PM.
 8      In studies in which various ambient and emission source particulates were instilled into  rats, the
 9      soluble metal content did appear to be the primary determinant of lung injury (Costa and Dreher,
10      1997).  However, one published study has compared the effects of inhaled ROFA (at 1 mg/m3) to
11      concentrated ambient PM (four experiments, at mean concentrations of 475 to 900 //g/m3) in
12      normal and SO2-induced bronchitic rats.  A statistically significant increase in at least one lung
13      injury marker was seen in bronchitic rats with only one out of four of the concentrated ambient
14      exposures; whereas inhaled ROFA had no effect even though the content of soluble iron,
15      vanadium, and nickel was much higher in the ROFA sample than in the concentrated ambient
16      PM.
17
18      Ultraftne Particles
19          When this subject was reviewed in the 1996 PM AQCD (U. S. Environmental Protection
20      Agency, 1996a), it was not known whether the pulmonary toxicity of freshly generated ultrafme
21      teflon particles was due to particle size or a result of absorbed fumes.  Subsequent studies with
22      other types of ultrafme particles have shown that the chemical constituents of ultrafmes
23      substantially modulate their toxicity. For example, Kuschner et al. (1997) have established that
24      inhalation of MgO particles produces far fewer respiratory effects than does ZnO.  Also,
25      inhalation exposure of normal rats to ultrafme carbon particles generated by electric arc  discharge
26      (100 //g/m3 for 6 h) caused minimal lung inflammation (Elder et al., 2000a,b), compared to
27      ultrafme Teflon or metal particles.  On the other hand, instillation of 125 //g of ultrafme carbon
28      black (20 nm) caused substantially more inflammation than  did the same dose of fine particles of
29      carbon black (200 to 250 nm), suggesting that ultrafme particles may cause more inflammation
30      than larger particles (Li et al., 1997). However, the chemical constituents of the two sizes of
31      carbon black used in this study were not analyzed, and it cannot be assumed that the chemical

        April 2002                               7-107       DRAFT-DO NOT QUOTE  OR CITE

-------
 1      composition was the same for the two sizes.  Thus, there is still insufficient toxicological
 2      evidence to conclude that ambient concentrations of ultrafme particles contribute to the health
 3      effects of paniculate air pollution. With acid aerosols, studies of low concentrations of sulfuric
 4      acid ultrafme metal oxide particles have demonstrated effects in the lung. However, it is possible
 5      that inhaled ultrafme particles may have systemic effects that are independent of effects on the
 6      lung.
 7
 8      Bioaerosols
 9           Recent studies support the conclusion of the 1996 PM AQCD (U. S. Environmental
10      Protection Agency, 1996a), which stated that bioaerosols,  at concentrations present in the
11      ambient environment, would not account for the reported health effects of ambient PM.
12      Dose-response studies in healthy volunteers exposed to 0.55 and 50 //g endotoxin, by the
13      inhalation route, showed a threshold for pulmonary and systemic effects for endotoxin between
14      0.5 and 5.0 //g (Michel et al., 1997). Monn and Becker (1999) examined effects of size
15      fractionated outdoor PM on human monocytes and found cytokine induction characteristic of
16      endotoxin activity in the coarse-size fraction but not in the fine fraction. Available information
17      suggests that ambient concentrations of endotoxin are very low and do not exceed 0.5 ng/m3.
18
19      Diesel Exhaust Particles
20           As described in Section 7.2.1.2, there is growing toxicological evidence that diesel PM
21      exacerbates the allergic response to inhaled antigens.  The organic fraction of diesel exhaust has
22      been linked to eosinophil degranulation and induction of cytokine production, suggesting that the
23      organic constituents of diesel PM are the responsible part for the immune effects. It is not known
24      whether the adjuvant-like activity of diesel PM is unique or whether other combustion particles
25      have similar effects. It is important to compare the immune effects of other source-specific
26      emissions, as well as concentrated ambient PM, to diesel PM to determine the extent to which
27      exposure to diesel exhaust may contribute to the incidence and severity of allergic rhinitis and
28      asthma.
        April 2002                                7-108        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Organic Compounds
 2           Published research on the acute effects of particle-associated organic carbon constituents is
 3      conspicuous by its relative absence, except for diesel exhaust particles. Like metals, organics are
 4      common constituents of combustion-generated particles and have been found in ambient PM
 5      samples over a wide geographical range. Organic carbon constituents comprise a substantial
 6      portion of the mass of ambient PM (10 to 60% of the total dry mass [Turpin, 1999]).  The
 7      organic fraction of ambient PM has been evaluated for its mutagenic effects. Although the
 8      organic fraction of ambient PM is a poorly characterized heterogeneous mixture of an unknown
 9      number of different compounds, strategies have been proposed for examining the health effects
10      of this potentially important constituent (Turpin, 1999).
11
12      Ambient Particle Studies
13           Ambient particle studies should be the most relevant in understanding the susceptibility of
14      individuals to PM and the underlying mechanisms.  Studies have used collected urban PM for
15      intratracheal administration to healthy and compromised animals.  Despite the difficulties in
16      extrapolating from the bolus delivery used in such studies, they have provided strong evidence
17      that the chemical composition of ambient particles can have a major influence on toxicity.  More
18      recent work with inhaled concentrated ambient PM has observed cardiopulmonary changes in
19      rodents and dogs at high concentrations of fine PM. No comparative studies to examine the
20      effects of ultrafine and coarse ambient PM have been done, although a new ambient particle
21      concentrator developed by Sioutas and colleagues should permit the direct toxicological
22      comparison of various ambient particle  sizes. Importantly, it has become evident that, although
23      the concentrated ambient PM studies can provide important dose-response information, identify
24      susceptibility factors in animal models,  and permit examination of mechanisms related to PM
25      toxicity, they are not particularly well suited for the identification of toxic components in urban
26      PM.  Because only a limited number of exposures using concentrated ambient PM can be
27      reasonably conducted by a given laboratory in a particular urban environment, there may be
28      insufficient information to conduct a factor analysis on an exposure/response matrix. This may
29      also hinder principal component analysis techniques that are useful in identifying particle
30      components responsible for adverse outcomes.
31

        April 2002                                7-109        DRAFT-DO NOT QUOTE OR CITE

-------
 1      7.7.1.2 Susceptibility
 2           Progress has been made in understanding the role of individual susceptibility to ambient
 3      PM effects. Studies have consistently shown that older animals or animals with certain types of
 4      compromised health, either genetic or induced, are more susceptible to instilled or inhaled
 5      particles, although the increased animal-to-animal variability in these models has created
 6      problems.  Moreover, because PM seems to affect broad categories of disease states, ranging
 7      from cardiac arrhythmias to pulmonary infection, it can be difficult to know what disease models
 8      to use in understanding the biological plausibility of the adverse health effects of PM.  Thus, the
 9      identification of susceptible animal models has been somewhat slow, but overall it represents
10      solid progress when one considers that data from millions of people are necessary in
11      epidemiology studies to develop the statistical power to detect small increases in PM-related
12      morbidity and mortality.
13
14      7.7.2 Mechanisms of Action
15           The mechanisms that underlie the biological responses to ambient PM are not clear.
16      Various toxicologic studies using particulate matter having diverse physicochemical
17      characteristics have shown that these characteristics have a great impact on the specific response
18      that is observed.  Thus, there are multiple biological mechanisms that may be responsible for
19      observed morbidity/mortality due to exposure to ambient PM, and these mechanisms may be
20      highly dependent on the type of particle in the exposure atmosphere. However, it should be
21      noted that many controlled exposure studies used particle concentrations much higher than those
22      typically occurring in ambient air. Thus, some of the mechanisms elicited may not occur with
23      exposure to lower levels.  Clearly, controlled exposure studies have not as yet been able to
24      unequivocally determine the particle characteristics and the toxicological mechanisms by which
25      ambient PM may affect biological systems.
26
        April 2002                                7-110       DRAFT-DO NOT QUOTE OR CITE

-------
  1       REFERENCES

  2       Adamson, I. Y. R.; Prieditis, H.; Hedgecock, C.; Vincent. R. (2000) Zinc is the toxic factor in the lung response to
  3             an atmospheric paniculate sample. Toxicol. Appl. Pharmacol. 166: 111-119.
  4       Amdur, M. O.; Chen, L. C. (1989) Furnace-generated acid aerosols: speciation and pulmonary effects.
  5             In: Symposium on the health effects of acid aerosols; October 1987; Research Triangle Park, NC. Environ.
  6             Health Perspect. 79: 147-150.
  7       Amdur, M. O.; Dubriel, M.; Creasia, D. A. (1978) Respiratory response of guinea pigs to low levels of sulfuric acid.
  8             Environ. Res. 15: 418-423.
  9       Ball, B. R.; Smith, K. R.; Veranth, J. M.; Aust, A. E. (2000) Bioavailability of iron from coal fly ash: mechanisms
10             of mobilization and of biological effects. In: Grant, L. D., ed. PM2000: paniculate matter and health.
11             Inhalation Toxicol. 12(suppl. 4): 209-225.
12       Bayram, H.; Devalia, J. L.; Khair, O. A.; Abdelaziz, M. M.; Sapsford, R. J.; Sagai, M.; Davies, R. J. (1998a)
13             Comparison of ciliary activity and inflammatory mediator release from bronchial epithelial cells of nonatopic
14             nonasthmatic subjects and atopic asthmatic patients and the effect of diesel exhaust particles in vitro.
15             J. Allergy Clin. Immunol.  102: 771-782.
16       Bayram, H.; Devalia, J. L.; Sapsford, R. J.; Ohtoshi, T.; Miyabara, Y.; Sagai, M.; Davies, R. J. (1998b) The effect
17             of diesel exhaust particles on cell function and release of inflammatory mediators from human bronchial
18             epithelial cells in vitro. Am. J. Respir. Cell Mol. Biol. 18: 441-448.
19       Becker, S.; Soukup, J. M. (1998) Decreased CD 1 IB expression, phagocytosis, and oxidative burst in urban
20             paniculate pollution-exposed human monocytes and alveolar macrophages. J. Toxicol. Environ. Health Part
21             A 55: 455-477.
22       Becker, S.; Soukup, J. M.; Gilmour, M. L; Devlin, R.  B. (1996) Stimulation of human and rat alveolar macrophages
23             by urban air particulates: effects on oxidant radical generation and cytokine production. Toxicol. Appl.
24             Pharmacol. 141:637-648.
25       Blomberg, A.; Sainsbury, C.; Rudell, B.; Frew, A. J.; Holgate, S. T.; Sandstrom, T.; Kelly, F. J. (1998) Nasal cavity
26             lining fluid ascorbic acid concentration increases in healthy human volunteers following short term exposure
27             to diesel exhaust. Free Radical Res. 28: 59-67.
28       Bolarin, D. M.; Bhalla, D. K.; Kleinman, M. T.  (1997) Effects of repeated exposures of geriatric rats to ozone and
29             particle-containing atmospheres: an analysis of bronchoalveolar lavage and plasma proteins. Inhalation
30             Toxicol. 9: 423-434.
31       Bonner, J. C.; Rice, A. B.; Lindroos, P. M.; O'Brien, P. O.; Dreher, K. L.; Rosas, L; Alfaro-Moreno, E.;
32             Osornio-Vargas, A. R. (1998) Induction of the lung myofibroblast PDGF receptor system by urban ambient
33             particles from Mexico City. Am. J. Respir. Cell Mol. Biol. 19: 672-680.
34       Bouthillier, L.; Vincent, R.; Goegan, P.; Adamson, I. Y. R.; Bjarnason, S.; Stewart, M.; Guenette, J.; Potvin, M.;
3 5             Kumarathasan, P. (1998) Acute effects of inhaled urban particles and ozone: lung morphology, macrophage
36             activity, and plasma endothelin-1. Am. J. Pathol. 153: 1873-1884.
37       Brain, J. D.; Long, N. C.; Wolfthal, S. F.; Dumyahn, T.; Dockery, D. W. (1998) Pulmonary toxicity in hamsters of
38             smoke particles from Kuwaiti oil fires. Environ. Health Perspect. 106: 141-146.
39       Broeckaert, F.; Buchet, J. P.; Huaux, F.; Lardot, C.; Lison, D.; Yager, J. W. (1997) Reduction of the ex vivo
40             production of tumor necrosis factor alpha by alveolar phagocytes after administration of coal fly ash and
41             copper smelter dust. J. Toxicol. Environ. Health 51:  189-202.
42       Brook, R. D.; Brook, J. R.; Urch, B.; Vincent, R.; Rajagopalan, S.;  Silverman, F. (2002) Inhalation of fine
43             paniculate air pollution and ozone causes acute arterial vasoconstriction in health adults. Circulation 105:
44             1534-1536.
45       Calderon-Garciduenas, L.; Osorno-Velaquez, A.; Bravo-Alvarez, H.; Delgado-Chavez, R.; Barrios-Marquez, R.
46             (1992) Histopathologic changes of the nasal mucosa in southwest metropolitan  Mexico City inhabitants. Am.
47             J. Pathol. 140: 225-232.
48       Calderon-Garciduenas, L.; Rodriguez-Alcaraz, A.; Garcia, R.;  Sanchez, G.; Barragan,  G.; Camacho, R.; Ramirez, L.
49             (1994) Human nasal mucosal changes after exposure to urban pollution. Environ. Health Perspect. 102:
50             1074-1080.
51       Calderon-Garciduenas, L.; Rodriguez-Alcaraz, A.; Garcia, R.;  Ramirez, L.; Barragan,  G. (1995) Nasal inflammatory
52             responses in children exposed to a polluted urban atmosphere. J. Toxicol. Environ. Health 45: 427-437.
         April 2002                                      7-111         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Calderon-Garciduenas, L.; Osnaya-Brizuela, N.; Ramirez-Martinez, L.; Villarreal-Calderon, A. (1996) DNA strand
  2             breaks in human nasal respiratory epithelium are induced upon exposure to urban pollution. Environ. Health
  3             Perspect. 104: 160-168.
  4       Calderon-Garciduenas, L.; Devlin, R. B.; Miller, F. J. (2000a) Respiratory tract pathology and cytokine imbalance
  5             in clinically healthy children chronically and sequentially exposed to air pollutants. Med. Hypotheses 55:
  6             373-378.
  7       Calderon-Garciduenas, L.; Mora-Tascareno, A.; Chung, C. J.; Valencia, G.; Fordham, L. A.; Garcia, R.; Osnaya,
  8             N.; Romero, L.; Acuna, H.; Villarreal-Calderon, A.; Devlin, R. B.; Koren, H. S. (2000b) Exposure to air
  9             pollution is associated with lung hyperinflation in healthy children and adolescents in southwest Mexico
10             City: a pilot study. Inhalation Toxicol. 12: 537-561.
11       Calderon-Garciduenas, L.; Valencia-Salazar, G.; Rodriguez-Alcaraz,  A.; Gambling, T. M.; Garcia, R.; Osnaya, N.;
12             Villarreal-Calderon, A.; Devlin, R. B.; Carson, J. L. (2001a) Ultrastructural nasal pathology in children
13             chronically and sequentially exposed to air pollutants. Am. J. Respir. Cell Mol. Biol. 24: 132-138.
14       Calderon-Garciduenas, L.; Rodriguez-Alcaraz, A.; Valencia-Salazar,  G.; Mora-Tascareno, A.; Garcia, R.; Osnaya,
15             N.; Villarreal-Calderon, A.; Devlin, R. B.; Van Dyke, T. L. (200Ib) Nasal biopsies of children exposed to air
16             pollutants. Toxicol. Pathol. 29: 558-564.
17       Calderon-Garciduenas, L.; Mora-Tiscareno, A.; Fordham, L. A.; Chung, C. J.; Garcia, R.; Osnaya, N.; Hernandez,
18             J.; Acuna, H.; Gambling, T. M.; Villarreal-Calderon, A.; Carson, J.; Koren, H. S.; Devlin, R. B. (200Ic)
19             Canines as sentinel species for assessing chronic exposures to  air pollutants: part 1. respiratory pathology.
20             Toxicol. Sci.  61:342-355.
21       Calderon-Garciduenas, L.; Gambling, T. M.; Acuna, H.; Garcia, R.; Osnaya, N.; Monroy, S.; Villarreal-Calderon,
22             A.; Carson, J.; Koren, H. S.; Devlin, R. B. (2001d) Canines as sentinel species for assessing chronic
23             exposures to air pollutants: part 2.  cardiac pathology. Toxicol.  Sci. 61: 356-367.
24       Campen, M. J.; Costa, D. L.; Watkinson,  W. P. (2000) Cardiac and thermoregulatory toxicity of residual oil fly ash
25             in cardiopurmonary-compromised rats. In:  Phalen, R. F., ed. Inhalation Toxicology: proceedings of the third
26             colloquium on paniculate air pollution and human health (second special issue); June, 1999; Durham, NC.
27             Inhalation Toxicol. 12(suppl. 2): 7-22.
28       Campen, M. J.; Nolan, J. P.; Schladweiler, M. C. J.; Kodavanti, U. P.; Evansky, P. A.; Costa, D. L.; Watkinson, W.
29             P. (2001) Cardiovascular and thermoregulatory effects of inhaled PM-associated transition metals: a potential
30             interaction between nickel and vanadium surf ate. Toxicol. Sci. 64: 243-252.
31       Campen, M. J.; Nolan, J. P.; Schladweiler, M. C. J.; Kodavanti, U. P.; Costa, D. L.; Watkinson, W. P. (2002)
32             Cardiac and thermoregulatory effects of instilled paniculate matter-associated transition metals in healthy and
33             cardiopurmonary-compromised rats. J. Toxicol. Environ. Health (special issue): in press.
34       Carter, J. D.; Ohio, A. J.; Samet, J. M.; Devlin, R. B. (1997) Cytokine production by human airway epithelial cells
3 5             after exposure to an air pollution particle is metal-dependent. Toxicol.  Appl. Pharmacol. 146:  180-188.
36       Castellan, R. M.; Olenchock, S. A.; Hankinson, J. L.; Millner, P. D.;  Cocke, J. B.; Bragg, C. K.; Perkins, H. H., Jr.;
37             Jacobs, R. R. (1984) Acute bronchoconstriction induced by cotton dust: dose-related responses to endotoxin
38             and other dust factors. Ann. Intern. Med. 102: 157-163.
39       Castellan, R. M.; Olenchock, S. A.; Kinsley, K. B.; Hankinson, J. L. (1987) Inhaled endotoxin and decreased
40             spirometric values: an exposure-response relation for cotton dust. N. Engl. J. Med. 317: 605-610.
41       Chang, M. C.; Sioutas, C.; Kim, S.; Gong, H., Jr.; Linn, W. S. (2000) Reduction of nitrate losses from filter and
42             impactor samplers by means of concentration enrichment. Atmos. Environ. 34:  85-98.
43       Chen, L. C.; Fine, J. M.; Qu, Q.-S.; Amdur, M. O.; Gordon, T. (1992) Effects of fine and ultrafine sulfuric acid
44             aerosols in guinea pigs: alterations in alveolar macrophage function and intracellular pH. Toxicol. Appl.
45             Pharmacol. 113: 109-117.
46       Chen, L. C.; Fang, C. P.; Qu, Q. S.; Fine, J. M.; Schlesinger, R. B. (1993) A novel system for the in vitro exposure
47             of pulmonary cells to acid  sulfate aerosols. Fundam. Appl. Toxicol. 20: 170-176.
48       Churg, A.; Brauer, M.; Keeling, B. (1996) Ozone enhances the uptake of mineral particles by tracheobronchial
49             epithelial cells in organ culture. Am. J. Respir. Crit. Care Med. 153: 1230-1233.
50       Churg, A.; Stevens, B.; Wright, J. L.  (1998) Comparison of the uptake of fine and ultrafine TiO2 in a tracheal
51             explant system. Am. J. Physiol. 274: L81-L86.
52       Clarke, R. W.; Catalano, P. J.; Koutrakis, P.; Krishna Murthy, G. G.;  Sioutas, C.; Paulauskis, J.; Coull, B.;
53             Ferguson, S.; Godleski, J. J. (1999) Urban air paniculate inhalation alters pulmonary function and induces
54             pulmonary inflammation in a rodent model of chronic bronchitis. Inhalation Toxicol. 11: 637-656.
         April 2002                                     7-112         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Clarke, R. W.; Coull, B.; Reinisch, U.; Catalano, P.; Killingsworth, C. R.; Koutrakis, P.; Kavouras, I.; Murthy, G.
  2             G. K.; Lawrence, J.; Lovett, E.; Wolf son, J. M; Verrier, R. L.; Godleski, J. J. (2000a) Inhaled concentrated
  3             ambient particles are associated with hematologic and bronchoalveolar lavage changes in canines. Environ.
  4             Health Perspect. 108: 1179-1187.
  5       Clarke, R. W.; Catalano, P.; Coull, B.; Koutrakis, P.; Krishna Murthy, G. G.; Rice, T.; Godleski, J. J. (2000)
  6             Age-related responses in rats to concentrated urban air particles (CAPs). In: Phalen, R. F., ed. Inhalation
  7             Toxicology: proceedings of the third colloquium on paniculate air pollution and human health (first special
  8             issue); June, 1999; Durham, NC. Inhalation Toxicol.  12(suppl. 1): 73-84.
  9       Cohen, M. D.; Becker, S.; Devlin, R.; Schlesinger, R. B.; Zelikoff, J. T. (1997) Effects of vanadium upon
10             polyl:C-induced responses in rat lung and alveolar macrophages. J. Toxicol. Environ. Health 51: 591-608.
11       Cormier, Y.; Laviolette, M.; Bedard, G.; Dosman, J.; Israel-Assayag, E. (1998) Effect of route of breathing on
12             response to exposure ina swine confinement building. Am. J. Respir. Crit. Care Med. 157:  1512-1521.
13       Costa, D. L.; Dreher, K. L. (1997) Bioavailable transition metals in paniculate matter mediate cardiopulmonary
14             injury in healthy and compromised animal models. In: Driscoll, K. E.; Oberdorster, G., eds. Proceedings of
15             the sixth international meeting on the toxicology of natural and man-made fibrous and non-fibrous particles;
16             September 1996; Lake Placid, NY. Environ. Health Perspect. Suppl. 105(5): 1053-1060.
17       Creutzenberg, O.; Bellmann, B.; Muhle, H.; Dasenbrock, C.; Morrow, P.; Mermelstein, R. (1998) Lung clearance
18             and retention of toner, TiO2, and crystalline silica, utilizing a tracer technique during chronic inhalation
19             exposure in Syrian Golden hamsters. Inhalation Toxicol. 10:731-751.
20       Diaz-Sanchez, D.; Tsien, A.;  Casillas, A.; Dotson, A. R.; Saxon, A. (1996) Enhanced nasal cytokine production in
21             human beings after in vivo challenge with diesel exhaust particles. J. Allergy Clin. Immunol. 98: 114-123.
22       Diaz-Sanchez, D.; Tsien, A.; Fleming, J.; Saxon, A. (1997) Combined diesel exhaust paniculate and ragweed
23             allergen challenge markedly enhances human in vivo  nasal ragweed-specific IgE and skews cytokine
24             production to a T helper cell 2-type pattern. J. Immunol. 158: 2406-2413.
25       Dick, C. A. J.; Dennekamp, M.; Howarth, S.; Cherrie, J. W.; Seaton, A.; Donaldson, K.; Stone, V.; (2001)
26             Stimulation of IL-8 release from epithelial cells by gas cooker PM10: a pilot study. Occup. Environ. Med. 58:
27             208-210.
28       Diociaiuti, M.; Balduzzi, M.; De Berardis, B.; Cattani, G.; Stacchini, G.; Ziemacki, G.; Marconi, A.; Paoletti, L.
29             (2001) The two PM2 5 (fine) and PM2 5-10 (coarse) fractions: evidence of different biological activity. Environ.
30             Res. A 86: 254-262.
31       Donaldson, K.; Brown, D. M.; Mitchell, C.; Dineva, M.; Beswick, P. H.; Gilmour, P.; MacNee,  W. (1997) Free
32             radical activity of PM10: iron-mediated generation of hydroxyl radicals. In: Driscoll, K. E.; Oberdorster, G.,
3 3             eds. Proceedings of the sixth international meeting on the toxicology of natural and man-made fibrous and
34             non-fibrous particles; September 1996; Lake Placid, NY. Environ. Health Perspect. Suppl. 105(5):
35             1285-1289.
36       Dong, W.; Lewtas, J.; Luster, M. I. (1996) Role of endotoxin in tumor necrosis factor aexpression from alveolar
37             macrophages treated with urban air particles. Exp. Lung Res. 22: 577-592.
38       Dormans, J. A. M. A.; Steerenberg, P. A.; Arts, J. H. E.; VanBree, L.; De Klerk, A.; Verlaan, A. P. J.; Bruijntjes,
39             J. P.; Beekhof, P.; Van Soolingen, D.; Van Loveren, H. (1999) Pathological and immunological effects of
40             respirable coal fly ash in male wistar rats. Inhalation Toxicol. 11:  51-69.
41       Dreher, K. L.; Jaskot, R. H.; Lehmann, J. R.; Richards, J. H.; McGee, J. K.; Ohio, A. J.; Costa, D.  L. (1997)
42             Soluable transition metals mediate residual oil fly ash induced acute lung injury. J. Toxicol. Environ. Health
43             50:285-305.
44       Driscoll, K. E.; Costa, D. L.; Hatch, G.; Henderson, R.; Oberdorster, G.; Salem, H.; Schlesinger, R. B. (2000)
45             Intratracheal instillation as an exposure technique for the evaluation of respiratory tract toxicity: uses and
46             limitations. Toxicol. Sci. 55:  24-35.
47       Dye, J. A.; Adler, K. B.; Richards, J. H.; Dreher, K. L. (1997) Epithelial injury induced by exposure to residual oil
48             fly-ash particles: role of reactive oxygen species? Am. J. Respir. Cell Mol. Biol. 17: 625-633.
49       Dye, J. A.; Adler, K. B.; Richards, J. H.; Dreher, K. L. (1999) Role of soluble metals in oil fly ash-induced airway
50             epithelial injury and cytokine gene expression. Am. J. Physiol. 277: L498-L510.
51       Dye, J. A.; Lehmann, J. R.; McGee, J. K.; Winsett, D. W.; Ledbetter, A.  D.; Everitt, J. I.; Ohio, A. J.; Costa, D. L.
52             (2001) Acute pulmonary toxicity of paniculate matter filter extracts in rats: coherence with epidemiological
53             studies in Utah Valley residents. Environ. Health Perspect. 109(suppl. 3): 395-403.
54       Elder, A. C. P.; Gelein, R.; Finkelstein J. N.; Cox,  C.; Oberdorster, G. (2000a) Endotoxin priming affects the lung
55             response to ultrafine particles and ozone in young and old rats. In: Phalen, R. F., ed. Inhalation Toxicology:


         April 2002                                     7-113         DRAFT-DO NOT QUOTE OR CITE

-------
  1             proceedings of the third colloquium on paniculate air pollution and human health (first special issue); June,
  2             1999; Durham, NC. Inhalation Toxicol. 12(suppl.  1): 85-98.
  3       Elder, A. C. P.; Gelein, R.; Finkelstein, J. N.;  Cox, C.; Oberdorster, G. (2000b) Pulmonary inflammatory response
  4             to inhaled ultrafine particles is modified by age, ozone exposure, and bacterial toxin. In: Grant, L. D., ed.
  5             PM2000: paniculate matter and health. Inhalation Toxicol. 12(suppl. 4): 227-246.
  6       Fabiani, R.; Pampanella, L.; Minelli, A.; Mezzasoma, L; Vecchiarelli, A.; Morozzi, G. (1997) Effect of airborne
  7             paniculate extracts onmonocyte oxidative metabolism. J. Environ. Pathol. Toxicol. Oncol. 16: 195-199.
  8       Ferin, J.; Oberdorster, G.; Penney, D. P.; Soderholm, S. C.; Gelein, R.; Piper, H. C. (1990) Increased pulmonary
  9             toxicity of ultrafine particles? I. Particle clearance, translocation, morphology. J. Aerosol Sci. 21: 381-384.
10       Ferin, J.; Oberdorster, G.; Penney, D. P. (1992) Pulmonary retention of ultrafine and fine particles in rats. Am. J.
11             Respir. Cell Mol. Biol. 6: 535-542.
12       Finkelstein, J. N.; Johnston, C.; Barrett, T.; Oberdorster, G. (1997) Particulate-cell interactions and pulmonary
13             cytokine expression. In: Driscoll, K. E.; Oberdorster, G., eds. Proceedings of the sixth international meeting
14             on the toxicology of natural and man-made fibrous and non-fibrous particles; September 1996; Lake Placid,
15             NY. Environ. Health Perspect. Suppl. 105(5): 1179-1182.
16       Folinsbee, L. J.; Kim, C. S.; Kehrl, H. R.; Prah, J. D.; Devlin, R. B. (1997) Methods in human inhalation toxicology.
17             In: Massaro, E. J., ed. Handbook of human toxicology. Boca Raton, FL: CRC Press LLC; pp. 607-670.
18       Fortoul, T. L; Osorio, L.  S.; Tovar, A. T.; Salazar, D.; Castilla, M. E.; Olaiz-Fernandez, G. (1996) Metals in lung
19             tissue from autopsy cases in Mexico City residents: comparison of cases from the 1950s and the 1980s.
20             Environ. Health Perspect. 104: 630-632.
21       Frampton, M. W.; Voter, K. Z.; Morrow, P. E.; Roberts, N. J., Jr.; Culp, D. J.; Cox, C.; Utell, M. J. (1992) Sulfuric
22             acid aerosol exposure in humans assessed by bronchoalveolar lavage. Am. Rev. Respir. Dis. 146: 626-632.
23       Frampton, M. W.; Morrow, P. E.; Cox, C.; Levy, P. C.; Condemi, J. J.; Speers, D.; Gibb, F. R.; Utell, M. J.  (1995)
24             Sulfuric acid aerosol followed by ozone exposure in healthy and asthmatic subjects. Environ. Res. 69: 1-14.
25       Frampton, M. W.; Ohio, A. J.; Samet, J. M.; Carson, J. L.; Carter, J. D.; Devlin, R. B. (1999) Effects of aqueous
26             extracts of PM10 filters from the  Utah Valley on human airway epithelial cells. Am. J. Physiol.
27             277: L960-L967.
28       Fujimaki, H.; Katayama, N.; Wakamori, K. (1992) Enhanced histamine release from lung mast cells of guinea pigs
29             exposed to sulfuric acid aerosols. Environ. Res. 58: 117-123.
30       Gardner, S. Y.; Lehmann, J. R.; Costa,  D. L. (2000) Oil fly ash-induced elevation of plasma fibrinogen levels in
31             rats. Toxicol. Sci.  56: 175-180.
32       Gavett, S. H.; Madison, S. L.; Dreher, K. L.; Winsett, D. W.; McGee, J. K.; Costa, D. L. (1997) Metal and sulfate
33             composition of residual oil fly ash determines airway hyperreactivity and lung injury in rats. Environ. Res.
34             72: 162-172.
35       Gavett, S. H.; Madison, S. L.; Stevens,  M. A.; Costa, D. L. (1999) Residual oil fly ash amplifies allergic cytokines,
36             airway responsiveness, and inflammation in mice.  Am. J. Respir. Crit. Care Med. 160: 1897-1904.
37       Gearhart, J. M.; Schlesinger, R. B. (1986) Sulfuric acid-induced airway hyperresponsiveness. Fundam. Appl.
38             Toxicol. 7: 681-689.
39       Gercken, G.; Berg, I.; Dorger, M.; Schluter, T. (1996) Mechanisms of particle-induced activation of alveolar
40             macrophages. Toxicol. Lett. 88:  121-129.
41       Ohio, A. J.; Devlin, R. B. (2001) Inflammatory lung injury after bronchial instillation of air pollution particles. Am.
42             J. Respir. Crit. Care Med. 164: 704-708.
43       Ohio, A. J.; Meng, Z. H.; Hatch, G.  E.; Costa, D. L. (1997a) Luminol-enhanced chemiluminescence after in vitro
44             exposures of rat alveolar macrophages to  oil fly ash is metal dependent. Inhalation Toxicol. 9: 255-271.
45       Ohio, A. J.; Piantadosi, C. A.; Crumbliss, A. L. (1997b) Hypothesis: iron chelation plays a vital role in neutrophilic
46             inflammation.  BioMetals  10: 135-142.
47       Ohio, A. J.; Carter, J. D.; Richards,  J. H.; Brighton, L. E.; Lay, J. C.; Devlin, R. B. (1998a) Disruption of normal
48             iron homeostasis after bronchial instillation of an iron-containing particle. Am. J. Physiol. 274: L396-L403.
49       Ohio, A. J.; Richards, J. H.; Dittrich, K. L.; Samet, J. M.  (1998b) Metal storage and transport proteins increase after
50             exposure of the rat lung to an air pollution particle. Toxicol. Pathol.  26: 388-394.
51       Ohio, A. J.; Carter, J. D.; Samet, J. M.; Reed,  W.; Quay, J.; Dailey, L. A.; Richards, J. H.; Devlin, R. B. (1998c)
52             Metal-dependent expression of ferritin and lactoferrin by respiratory epithelial cells. Am. J. Physiol.
53             274: L728-L736.
         April 2002                                     7-114         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Ohio, A. J.; Stonehuerner, J.; Dailey, L. A.; Carter, J. D. (1999a) Metals associated with both the water-soluble and
  2             insoluble fractions of an ambient air pollution particle catalyze an oxidative stress. Inhalation Toxicol. 11: 37-49.
  3       Ohio, A. J.; Carter, J. D.; Dailey, L. A.; Devlin, R. B.; Samet, J. M. (1999b) Respiratory epithelial cells demonstrate
  4             lactoferrin receptors that increase after metal exposure. Am. J. Physiol. 276: L933-L940.
  5       Ohio, A. J.; Stoneheurner, J.; McGee, J. K.; Kinsey, J. S. (1999c) Sulfate content correlates with iron concentrations
  6             in ambient air pollution particles. Inhalation Toxicol. 11: 293-307.
  7       Ohio, A. J.; Kim, C.; Devlin, R. B. (2000a) Concentrated ambient air particles induce mild pulmonary inflammation
  8             in healthy human volunteers. Am. J. Respir. Crit. Care Med. 162: 981-988.
  9       Ohio, A. J.; Carter, J. D.; Richards, J. H.; Crissman, K. M.; Bobb,  H. H.; Yang, F. (2000b) Diminished injury in
10             hypotransferrinemic mice after exposure to a metal-rich particle. Am. J. Physiol. 278: L1051-L1061.
11       Ohio, A. J.; Gilbey, J. G.; Roggli, V. L.; Richards, J. H.; McGee, J. K.; Carson, J. L.; Devlin, R. B.; Cascio, W. E.
12             (2001) Diffuse alveolar damage after exposure to an oil fly ash. Am. J. Respir. Crit. Care Med.  164:
13             1514-1518.
14       Gift, J. S.; Faust, R. A. (1997) Noncancer inhalation toxicology of crystalline silica: exposure-response assessment.
15             J. Exposure Anal. Environ. Epidemiol. 7: 345-358.
16       Gilmour, M. L; Taylor, F. G. R.; Baskerville, A.; Wathes, C. M. (1989a) The effect of titanium dioxide inhalation
17             on the pulmonary clearance of Pasteurella haemolytica in the mouse. Environ. Res. 50: 157-172.
18       Gilmour, M. L; Taylor, F. G. R.; Wathes, C. M. (1989b) Pulmonary clearance of Pasteurella haemolytica and
19             immune responses in mice following exposure to titanium dioxide. Environ. Res. 50: 184-194.
20       Gilmour, P. S.; Brown, D. M.; Lindsay, T. G.; Beswick, P. H.; MacNee, W.; Donaldson, K. (1996) Adverse health
21             effects of PM10 particles: involvement of iron in generation of hydroxyl radical. Occup. Environ. Med.
22             53:817-822.
23       Godleski, J. J.; Verrier, R. L.; Koutrakis, P.; Catalano, P. (2000) Mechanisms of morbidity and mortality from
24             exposure to ambient air particles. Cambridge, MA: Health Effects Institute; research report no.  91.
25       Goldsmith, C.-A.; Frevert, C.; Imrich, A.; Sioutas, C.; Kobzik, L. (1997) Alveolar macrophage interaction with air
26             pollution particulates. In: Driscoll, K. E.; Oberdorster, G., eds. Proceedings of the sixth international meeting
27             on the toxicology of natural and man-made fibrous and non-fibrous particles; September 1996;  Lake Placid,
28             NY. Environ. Health Perspect. Suppl. 105(5): 1191-1195.
29       Goldsmith, C.-A. W.; Imrich, A.; Danaee, H.; Ning, Y. Y.; Kobzik, L. (1998) Analysis of air pollution
30             particulate-mediated oxidant stress in alveolar macrophages. J. Toxicol. Environ. Health Part A 54: 529-545.
31       Goldsmith, C.-A. W.; Hamada, K.; Ning, Y. Y.; Qin, G.; Catalano, P.; Murthy, G. G. K.; Lawrence, J.; Kobzik, L.
32             (1999) The effects of environmental aerosols on airway hyperresponsiveness in a murine model of asthma.
33             Inhalation Toxicol. 11: 981-998.
34       Gong, H., Jr.; Sioutas, C.; Linn, W. S.; Clark, K. W.; Terrell, S. L.; Terrell, L. L.; Anderson, K. R.; Kim, S.; Chang,
3 5             M.-C. (2000) Controlled human exposures to concentrated ambient fine particles in metropolitan Los
36             Angeles: methodology and preliminary health-effect findings. In: Phalen, R. F., ed. Inhalation Toxicology:
3 7             proceedings  of the third colloquium on paniculate air pollution and human health (first special issue); June,
38             1999; Durham, NC. Inhalation Toxicol. 12(suppl.  1): 107-119.
39       Gordon, T.; Nadziejko, C.; Schlesinger, R.; Chen, L. C. (1998) Pulmonary and cardiovascular effects of acute
40             exposure to concentrated ambient paniculate matter in rats.  Toxicol. Lett. 96-97: 285-288.
41       Gordon, T.; Nadziejko, C.; Chen, L. C.; Schlesinger, R. (2000) Effects of concentrated ambient particles in rats and
42             hamsters: an exploratory study. Cambridge, MA: Health Effects Institute; research report no. 93.
43       Grabowski, G. M.;  Paulauskis, J. D.; Godleski, J. J. (1999) Mediating phosphorylation events in the
44             vanadium-induced respiratory burst of alveolar macrophages. Toxicol. Appl. Pharmacol. 156: 170-178.
45       Gulisano, M.; Marceddu, S.; Barbara, A.; Pacini, A.; Buiatti, E.; Martini, A.; Pacini, P. (1997) Damage to the
46             nasopharyngeal mucosa induced by current levels of urban air pollution: a field study in lambs.  Eur. Respir.
47             J. 10:  567-572.
48       Hahon, N.; Booth, J. A.; Green, F.; Lewis, T. R. (1985) Influenza virus infection in mice after exposure to coal dust
49             and diesel engine  emissions. Environ. Res. 37: 44-60.
50       Hamada, K.; Goldsmith, C.-A.; Kobzik, L. (1999) Increased airway hyperresponsiveness and inflammation in a
51             juvenile mouse model of asthma exposed to air-pollutant aerosol. J. Toxicol. Environ. Health 58: 129-143.
52       Hamada, K.; Goldsmith, C.-A.; Goldman, A.; Kobzik, L. (2000) Resistance of very young mice to inhaled allergen
53             sensitization Is overcome by coexposure to an air-pollutant aerosol. Am. J. Respir. Crit. Care Med.
54             161: 1285-1293.
         April 2002                                     7-115         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Harder, S. D.; Soukup, J. M; Ohio, A. J.; Devlin, R. B.; Becker, S. (2001) Inhalation of PM25 does not modulate
  2             host defense or immune parameters in blood or lung of normal human subjects. Environ. Health Perspect.
  3             109(suppl. 4): 599-604.
  4       Harkema, J. R.; Morgan, K. T.; Gross, E. A.; Catalano, P. J.; Griffith, W. C. (1994) Consequences of prolonged
  5             inhalation of ozone on F344/N rats: collaborative studies. Part VII: effects on the nasal mucociliary
  6             apparatus. Cambridge, MA: Health Effects Institute; research report no. 65.
  7       Harkema, J. R.; Barr, E. B.; Hotchkiss, J. A. (1997a) Responses of rat nasal epithelium to short- and long-term
  8             exposures of ozone: image analysis of epithelial injury, adaptation and repair. Microsc. Res. Tech. 36:
  9             276-286.
10       Harkema, J. R.; Hotchkiss, J. A.; Griffith, W. C. (1997b) Mucous cell metaplasia in rat nasal epithelium after a
11             20-month exposure to ozone: a morphometric study of epithelial differentiation. Am. J. Respir.  Cell Mol.
12             Biol. 16: 521-530.
13       Health Effects Institute. (1995) Diesel exhaust: a critical analysis of emissions, exposure, and health effects:
14             a special report of the Institute's Diesel Working Group. Cambridge, MA: Health Effects Institute.
15       Heck, H.  d'A.; Casanova, M.; Starr, T. B. (1990) Formaldehyde toxicity~new understanding. Crit. Rev. Toxicol. 20:
16             397-426.
17       Heyder, J.; Beck-Speier, I.; Busch, B.; Dirscherl, P.; Heilmann, P.; Perron, G. A.; Josten, M.; Karg, E.; Kreyling,
18             W. G.; Lenz, A.-G.; Maier, K. L.; Miaskowski, U.; Platz, S.; Reitmeir, P.; Schulz, H.; Takenaka, S.;
19             Ziesenis, A. (1999) Health effects of sulfur-related environmental air pollution. I. Executive summary.
20             Inhalation Toxicol. 11: 343-359.
21       Hitzfeld,  B.; Friedrichs, K. H.; Ring, J.; Behrendt, H. (1997) Airborne paniculate matter modulates the production
22             of reactive oxygen species in human polymorphonuclear granulocytes. Toxicology 120: 185-195.
23       Holian, A.; Hamilton, R. F., Jr.; Morandi, M. T.; Brown, S. D.; Li, L. (1998) Urban particle-induced apoptosis and
24             phenotype shifts in human alveolar macrophages. Environ. Health Perspect. 106: 127-132.
25       Hunter, R. J. (1981) Zeta potential in colloid science: principles and applications. Orlando, FL: Academic Press, pp.
26             1-124. [Ottewill, R. H.; Rowell, R. L., eds. Colloid Science].
27       Imrich, A.; Ning, Y. Y.; Kobzik, L. (2000) Insoluble components of concentrated air particles mediate alveolar
28             macrophage responses in vitro. Toxicol.  Appl. Pharmacol. 167: 140-150.
29       Irsigler, G. B.; Visser, P. J.; Spangenberg, P. A. L. (1999) Asthma and chemical bronchitis in vanadium plant
30             workers. Am. J. Ind. Med. 35: 366-374.
31       Jacobs, J.; Kreutzer, R.; Smith,  D. (1997) Rice burning and asthma hospitalizations, Butte County, California,
32             1983-1992. Environ. Health Perspect.  105: 980-985.
33       Jakab, G. J. (1992) Relationship between carbon black particulate-bound formaldehyde, pulmonary antibacterial
34             defenses, and alveolar macrophage phagocytosis. Inhalation Toxicol. 4: 325-342.
3 5       Jakab, G. J. (1993) The toxicologic interactions resulting from inhalation of carbon black and acrolein on
36             pulmonary antibacterial and antiviral defenses. Toxicol. Appl. Pharmacol. 121: 167-175.
37       Jakab, G. J.; Hemenway, D. R.  (1993) Inhalation coexposure to carbon black and acrolein suppresses alveolar
3 8             macrophage phagocytosis and TNF-a release and modulates peritoneal macrophage phagocytosis. Inhalation
39             Toxicol. 5:  275-289.
40       Jakab, G. J.; Clarke, R. W.; Hemenway,  D. R.; Longphre, M. V.; Kleeberger, S. R.; Frank, R. (1996) Inhalation of
41             acid coated carbon black particles impairs alveolar macrophage phagocytosis. Toxicol. Lett. 88: 243-248.
42       Jiang, N.; Dreher, K. L.; Dye, J. A.; Richards, J. H.; Martin, L. D.; Adler, K. B. (2000) Residual oil fly ash induces
43             cytotoxicity and mucin secretion by guinea pig tracheal epithelial cells via an oxidant-mediated mechanism.
44             Toxicol. Appl. Pharmacol. 163: 221-230.
45       Johnston, C. J.; Finkelstein, J. N.; Gelein, R.; Baggs, R.; Oberdorster, G. (1996) Characterization of the early
46             pulmonary inflammatory response associated with PTFE fume exposure. Toxicol. Appl. Pharmacol.
47             140: 154-163.
48       Johnston, C. J.; Finkelstein, J. N.; Gelein, R.; Oberdorster, G. (1998) Pulmonary inflammatory responses and
49             cytokine and antioxidant mRNA levels in the lungs of young and old C57BL/6 mice after exposure to Teflon
50             fumes. Inhalation Toxicol. 10: 931-953.
51       Kadiiska, M. B.; Mason, R. P.;  Dreher, K. L.; Costa, D. L.; Ohio, A. J. (1997) In vivo evidence of free radical
52             formation in the rat lung after exposure to an emission source air pollution particle. Chem.  Res. Toxicol.
53             10: 1104-1108.
         April 2002                                     7-116         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Kennedy, T.; Ohio, A. I; Reed, W.; Samet, I; Zagorski, I; Quay, I; Carter, I; Dailey, L.; Hoidal, J. R.; Devlin,
  2             R. B. (1998) Copper-dependent inflammation and nuclear factor-KB activation by paniculate air pollution.
  3             Am. J. Respir. Cell Mol. Biol. 19: 366-378.
  4       Killingsworth, C. R.; Alessandrini, F.; Krishna Murthy, G. G.; Catalano, P. J.; Paulauskis, J. D.; Godleski, J. J.
  5             (1997) Inflammation, chemokine expression, and death in monocrotaline-treated rats following fuel oil fly
  6             ash inhalation. Inhalation Toxicol. 9: 541-565.
  7       Kim, S.; Chang, M.-C.; Kim, D.; Sioutas, C. (2000a) A new generation of portable coarse, fine, and ultrafine
  8             particle  concentrators for use in inhalation toxicology. In: Phalen, R. F., ed. Inhalation Toxicology:
  9             proceedings of the third colloquium on paniculate air pollution and human health (first special issue); June,
10             1999; Durham, NC. Inhalation Toxicol. 12 (suppl. 1): 121-137.
11       Kim, S.; Sioutas, C.; Chang, M.-C.; Gong, H., Jr. (2000b) Factors affecting the stability of the performance of
12             ambient fine-particle concentrators.  In: Grant, L. D., ed. PM2000: paniculate matter and health. Inhalation
13             Toxicol. 12(suppl. 4): 281-298.
14       Kimmel, T. A.; Chen, L. C.; Bosland, M. C.; Nadziejko, C. (1997) Influence of acid aerosol droplet size on
15             structural changes in the rat lung caused by acute exposure  to sulfuric acid and ozone. Toxicol. Appl.
16             Pharmacol. 144: 348-355.
17       Kitabatake, M.; Imai, M.; Kasama, K.; Kobayashi, I.; Tomita, Y.; Yoshida, K. (1979) Effects of air pollutants on the
18             experimental induction of asthma attacks in guinea pigs: sulfuric acid mist and mixture of the mist and sulfur
19             dioxide. Mie Med. J. 29: 29-36.
20       Kleeberger, S.  R.; Levitt, R. C.; Zhang, L.-Y.; Longphre, M.; Harkema, J.; Jedlicka, A.; Eleff, S. M.;
21             DiSilvestre, D.; Holroyd, K. J. (1997) Linkage analysis of susceptibility to ozone-induced lung inflammation
22             in inbred mice. Nat. Genet. 17: 475-478.
23       Kleinman, M. T.; Leaf, D. A.; Kelly, E.; Caiozzo, V.; Osann, K.; O'Niell, T. (1998) Urban angina in the mountains:
24             effects of carbon monoxide and mild hypoxemia on subjects with chronic stable angina. Arch. Environ.
25             Health 53: 388-397.
26       Kleinman, M. T.; Mautz, W. J.; Bjarnason, S. (1999) Adaptive and non-adaptive responses in rats exposed to ozone,
27             alone and in mixtures, with acidic aerosols. Inhalation Toxicol. 11:  249-264.
28       Kleinman, M. T.; Bufalino, C.; Rasmussen, R.; Hyde, D.; Bhalla, D. K.; Mautz, W. J. (2000) Toxicity of chemical
29             components of ambient fine paniculate matter (PM2 5) inhaled by aged rats. J. Appl. Toxicol. 20: 357-364.
30       Knox, R. B.; Suphioglu, C.; Taylor, P.; Desai, R.; Watson, H. C.; Peng, J.  L.; Bursill, L. A. (1997) Major grass
31             pollen allergen Lol p 1 binds to diesel exhaust particles: implications for asthma and air pollution. Clin. Exp.
32             Allergy 27: 246-251.
33       Kobzik, L. (1995) Lung macrophage uptake of unopsonized environmental particulates: role of scavenger-type
34             receptors. J. Immunol. 155: 367-376.
35       Kobzik, L.; Goldsmith, C.-A. W.; Ning, Y. Y.; Qin, G.; Morgan, B.; Imrich, A.; Lawrence J.; Murthy, G. G. K.;
36             Catalano, P. J. (2001) Effects of combined ozone and air pollution particle exposure in mice. Boston, MA:
37             Health Effects Institute; research report no. 106. Available: http://www.healtheffects.org/pubs-recent.htm [11
38             April 2002].
39       Kodavanti, U.  P.; Costa, D. L. (1999) Animal models to study for  pollutant effects. In: Holgate, S. T.; Samet,  J. M.;
40             Koren, H.  S.; Maynard, R. L., eds. Air pollution and health. New York, NY: Academic Press; pp. 165-197.
41       Kodavanti, U.  P.; Jaskot, R. H.; Bonner, J.; Badgett, A.; Dreher, K. L. (1996) Eosinophilic lung inflammation in
42             particulate-induced lung injury: technical consideration in isolating RNA for gene expression studies.
43             Exp. Lung Res. 22: 541-554.
44       Kodavanti, U.  P.; Jaskot, R. H.; Su, W. Y.; Costa, D. L.; Ohio, A.  J.; Dreher, K. L. (1997a) Genetic variability in
45             combustion particle-induced chronic lung injury. Am. J. Physiol. 272: L521-L532.
46       Kodavanti, U.  P.; Jaskot, R. H.; Costa, D. L.; Dreher, K. L. (1997b) Pulmonary proinflammatory gene induction
47             following acute exposure to residual oil fly ash:  roles of particle-associated metals. Inhalation Toxicol.
48             9:679-701.
49       Kodavanti, U.  P.; Hauser, R.; Christian!, D. C.; Meng, Z. H.; McGee, J.; Ledbetter, A.; Richards, J.; Costa, D. L.
50             (1998a) Pulmonary responses to oil fly ash particles in the rat differ by virtue of their specific soluble metals.
51             Toxicol. Sci. 43:204-212.
52       Kodavanti, U.  P.; Costa, D. L.; Bromberg, P. A. (1998b) Rodent models of cardiopulmonary disease: their potential
53             applicability in studies of air pollutant susceptibility. Environ. Health Perspect.  106(suppl. 1): 111-130.
54       Kodavanti, U.  P.; Jackson, M. C.; Ledbetter, A. D.; Richards, J. R.; Gardner, S. Y.; Watkinson, W. P.; Campen,
55             M. J.; Costa, D. L. (1999) Lung injury from intratracheal and inhalation exposures to residual oil fly ash in a


         April 2002                                     7-117         DRAFT-DO NOT  QUOTE OR CITE

-------
  1             rat model of monocrotaline-induced pulmonary hypertension. J. Toxicol. Environ. Health Part A
  2             57:  101-121.
  3       Kodavanti, U. P.; Mebane, R.; Ledbetter, A.; Krantz, T.; McGee, J.; Jackson, M. C.; Walsh, L.; Hilliard, H.; Chen,
  4             B. Y.; Richards, J.; Costa, D. L. (2000a) Variable pulmonary responses from exposure to concentrated
  5             ambient air particles in a rat model of bronchitis. Toxicol. Sci. 54: 441-451.
  6       Kodavanti, U. P.; Schladweiler, M. C.; Ledbetter, A. D.; Watkinson, W. P.; Campen, M. J.; Winsett, D. W.;
  7             Richards, J. R.; Crissman, K. M.; Hatch, G. E.; Costa, D. L. (2000b) The spontaneously hypertensive rat as a
  8             model of human cardiovascular disease: evidence of exacerbated cardiopulmonary injury and oxidative stress
  9             from inhaled emission paniculate matter. Toxicol. Appl. Pharmacol. 164: 250-263.
10       Kodavanti, U. P.; Schladweiler, M. C. J.; Richards, J. R.; Costa, D. L. (2001) Acute lung injury from intratracheal
11             exposure to fugitive residual oil fly ash and its constituent metals in normo- and spontaneously hypertensive
12             rats. Inhalation Toxicol. 13: 37-54.
13       Kodavanti, U. P.; Schladweiler, M. C.; Ledbetter, A. D.; Hauser, R.; Christiani, D. C.; McGee, J.; Richards, J. R.;
14             Costa, D. L. (2002) Temporal association between pulmonary and systemic effects of paniculate matter in
15             healthy and cardiovascular compromised rats. J. Toxicol. Environ. Health: in press.
16       Kuschner, W. G.; Wong, H.; D'Alessandro, A.; Quinlan, P.; Blanc, P. D. (1997) Human pulmonary responses to
17             experimental inhalation of high concentration fine and ultrafine magnesium oxide particles. Environ. Health
18             Perspect. 105:  1234-1237.
19       Lambert, A. L.; Dong, W.; Winsett, D. W.; Selgrade, M. K.;  Gilmour, M. I. (1999) Residual oil fly ash exposure
20             enhances allergic sensitization to house dust mite. Toxicol. Appl. Pharmacol. 158: 269-277'.
21       Lambert, A. L.; Dong, W.; Selgrade, M. K.; Gilmour, M. I. (2000) Enhanced allergic sensitization by residual oil fly
22             ash particles is mediated by soluble metal constituents. Toxicol. Appl. Pharmacol. 165: 84-93.
23       Last, J. A.; Pinkerton, K. E. (1997) Chronic exposure of rats  to ozone and  sulfuric acid aerosol: biochemical and
24             structural responses. Toxicology 116: 133-146.
25       Last, J. A.; Hyde, D. M.; Chang, D. P. Y. (1984) A mechanism of synergistic lung damage by ozone and a
26             respirable aerosol. Exp. Lung Res. 7: 223-235.
27       Lay, J. C.; Bennett, W. D.; Kim, C. S.; Devlin, R. B.; Bromberg, P. A. (1998) Retention and intracellular
28             distribution of instilled iron oxide particles in human alveolar macrophages. Am. J. Respir. Cell Mol. Biol.
29             18:687-695.
30       Lay, J. C.; Zeman, K. L.; Ohio, A. J.; Bennett, W. D. (2001)  Effects of inhaled iron oxide particles on alveolar
31             epithelial permeability in normal subjects. Inhalation Toxicol. 13: 1065-1078.
32       Leduc, D.; Fally,  S.; De Vuyst, P.; Wollast, R.; Yernault, J.-C. (1995) Acute exposure to realistic acid fog: effects
33             on respiratory function and airway responsiveness in asthmatics. Environ. Res. 71: 89-98.
34       Lee, M. M.; Green, F. H. Y.; Roth, S. H.; Karkhanis, A.; Bjarnason, S. G.; Schurch, S. (1999) Sulfuric acid aerosol
3 5             induces changes in alveolar surface tension in the guinea pig but not in the rat. Exp. Lung Res. 25: 229-244.
36       Leikauf, G. D.; McDowell, S. A.; Gammon, K.; Wesselkamper, S. C.; Bachurski, C. J.;  Puga, A.; Wiest, J. S.;
37             Leikauf, J. E.; Prows, D. R. (2000) Functional genomics of particle-induced lung injury. Inhalation Toxicol.
38             12:59-73.
39       Leikauf, G.; McDowell, S. A.; Wesselkamper, S. C.; Miller,  C. R.; Hardie, W. D.; Gammon, K.; Biswas, P. P.;
40             Korfhagen, T. R.; Bachurski, C. J.; Wiest, J. S.; Willeke, K.; Bingham, E.; Leikauf, J. E.; Aronow, B. J.;
41             Prows, D. R. (2001) Pathogenomic mechanisms for paniculate matter induction of acute lung injury and
42             inflammation in mice. Boston, MA:  Health Effects Institute; research report no. 105. Available:
43             http://www.healtheffects.org/pubs-recent.htm [11 April 2002].
44       Li, X. Y.; Gilmour, P. S.; Donaldson, K.; MacNee, W. (1996) Free radical activity and pro-inflammatory effects of
45             particulate air pollution (PM10) in vivo and in vitro. Thorax 51:  1216-1222.
46       Li, X. Y.; Gilmour, P. S.; Donaldson, K.; MacNee, W. (1997) In vivo  and in vitro proinflammatory effects of
47             particulate air pollution (PM10). In: Driscoll, K. E.; Oberdorster, G., eds. Proceedings of the sixth
48             international meeting on the toxicology of natural and man-made  fibrous and non-fibrous particles;
49             September 1996; Lake Placid, NY. Environ. Health Perspect. Suppl. 105(5): 1279-1283.
50       Li, X. Y.; Brown, D.; Smith, S.; MacNee, W.; Donaldson,  K. (1999) Short-term inflammatory responses following
51             intratracheal instillation of fine and ultrafine carbon black in rats. Inhalation Toxicol. 11: 709-731.
52       Linn, W. S.; Shamoo, D. A.; Anderson, K.  R.; Peng, R.-C.; Avol, E. L.;  Hackney, J. D.  (1994) Effects  of prolonged,
53             repeated exposure to ozone, sulfuric acid, and their combination in healthy and asthmatic volunteers. Am. J.
54             Respir. Crit. Care Med. 150: 431-440.
         April 2002                                      7-118         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Linn, W. S.; Gong, H., Jr.; Shamoo, D. A.; Anderson, K. R.; Avol, E. L. (1997) Chamber exposures of children to
  2             mixed ozone, sulfur dioxide, and sulfuric acid. Arch. Environ. Health 52: 179-187.
  3       Lippmann, M. (2000) Environmental toxicants: human exposures and their health effects. 2nd ed. New York, NY:
  4             Wiley-Interscience.
  5       Lipsett, M.; Hurley, S.; Ostro, B. (1997) Air pollution and emergency room visits for asthma in Santa Clara County,
  6             California. Environ. Health Perspect.  105: 216-222.
  7       Lison, D.; Lardot, C.; Huaux, F.; Zanetti, G.; Fubini, B. (1997) Influence of particle surface area on the toxicity of
  8             insoluble manganese dioxide dusts. Arch. Toxicol. 71: 725-729.
  9       Long, C. M.; Suh, H. H.; Kobzik, L.; Catalano, P. I; Ning, Y. Y.; Koutrakis, P. (2001) A pilot investigation of the
10             relative toxicity of indoor and outdoor fine particles: in vitro effects of endotoxin and other paniculate
11             properties. Environ. Health Perspect.  109: 1019-1026.
12       Longphre, M.; Li, D.; Matovinovic, E.; Gallup, M.; Samet, J. M.; Basbaum, C. B. (2000) Lung mucin production is
13             stimulated by the air pollutant residual oil fly ash. Toxicol. Appl. Pharmacol. 162: 86-92.
14       Lorz, C.; Lopez, J. (1997) Incidence of air pollution in the pulmonary surfactant system of the pigeon (Columbia
15             livid). Anat. Rec. 249: 206-212.
16       Madden, M. C.; Thomas, M. J.; Ohio, A. J. (1999) Acetaldehyde (CH3CHO) production in rodent lung after
17             exposure to metal-rich particles. Free Radical Biol. Med. 26: 1569-1577.
18       Madden, M. C.; Richards, J. H.; Dailey, L. A.; Hatch, G. E.; Ohio, A. J. (2000) Effect of ozone on diesel exhaust
19             particle toxicity in rat lung. Toxicol. Appl. Pharmacol. 168:  140-148.
20       Madl, A. K.; Wilson, D. W.; Segall, H. J.; Pinkerton, K. E. (1998) Alteration in lung particle translocation,
21             macrophage function, and microfilament arrangement in monocrotaline-treated rats. Toxicol. Appl.
22             Pharmacol. 153:28-38.
23       Mauderly, J. L. (1993) Toxicological approaches to complex mixtures. Environ. Health Perspect.  Suppl.
24             101(4): 155-165.
25       Mautz, W. J.; Kleinman, M.  T.; Bhalla, D. K.;  Phalen, R. F. (2001) Respiratory tract responses to repeated
26             inhalation of an oxidant and acid gas-particle air pollutant mixture. Toxicol. Sci. 61: 331-341.
27       McDowell,  S. A.;  Gammon,  K.; Bachurski, C.  J.; Weist, J.  S.; Leikauf, J. E.; Prows, D. R. Leikauf, G. D. (2000)
28             Differential gene expression in the initiation and progression of nickel-induced acute lung injury.  Am. J.
29             Respir. Cell Mol. Biol. 23: 466-474.
30       McKenna, I. M.; Gordon, T.; Chen, L. C.; Anver, M. R.; Waalkes, M. P. (1998) Expression of metallothionein
31             protein in the lungs of Wistar rats and C57 and DBA mice exposed to cadmium oxide fumes. Toxicol. Appl.
32             Pharmacol. 153: 169-178.
33       Menetrez, M. Y.; Foarde, K. K.; Ensor, D. S. (2001) An analytical method for the measurement of nonviable
34             bioaerosols. J. Air Waste Manage. Assoc. 51: 1436-1442.
35       Michel,  O.; Nagy, A.-M.; Schroeven, M.; Duchateau, J.; Neve, J.; Fondu, P.; Sergysels, R. (1997) Dose-response
36             relationship to inhaled endotoxin in normal subjects. Am. J. Respir. Crit. Care Med. 156: 1157-1164.
37       Monn, C.; Becker, S. (1999) Cytotoxicity and induction of proinflammatory cytokines from human monocytes
3 8             exposed to fine (PM2 5) and coarse particles (PM10_2 5) in outdoor and indoor air. Toxicol. Appl. Pharmacol.
39             155:245-252.
40       Morgan, K. T.; Patterson, D. L.; Gross, E. A. (1986) Responses of the nasal mucociliary apparatus of F-344 rats to
41             formaldehyde gas. Toxicol. Appl. Pharmacol. 82:  1-13.
42       Moss, O. R.; Gross, E. A.; James, R. A.; Janszen, D. B.; Ross, P. W.; Roberts, K. C.; Howard, A. M.; Harkema, J.
43             R.; Calderon-Garciduenas, L.; Morgan,  K. T. (2001) Respiratory tract toxicity in rats exposed to Mexico City
44             air. Cambridge, MA: Health Effects Institute; research report no. 100. Available:
45             http://www.healtheffects.org/ [31 July, 2001].
46       Muggenburg, B. A.; Barr, E. B.; Cheng, Y. S.; Seagrave, JC.; Tilley, L. P.; Mauderly, J. L. (2000) Effect of inhaled
47             residual oil fly ash on the electrocardiogram of dogs. In: Grant, L. D., ed. PM2000: paniculate matter and
48             health. Inhalation Toxicol. 12(suppl. 4): 189-208.
49       Mukae, H.;  Hogg, J. C.; English, D.; Vincent, R.; VanEeden, S. F. (2000) Phagocytosis of paniculate air pollutants
50             by human alveolar macrophages stimulates the bone marrow. Am. J. Physiol. 279: L924-L931.
51       Mukae, H.;  Vincent, R.; Quinlan, K.; English, D.; Hards, J.; Hogg, J. C.; Van Eeden, S. F. (2001) The effect of
52             repeated exposure to paniculate air pollution (PM10) on the bone marrow. Am. J. Respir. Crit. Care Med.
53             163:201-209.
54       Murphy, S.  A.; BeruBe, K. A.; Pooley, F. D.; Richards, R. J.  (1998) The response of lung epithelium to well
55             characterised fine particles. Life Sci. 62: 1789-1799.


         April 2002                                     7-119        DRAFT-DO NOT QUOTE OR CITE

-------
  1       Nadadur, S. S.; Kodavanti, U. P. (2002) Altered gene expression profiles of rat lung in response to an emission
  2             paniculate and its metal constituents. J. Toxicol. Environ. Health: in press.
  3       Nadadur, S. S.; Schladweiler, M. C. J.; Kodavanti, U. P. (2000) A pulmonary rat gene array for screening altered
  4             expression profiles in air pollutant-induced lung injury. Inhalation Toxicol.  12: 1239-1254.
  5       Nadeau, D.; Vincent, R.; Kumarathasan, P.; Brook, J.; Dufresne, A. (1996) Cytotoxicity of ambient air particles to
  6             rat lung macrophages: comparison of cellular and functional assays. Toxicol. In Vitro 10: 161-172.
  7       Nakamura, T.; Hayashida, Y. (1992) Autonomic cardiovascular responses to smoke exposure in conscious rats. Am.
  8             J. Physiol. 262: R73 8-745.
  9       National Institutes of Health. (1997) Guidelines for the diagnosis and management of asthma: expert panel report 2.
10             Bethesda, MD:  U. S. Department of Health and Human Services, National Heart, Lung, and Blood Institute;
11             publication no.  97-4051.
12       Nemmar, A.; Delaunois, A.; Nemery, B.; Dessy-Doize, C.; Beckers, J.-F.; Sulon, J.; Gustin, P. (1999) Inflammatory
13             effect of intratracheal instillation of ultrafine particles in the rabbit: role of C-fiber and mast cells.
14             Toxicol. Appl. Pharmacol. 160: 250-261.
15       Nightingale, J. A.; Maggs,  R.; Cullinan, P.;  Donnelly, L. E.; Rogers, D. F.; Kinnersley, R.; Chung, K. F.; Barnes, P.
16             J.; Ashmore, M.; Newman-Taylor, A. (2000) Airway inflammation after controlled exposure to diesel
17             exhaust particulates. Am. J. Respir. Crit. Care Med. 162: 161-166.
18       Ning, Y.; Imrich, A.; Goldsmith, C. A.; Qin, G.; Kobzik, L.  (2000) Alveolar macrophage cytokine production in
19             response to air particles in vitro: role of endotoxin.  J. Toxicol. Environ. Health A 59: 165-180.
20       Nordenhall, C.; Pourazar, J.; Ledin, M. C.; Levin, J. O.; Sandstrom, T.; Adelroth, E. (2001) Diesel exhaust
21             enhances airway responsiveness in asthmatic subjects. Eur. Respir. J. 17: 909-915.
22       Norwood, J., Jr.; Ledbetter, A. D.; Doerfler, D. L.; Hatch, G. E. (2001) Residual oil fly ash inhalation in guinea
23             pigs: influence of absorbate and glutathione depletion. Toxicol. Sci. 61: 144-153.
24       Oberdorster, G. (1996) Effects of ultrafine particles in the lung and potential relevance to environmental particles.
25             In: Marijnissen, J. C. M.; Gradon, L., eds.  Proceedings of Aerosol inhalation, lung transport, deposition and
26             the relation to the environment: recent research frontiers, an international workshop; September, 1995;
27             Warsaw, Poland. London, United Kingdom: Kluwer Academic; pp.  165-174.
28       Oberdorster, G.;  Ferin, J.; Gelein, R.; Soderholm, S. C.; Finkelstein, J. (1992) Role of the alveolar macrophage in
29             lung injury: studies with ultrafine particles. Environ. Health Perspect. 97: 193-199.
30       Oberdorster, G.;  Cox,  C.; Gelein, R. (1997) Intratracheal instillation versus intratracheal inhalation of tracer
31             particles for measuring lung clearance function. Exp. Lung Res. 23: 17-34.
32       Oberdorster, G.;  Finkelstein, J. N.; Johnston, C.;  Gelein, R.; Cox, C.; Baggs, R.; Elder, A. C. P. (2000) Acute
33             pulmonary effects of ultrafine particles in rats and mice. Cambridge, MA: Health Effects Institute; research
34             report no. 96.
35       Oettinger, R.; Drumm, K.; Knorst, M.; Krinyak, P.; Smolarski, R.; Kienast, K. (1999) Production of reactive oxygen
36             intermediates by human macrophages exposed to soot particles and asbestos fibers and increase in NF-kappa
37             B p50/p!05 mRNA. Lung 177: 343-354.
38       Ohtoshi, T.; Takizawa, H.; Okazaki, H.; Kawasaki, S.; Takeuchi, N.; Ohta, K.; Ito, K. (1998) Diesel exhaust
39             particles stimulate human airway epithelial cells to produce cytokines relevant to airway inflammation in
40             vitro. J. Allergy Clin. Immunol. 101: 778-785.
41       Ohtsuka, Y.; Clarke, R. W.; Mitzner, W.; Branson, K.; Jakab, G. J.; Kleeberger, S. R. (2000a) Interstrain variation
42             in murine  susceptibility to inhaled acid-coated  particles. Am. J. Physiol. L469-L476.
43       Ohtsuka, Y.; Branson, K. J.; Jedlicka, A. E.; Mitzner, W.; Clarke, R. W.; Zhang, L.-Y.; Eleff, S. M.; Kleeberger,
44             S. R. (2000b) Genetic linkage analysis of susceptibility to particle exposure in mice. Am. J. Respir. Cell Mol.
45             Biol. 22: 574-581.
46       Oortgiesen, M.; Veronesi, B.; Eichenbaum,  G.; Kiser, P. F.; Simon, S. A. (2000) Residual oil fly ash and charged
47             polymers activate epithelial cells and nociceptive sensory neurons. Am. J. Physiol. 278: L683-L695.
48       Osier, M.; Oberdorster, G.  (1997) Intratracheal inhalation vs intratracheal instillation: differences  in particle effects.
49             Fundam. Appl.  Toxicol. 40: 220-227.
50       Osier, M.; Baggs, R. B.; Oberdorster, G. (1997) Intratracheal instillation versus intratracheal inhalation: influence of
51             cytokines  on inflammatory response. In: Driscoll, K. E.; Oberdorster, G., eds. Proceedings of the sixth
52             international meeting on the toxicology of natural and man-made fibrous and non-fibrous particles;
53             September 1996; Lake Placid, NY. Environ. Health Perspect. Suppl. 105(5): 1265-1271.
54       Osman, K.; Zejda, J. E.; Schtitz, A.; Mielzynska,  D.; Blinder, C. G.; Vahter, M. (1998) Exposure to lead and other
55             metals in children from Katowice district, Poland. Int. Arch. Occup. Environ. Health71: 180-186.


         April 2002                                     7-120         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Peters, A.; Wichmann, H. E.; Tuch, T.; Heinrich, J.; Heyder, J. (1997) Respiratory effects are associated with the
  2             number of ultrafine particles. Am. J. Respir. Crit. Care Med. 155: 1376-1383.
  3       Petrovic, S.; Urch, B.; Brook, J.; Datema, J.; Purdham, J.; Liu, L.; Lukic, Z.; Zimmerman, B.; Tofler, G.; Downar,
  4             E.; Corey, P.; Tarlo, S.; Broder, L; Dales, R.; Silverman, F. (1999) Cardiorespiratory effects of concentrated
  5             ambient PM2.5: a pilot study using controlled human exposures. In: Phalen, R. F., ed. Inhalation Toxicology:
  6             proceedings of the third colloquium on paniculate air pollution and human health (first special issue); June,
  7             1999; Durham, NC. Inhalation Toxicol. 12(suppl. 1): 173-188.
  8       Pierce, L. M.; Alessandrini, F.; Godleski, J. J.; Paulauskis, J. D. (1996) Vanadium-induced chemokine mRNA
  9             expression and pulmonary inflammation. Toxicol. Appl. Pharmacol.  138: 1-11.
10       Pinkerton, K. E.; Brody, A. R.; Miller, F. J.; Crapo, J.  D. (1989) Exposure to low levels of ozone results in
11             enhanced pulmonary retention of inhaled asbestos fibers. Am. Rev. Respir. Dis. 140: 1075-1081.
12       Pinto, M.; Birnbaum, S. C.; Kadar, T.; Goldberg, G. M. (1979) Lung injury in mice induced by factors acting
13             synergistically with inhaled paniculate antigen. Clin. Immunol. Immunopathol. 13: 361-368.
14       Pope, C. A., III. (1989) Respiratory disease associated with community air pollution and a steel mill, Utah Valley.
15             Am. J. Public Health 79: 623-628.
16       Pope, C. A., III. (1991) Respiratory hospital admissions associated with PM10 pollution in Utah, Salt Lake, and
17             Cache Valleys. Arch. Environ. Health 46: 90-97.
18       Prahalad, A. K.; Soukup, J. M.; Inmon, J.; Willis, R.; Ohio, A. J.; Becker, S.; Gallagher, J. E. (1999) Ambient air
19             particles: effects on cellular oxidant radical generation in relation to paniculate elemental chemistry. Toxicol.
20             Appl. Pharmacol. 158: 81-91.
21       Pritchard, R. J.; Ohio, A. J.; Lehmann, J. R.; Winsett, D. W.; Tepper, J. S.; Park, P.; Gilmour, M. I.; Dreher, K. L.;
22             Costa, D. L. (1996) Oxidant generation and lung injury after paniculate air pollutant exposure increase with
23             the concentrations of associated metals. Inhalation Toxicol. 8: 457-477.
24       Prows, D. R.; Shertzer, H. G.; Daly, M. J.; Sidman, C. L.; Leikauf, G. D. (1997) Genetic analysis of ozone-induced
25             acute lung injury in sensitive and resistant strains of mice. Nat. Genet. 17: 471-474.
26       Prows, D. R.; Leikauf, G. D. (2001) Quantitative trait  analysis of nickel-induced acute lung injury in mice. Am. J.
27             Respir. Cell Mol. Biol. 24: 740-746.
28       Quay, J. L.; Reed, W.; Samet, J.; Devlin, R. B. (1998) Air pollution particles induce IL-6 gene expression in human
29             airway epithelial cells via NF-KB activation. Am. J.  Respir. Cell Mol. Biol. 19: 98-106.
30       Rose, C. S.; Martyny, J. W.; Newman, L. S.; Milton, D. K.; King, T. E., Jr.; Beebe, J. L.; McCammon, J. B.;
31             Hoffman, R. E.; Kreiss, K. (1998) "Lifeguard lung": endemic granulomatous pneumonitis in an indoor
32             swimming pool. Am. J. Public Health 88: 1795-1800.
33       Rudell, B.; Sandstrom, T.; Hammarstrom, U.; Ledin, M.-L.; Horstedt, P.; Stjernberg, N. (1994) Evaluation of an
34             exposure setup for studying effects of diesel exhaust in humans. Int.  Arch. Occup. Environ. Health 66: 77-83.
35       Rudell, B.; Ledin, M.-C.; Hammarstrom, U.; Stjernberg, N.; Lundback, B.;  Sandstrom, T. (1996) Effects on
36             symptoms and lung function in humans experimentally exposed to diesel exhaust. Occup. Environ. Med. 53:
37             658-662.
38       Rudell, B.; Blomberg, A.; Helleday, R.; Ledin, M.-C.; Lundback, B.; Stjernberg, N.; Horstedt, P.; Sandstrom, T.
39             (1999) Bronchoalveolar inflammation after exposure to diesel exhaust: comparison between unfiltered and
40             particle trap filtered exhaust. Occup. Environ. Med. 56: 527-534.
41       Sagai, M.; Furuyama, A.; Ichinose, T. (1996) Biological effects of diesel exhaust particles (DEP). III. Pathogenesis
42             of asthma like symptoms in mice. Free Radical Biol. Med. 21: 199-209.
43       Salvi, S.; Blomberg, A.; Rudell, B.; Kelly, F.; Sandstrom, T.; Holgate, S. T.; Frew, A. (1999) Acute inflammatory
44             responses in the airways and peripheral blood after short-term exposure to diesel exhaust in healthy human
45             volunteers. Am. J. Respir. Crit. Care Med. 159: 702-709.
46       Salvi, S.  S.; Nordenhall, C.; Blomberg, A.; Rudell, B.; Pourazar, J.; Kelly, F. J.; Wilson, S.; Sandstrom, T.; Holgate,
47             S. T.; Frew, A. J. (2000) Acute exposure to diesel exhaust increases IL-8 and GRO-aproduction in healthy
48             human airways. Am. J. Respir. Crit. Care Med.  161: 550-557.
49       Samet, J. M.; Reed, W.; Ohio, A. J.; Devlin, R. B.; Carter, J. D.; Dailey, L.  A.; Bromberg, P. A.; Madden, M. C.
50             (1996) Induction of prostaglandin H synthase 2 in human airway epithelial cells exposed to residual oil fly
51             ash. Toxicol. Appl. Pharmacol. 141: 159-168.
52       Samet, J. M.; Stonehuerner, J.; Reed, W.; Devlin, R. B.; Dailey, L. A.; Kennedy, T. P.; Bromberg, P. A.; Ohio, A. J.
53             (1997) Disruption of protein tyrosine phosphate homeostasis in bronchial epithelial cells exposed to oil fly
54             ash. Am. J. Physiol. 272: L426-L432.
         April 2002                                      7-121         DRAFT-DO NOT QUOTE OR CITE

-------
  1       Samet, J. M; Graves, L. M; Quay, I; Dailey, L. A.; Devlin, R. B.; Ohio, A. I; Wu, W.; Bromberg, P. A.; Reed, W.
  2             (1998) Activation of MAPKs in human bronchial epithelial cells exposed to metals. Am. J. Physiol.
  3             275:L551-L558.
  4       Samet, J. M.; Ohio, A. J.; Costa, D. L.; Madden, M. C. (2000) Increased expression of cyclooxygenase 2 mediates
  5             oil fly ash-induced lung injury. Exp. Lung Res. 26: 57-69.
  6       Schluter, T.; Berg, L; Dorger, M.; Gercken, G. (1995) Effect of heavy metal ions on the release of reactive oxygen
  7             intermediates by bovine alveolar macrophages. Toxicology 98: 47-55.
  8       Schwartz, J. (1994) What are people dying of on high air pollution days? Environ. Res. 64: 26-35.
  9       Seagrave, J. C.; Nikula, K. J. (2000) Multiple modes of responses to air pollution paniculate materials in
10             A549 alveolar type II cells. In: Grant, L. D., ed. PM2000: paniculate matter and health. Inhalation Toxicol.
11             12(suppl. 4): 247-260.
12       Seaton, A.; MacNee, W.; Donaldson, K.; Godden, D. (1995) Paniculate  air pollution and acute health effects.
13             Lancet (8943): 176-178.
14       Shukla, A.; Timblin, C.; BeruBe, K.; Gordon, T.; McKinney, W.; Driscoll, K.; Vacek, P.; Mossman, B. T. (2000)
15             Inhaled paniculate matter causes expression of nuclear factor (NF)-icB-related genes and oxidant-dependent
16             NF-KB activation in vitro. Am. J. Respir. Cell Mol. Biol. 23: 182-187.
17       Silbajoris, R.; Ohio, A. J.; Samet, J. M.; Jaskot, R.; Dreher, K. L.; Brighton, L. E. (2000)  In vivo and in vitro
18             correlation of pulmonary map kinase activation following metallic exposure. Inhalation Toxicol. 12: 453-468.
19       Sindhu, R. K.; Mautz, W. J.; Kikkawa, Y. (1998) Chronic exposure to ozone and nitric acid vapor results in
20             increased levels of rat pulmonary putrescine. Arch. Toxicol. 72: 445-449.
21       Sioutas, C.; Koutrakis, P.; Burton, R. M.  (1995a) A technique to expose  animals to concentrated fine ambient
22             aerosols. Environ. Health Perspect. 103: 172-177.
23       Sioutas, D.; Koutrakis, P.; Ferguson, S. T.; Burton, R. M. (1995b) Development and evaluation of a prototype
24             ambient particle concentrator for inhalation exposure studies. Inhalation Toxicol. 7: 633-644.
25       Sioutas, C.; Kim, S.; Chang, M.; Terrell,  L. L.; Gong, H., Jr. (2000) Field evaluation of a modified DataRAM MIE
26             scattering monitor for real-time PM2 5 mass concentration measurements. Atmos. Environ.  34: 4829-4838.
27       Sjogren, B. (1997) Occupational exposure to dust: inflammation and  ischaemic heart disease. Occup. Environ. Med.
28             54:466-469.
29       Smith, K. R.; Aust, A. E. (1997) Mobilization of iron from urban particulates leads to generation of reactive oxygen
30             species in vitro and induction of ferritin synthesis in human lung epithelial cells. Chem. Res. Toxicol.
31             10:828-834.
32       Smith, K. R.; Veranth, J. M.; Lighty, J. S.; Aust, A. E. (1998) Mobilization of iron from coal fly ash was dependent
33             upon the particle size and the source of coal. Chem.  Res. Toxicol. 11: 1494-1500.
34       Soukup, J. M.;  Ohio, A. J.; Becker, S. (2000) Soluble components of Utah Valley paniculate pollution alter alveolar
35             macrophage function in vivo and in vitro.  Inhalation Toxicol. 12:  401-414.
36       Soukup, J. M.;  Becker, S. (2001) Human alveolar macrophage responses to air pollution particulates are associated
37             with insoluble components of coarse material, including paniculate endotoxin. Toxicol. Appl. Pharmacol.
38             171:20-26.
39       Spengler, J. D.; Thurston, G. D. (1983) Mass and elemental composition of fine and coarse particles in six U.S.
40             cities. J. AirPollut. Control Assoc. 33: 1162-1171.
41       Steerenberg, P. A.; Dormans, J. A. M. A.; Van Doom, C. C. M.; Middendorp, S.; Vos, J.  G.; VanLoveren, H.
42             (1999) A pollen model in the rat for testing adjuvant activity of air pollution components. Inhalation Toxicol.
43             11: 1109-1122.
44       Stringer,  B.; Kobzik, L. (1996) Alveolar macrophage uptake of the environmental paniculate titanium dioxide:
45             role of surfactant components. Am. J. Respir. Cell Mol. Biol. 14:  155-160.
46       Stringer,  B.; Kobzik, L. (1998) Environmental particulate-mediated cytokine production in lung epithelial cells
47             (A549):  role of preexisting inflammation and oxidant stress. J. Toxicol. Environ. Health Part A 55: 31-44.
48       Stringer,  B.; Imrich, A.; Kobzik, L. (1996) Lung epithelial cell (A549) interaction with unopsonized  environmental
49             particulates: quantitation of particle-specific binding and IL-8  production. Exp. Lung Res.  22: 495-508.
50       Su, W.-Y.; Jaskot, R. H.; Richards, J.; Abramson, S. R.; Woessner, J. F., Jr.; Yu, W.-H.; Dreher, K. L. (2000)
51             Induction of pulmonary matrilysin expression by combustion and ambient air particles. Am. J. Physiol. 279:
52             L152-L160.
53       Su, W.-Y.; Jaskot, R. H.; Dreher, K. L. (2000b) Paniculate matter induction of pulmonary gelatinase A, gelatinase
54             B, tissue inhibitor of metalloproteinase expression. In: Phalen, R.  F., ed. Inhalation Toxicology: proceedings
         April 2002                                      7-122        DRAFT-DO NOT QUOTE OR CITE

-------
  1             of the third colloquium on paniculate air pollution and human health (second special issue); June, 1999;
  2             Durham, NC. Inhalation Toxicol. 12(suppl. 2): 105-119.
  3       Sun, G.; Crissman, K.; Norwood, J.; Richards, J.; Slade, R.; Hatch, G. E. (2001) Oxidative interactions of synthetic
  4             lung epithelial lining fluid with metal-containing paniculate matter. Am. J. Physiol. 281: L807-L815.
  5       Takano, H.; Yoshikawa, T.; Ichinose, T.; Miyabara, Y.; Imaoka, K.; Sagai, M. (1997) Diesel exhaust particles
  6             enhance antigen-induced airway inflammation and local cytokine expression in mice. Am. J. Respir. Crit.
  7             Care Med. 156: 36-42.
  8       Takano, H.; Ichinose,  T.; Miyabara, Y.; Shibuya, T.; Lim, H.-B.; Yoshikawa, T. Sagai, M. (1998) Inhalation of
  9             diesel exhaust enhances allergen-related eosinophil recruitment and airway hyperresponsiveness in mice.
10             Toxicol. Appl.  Pharmacol. 150: 328-337.
11       Tan, W. C; Qiu, D.; Liam, B. L.; Ng, T. P.; Lee, S. H.; Van Eeden, S. F.; D'Yachkova, Y.; Hogg, J. C. (2000) The
12             human bone marrow response to acute air pollution caused by forest fires. Am. J. Respir. Crit. Care Med.
13             161: 1213-1217.
14       Terashima, T.; Wiggs, B.; English, D.; Hogg, J. C.; Van Eeden, S. F. (1997) Phagocytosis of small carbon particles
15             (PM10) by alveolar macrophages stimulates the release of polymorphonuclear leukocytes from bone marrow.
16             Am. J. Respir. Crit.  Care Med. 155: 1441-1447.
17       Thurston, G. D.; Lippmann, M.; Scott, M. B.; Fine, J. M. (1997) Summertime haze air pollution and children with
18             asthma. Am.  J.  Respir. Crit. Care Med. 155: 654-660.
19       Timblin, C.; BeruBe, K.; Chrug, A.; Driscoll, K.; Gordon, T.; Hemenway, D.; Walsh, E.; Cummins, A. B.;
20             Vacek, P.; Mossman, B. (1998) Ambient paniculate matter  causes activation of the c-jun
21             kinase/stress-activated protein kinase cascade and DNA synthesis in lung epithelial cells. Cancer Res.
22             58:4543-4547.
23       Tsien, A.;  Diaz-Sanchez, D.; Ma, J.; Saxon, A. (1997) The organic component of diesel exhaust particles and
24             phenanthrene, a major polyaromatic hydrocarbon constituent, enhances IgE production by IgE-secreting
25             EBV-transformed human B cells in vitro. Toxicol. Appl. Pharmacol. 142: 256-263.
26       Tsuchiyama, F.; Hisanaga, N.; Shibata, E.; Aoki, T.; Takagi, H.; Ando, T.; Takeuchi, Y. (1997) Pulmonary metal
27             distribution in urban dwellers. Int. Arch. Occup. Environ. Health 70: 77-84.
28       Turpin, B. J. (1999) Options for characterizing organic paniculate matter. Environ. Sci. Technol. 33: 76A-79A.
29       U.S. Environmental Protection Agency. (1982) Air quality criteria for paniculate matter and sulfur oxides. Research
3 0             Triangle Park, NC: Office of Health and Environmental Assessment, Environmental Criteria and Assessment
31             Office; EPA  report no. EPA-600/8-82-029aF-cF. 3v. Available from: NTIS, Springfield, VA; PB84-156777.
32       U.S. Environmental Protection Agency. (1989) An acid aerosols issue paper: health effects and aerometrics.
33             Research Triangle Park, NC: Office of Health and Environmental Assessment, Environmental Criteria and
34             Assessment Office;  EPA report no. EPA-600/8-88-005F. Available from: NTIS, Springfield, VA;
35             PB91-125864.
36       U.S. Environmental Protection Agency. (1996a) Air quality criteria for paniculate matter. Research Triangle Park,
37             NC: National Center for Environmental Assessment-RTF Office; report nos. EPA/600/P-95/00laF-cF. 3v.
38       U.S. Environmental Protection Agency. (1996b) Ambient levels and noncancer health effects of inhaled crystalline
39             and amorphous silica: health issue assessment [draft]. Research Triangle Park, NC: National Center for
40             Environmental  Assessment; EPA report no. EPA/600/R-95/115.
41       U.S. Environmental Protection Agency. (2000) Health assessment document for diesel exhaust [SAB review draft].
42             Washington,  DC: Office of Research and Development, National Center for Environmental
43             Assessment-Washington Office; report no. EPA/600/8-90/057E. Available:
44             http://cfpub.epa.gov/ncea/cfm/nceahome.cfm  [29 March, 2002].
45       Utell, M. J.; Morrow,  P. E.; Hyde, R. W. (1983) Latent development of airway hyperreactivity in human subjects
46             after sulfuric acid aerosol exposure. J. Aerosol Sci. 14: 202-205.
47       Van Eeden, S. F.; Tan, W.  C.; Suwa, T.; Mukae, H.; Terashima, T.; Fujii, T.; Qui, D.; Vincent, R.; Hogg, J. C.
48             (2001) Cytokines involved in the systemic inflammatory response induced by exposure to paniculate matter
49             air pollutants (PM10). Am. J. Respir. Crit. Care Med. 164: 826-830.
50       VanMaanen, J. M.; Borm, P. J.; Knaapen, A.; Van Herwijnen, M.; Schilderman, P. A.; Smith, K. R.; Aust, A. E.;
51             Tomatis, M.; Fubini, B. (1999) In vitro effects of coal fly ashes: hydroxyl radical generation, iron release,
52             and DNA damage and toxicity in rat lung epithelial cells. Inhalation Toxicol. 11: 1123-1141.
53       Van Zijverden, M.;  Granum, B. (2000a) Adjuvant activity of paniculate pollutants in different mouse models.
54             Toxicology 152: 69-77.
         April 2002                                     7-123         DRAFT-DO NOT QUOTE OR CITE

-------
  1       VanZijverden, M; Van der Fiji, A.; Bol, M; VanPinxteren, F. A.; De Haar, C.; Penninks, A. H.; Van Loveren, H.;
  2             Pieters, R. (2000b) Diesel exhaust, carbon black, and silica particles display distinct Thl/Th2 modulating
  3             activity. Toxicol. Appl. Pharmacol. 168: 131-139.
  4       Van Zijverden, M.; De Haar, C.; Van Beelen, A.; Van Loveren, H.; Penninks, A.; Pieters, R. (2001)
  5             Co-administration of antigen and particles optimally stimulates the immune response in an intranasal
  6             administration model in mice. Toxicol. Appl. Pharmacol. 177: 174-178.
  7       Vanda, B.; de Buen, N.; Jasso, R.; Valero, G.; Vargas, M. H.; Olmos, R.; Arreola, J. L.; Santillan, P.; Alonso, P.
  8             (1998) Inflammatory cells and ferruginous bodies in bronchoalveolar lavage in urban dogs. Acta Cytol.
  9             42: 939-944.
10       Veronesi, B.; Oortgiesen, M. (2001) Neurogenic inflammation and paniculate matter (PM) air pollutants.
11             Neurotoxicology 22: 795-810.
12       Veronesi, B.; Carter, J. D.; Devlin, R. B.; Simon, S. A.; Oortgiesen, M. (1999a) Neuropeptides and capsaicin
13             stimulate the release of inflammatory cytokines in a human bronchial epithelial cell line. Neuropeptides 33:
14             447-456.
15       Veronesi, B.; Oortgiesen, M.; Carter, J. D.; Devlin, R. B.  (1999b) Paniculate matter initiates inflammatory cytokine
16             release by activation of capsaicin and acid receptors in a human bronchial epithelial cell line. Toxicol. Appl.
17             Pharmacol. 154:  106-115.
18       Veronesi, B.; Oortgiesen, M.; Roy, J.; Carter, J. D.; Simon, S. A.; Gavett, S. H. (2000) Vanilloid (capsaicin)
19             receptors influence inflammatory sensitivity in response to paniculate matter. Toxicol. Appl. Pharmacol. 169:
20             66-76.
21       Veronesi, B.; De Haar, C.; Roy, J.; Oortgiesen, M. (2002a) Paniculate matter inflammation and receptor sensitivity
22             are target cell specific. Inhalation Toxicol. 14:  101-125.
23       Veronesi, B.; De Haar, C.; Lee, L.; Oortgiesen, M. (2002b) The surface charge of paniculate matter related to
24             cytokine release in human bronchial epithelial cells (BEAS-2B). Toxicol. Appl. Pharmacol. 179: XX
25       Vincent, R.; Bjarnason, S. G.; Adamson, I. Y. R.; Hedgecock, C.; Kumarathasan, P.;  Guenette, J.; Potvin, M.;
26             Goegan, P.; Bouthillier, L. (1997) Acute pulmonary toxicity of urban paniculate matter and ozone. Am. J.
27             Pathol. 151:  1563-1570.
28       Vincent, R.; Kumarathasan, P.; Goegan, P.; Bjarnason, S. G.; Guenette, J.; Berube, D.; Adamson, I. Y.; Desjardins,
29             S.; Burnett, R. T.; Miller, F. J.; Battistini, B. (2001) Inhalation toxicology of urban ambient paniculate
30             matter: acute cardiovascular effect in rats. Boston,  MA:  Health Effects Institute; research report no. 104.
31             Available: http://www.healtheffects.org/Pubs/Vincent.pdf [29 January 02].
32       Vogelzang, P. F. J.; Van Der Gulden, J. W. J.; Folgering, H.; Kolk, J. J.;  Heederik, D.; Preller, L.; Tielen, M. J. M.;
33             Van Schayck, C. P. (1998) Endotoxin exposure as  a major determinant of lung function decline in pig
34             farmers. Am. J. Respir. Crit. Care Med. 157: 15-18.
35       Wagner, J. G.; Hotchkiss, J. A.; Harkema, J. R. (2001) Effects of ozone and endotoxin coexposure on rat airway
36             epithelium: potentiation of toxicant-induced alterations. Environ. Health Perspect. 109(suppl. 4): 591-598.
37       Walters, D. M.; Breysse, P. N.; Wills-Karp, M. (2001) Ambient urban Baltimore particulate-induced airway
38             hyperresponsiveness and inflammation in mice. Am. J. Respir. Crit. Care Med. 164: 1438-1443.
39       Watkinson, W. P.; Campen, M. J.; Costa, D. L. (1998) Cardiac arrhythmia induction after exposure to residual oil
40             fly ash particles in a rodent model of pulmonary hypertension. Toxicol. Sci. 41: 209-216.
41       Watkinson, W. P.; Campen, M. J.; Nolan, J. P.; Kodavanti, U. P.; Dreher, K. L.; Su, W.-Y.; Highfill, J. W.; Costa,
42             D. L. (2000a) Cardiovascular effects following exposure to paniculate matter in healthy and
43             cardiopulmonary-compromised rats. In: Heinrich, U.; Mohr, U., eds. Relationships between acute and
44             chronic effects of air pollution. Washington, DC: ISLI Press, pp. 447-463.
45       Watkinson, W. P.; Campen, M. J.; Dreher, K. L.; Su, W.-Y.; Kodavanti, U.P.; Highfill, J. W.;  Costa, D. L. (2000b)
46             Thermoregulatory effects following exposure to paniculate matter in healthy and
47             cardiopulmonary-compromised rats. J. Therm. Biol. 25: 131-137.
48       Wesselkamper, S. C.; Prows, D. R.; Biswas, P.; Willeke, K.; Bingham, E.; Leikauf, G. D. (2000) Genetic
49             susceptibility to irritant-induced acute lung injury in mice. Am. J. Physiol. 279: L575-L582.
50       Wheatley, L. M.; Platts-Mills, T. A. E. (1996) Perennial allergens and the asthma epidemic. Sci.  Med. 3: 6-13.
51       Woodin, M. A.; Hauser, R.; Liu, Y.; Smith, T. J.; Siegel, P. D.; Lewis, D. M.; Tollerud, D. J.;  Christiani, D. C.
52             (1998) Molecular markers of acute upper airway inflammation in workers exposed to fuel-oil ash. Am. J.
53             Respir. Crit. Care Med. 158: 182-187.
54       Wu, W.; Samet, J. M.; Ohio, A. J.; Devlin, R. B. (2001) Activation of the EOF receptor signaling pathway in human
55             airway epithelial cells exposed to Utah Valley PM. Am. J. Physiol. 281: L483-L489.


         April 2002                                     7-124         DRAFT-DO NOT  QUOTE OR CITE

-------
 1      Zelikoff, J. T.; Sisco, M; Cohen, M. D.; Frampton, M. W.; Utell, M. J; Schlesinger, R. B. (1994) Interspecies
 2            comparison of immunotoxicity of inhaled sulfuric acid. II. New Zealand white rabbits. In: 1994 international
 3            conference sponsored by the American Lung Association and the American Thoracic Society; May; Boston,
 4            MA. Am. J. Respir. Crit. Care Med. 149: A621.
 5      Zelikoff, J. T.; Frampton, M. W.; Cohen, M. D.; Morrow, P. E.; Sisco, M.; Tsai, Y.; Utell, M. J.; Schlesinger, R. B.
 6            (1997) Effects of inhaled sulfuric acid aerosols on pulmonary immunocompetence: a comparative study in
 7            humans and animals. Inhalation Toxicol. 9: 731-752.
 8      Zhang, T.; Huang, C.; Johns, E. J. (1997) Neural regulation of kidney function by the somatosensory system in
 9            normotensive and hypertensive rats. Am. J. Physiol. 273: R1749-R1757.
10      Zock, J.-P.; Hollander, A.; Heederik, D.; Douwes, J. (1998) Acute lung function changes and low endotoxin
11            exposures in the potato processing industry. Am. J. Ind. Med. 33: 384-391.
        April 2002                                    7-125        DRAFT-DO NOT QUOTE OR CITE

-------
 i        8.  EPIDEMIOLOGY OF HUMAN HEALTH EFFECTS
 2               FROM AMBIENT PARTICULATE MATTER
 3
 4
 5      8.1  INTRODUCTION
 6          Epidemiology studies linking community ambient PM concentrations to adverse health
 7      effects played an important role in the 1996 PM Air Quality Criteria Document (PM AQCD), and
 8      continue to play an important role. Those studies are indicative of measurable excesses in
 9      pulmonary function decrements, respiratory symptoms, hospital and emergency department
10      admissions, and mortality in human populations being associated with ambient levels of PM25,
11      PM10_2 5, PM10, and other indicators of PM exposure.  The numerous more recent epidemiologic
12      studies reviewed in this chapter generally identify more cities where ambient PM-relationships
13      with morbidity and mortality have been found and, thereby, both extend the earlier findings and
14      provide an expanded evidence base that substantiates health effects being associated with
15      exposures to PM at concentrations currently encountered in the United States.
16          The epidemiology studies presented here should be considered in combination with the
17      ambient concentration information presented in Chapter 3,  the studies of human PM exposure in
18      Chapter 5, and the discussions of PM dosimetry and toxicology in Chapters 6 and 7. The
19      contribution of the epidemiology studies is to evaluate associations between health effects and
20      exposures of human populations to ambient PM and to help identify susceptible subgroups and
21      associated risk factors. Chapter 9 provides a concise interpretive synthesis of the information.
22          This chapter opens with a brief overview of key general features of the several types of
23      epidemiologic studies assessed in  the chapter and a discussion of important general
24      methodological issues that must be considered in their critical assessment. After this brief
25      introduction,  Section 8.2 assesses  studies of PM effects on  mortality. Section 8.3 evaluates
26      studies of morbidity as a health endpoint.  Section 8.4 then provides an interpretive assessment of
27      the overall PM epidemiologic data base in relation  to a variety of key issues and potential
28      inferences associated with studies reviewed in Sections 8.2 and 8.3.  The overall key findings and
29      conclusions for this chapter are then summarized in Section 8.5.
30

        April 2002                                8-1         DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.1.1  Types of Epidemiology Studies Reviewed
 2           Definitions of various types of epidemiology studies used here were provided in the 1996
 3      PM AQCD (U.S. Environmental Protection Agency, 1996a) and are briefly summarized here.
 4      Briefly, the epidemiology studies are divided into mortality studies and morbidity studies.
 5      Mortality studies evaluating PM effects on total (non-accidental) mortality and cause-specific
 6      mortality have provided the most unambiguous evidence of a clearly adverse endpoint. The
 7      morbidity studies further substantiate PM effects on a wide range of health endpoints, such as:
 8      cardiovascular and respiratory-related hospital admissions, medical visits, reports of respiratory
 9      symptoms, self-medication in asthmatics, changes in pulmonary function tests (PFT), low
10      birthweight infants, etc.
11           The epidemiology strategies most commonly used in PM health studies are of four types:
12      (1) ecologic studies; (2) time-series semi-ecologic studies; (3) longitudinal panel and prospective
13      cohort studies; and (4) case-control and crossover studies. All of these are observational studies
14      rather than experimental studies, since participants are not assigned at random to air pollution
15      exposures.  In general, the exposure of the participant is not directly observed, and the
16      concentration of airborne particles and other air pollutants at one or more stationary air monitors
17      is used as a proxy for individual exposure to ambient air pollution.
18           In ecologic studies., the responses are at a community level (for example, annual mortality
19      rates), as are the exposure indices (for example, annual average particulate matter concentrations)
20      and covariates (for example, the percentage of the population greater than 65 years of age).
21      No individual data is used in the analysis, therefore the relation between health effect and
22      exposure calculated across  different communities may not reflect individual-level associations
23      between health outcome and exposure.  The use of proxy measures for individual exposure and
24      covariates or effects modifiers may also bias the results, and within-city or within-unit
25      confounding may be overlooked.
26           Time series studies are more informative because they allow study of associations between
27      changes in outcomes and changes in exposure indicators preceding or simultaneous with the
28      outcome. The temporal relationship supports a conclusion of a causal relation, even when both
29      the outcome (for example, the number of non-accidental deaths in a city during a day) and the
30      exposure (for example, daily air pollution concentration) are community indices.

        April 2002                                 8-2        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Prospective cohort (or panel) studies use data from individuals, including health status
 2      (where available), individual exposure (not usually available), and individual covariates or risk
 3      factors, observed over time.  The participants in a prospective cohort study are ideally recruited,
 4      using a simple or stratified random sample so as to represent a target population for which
 5      individual or community exposure of the participants is known before and during the interval up
 6      to the time the health endpoint occurs.  The use of individual-level data is believed to give
 7      prospective cohort studies greater inferential strength than other epidemiology strategies, but the
 8      use of community-level or estimated exposure data may weaken this advantage, as in time-series
 9      studies.
10           Case-control studies are retrospective studies in that exposure is determined after the health
11      endpoint occurs (this is common in occupational health studies).  As Rothman and Greenland
12      (1998) describe it, "Case-control studies are best understood by defining a source population,
13      which represents a hypothetical study population in which a cohort study might have been
14      conducted ... In a case-control study, the cases are identified and their exposure status is
15      determined just as in a cohort study . . . [and] a control group of study subjects is sampled from
16      the entire source population that gives rise to the cases ... the cardinal requirement of control
17      selection is that the controls must be sampled independently  of their exposure status."
18           The case-crossover design is suited to the study of a transient effect of an intermittent
19      exposure on the subsequent risk of a rare acute-onset disease hypothesized to occur a short time
20      after exposure.  In the original development of the method, effect estimates were based on
21      within-subject comparisons of exposures associated with incident disease events with exposures
22      at times before the occurrence of disease, using matched case-control methods or methods for
23      stratified follow-up studies with spare data within each stratum.  The principle of the analysis is
24      that the exposures of cases just before the event are compared with the distribution of exposure
25      estimated from some separate time period. This distribution is assumed to be representative of
26      the distribution of exposures for those individuals while they are at risk of developing the
27      outcome of interest.
28           When measurements of exposure or potential effect modifiers are available on an
29      individual level, it is possible to incorporate this information into a case-crossover study unlike a
30      time-series analysis.  A disadvantage of the case-crossover design, however, is the potential for
31      bias due to time trends in the exposure time-series.  Since case-crossover comparisons are made

        April 2002                                 8-3        DRAFT-DO NOT QUOTE  OR CITE

-------
 1      between different points in time, the case-crossover analysis implicitly depends on an assumption
 2      that the exposure distribution is stable over time (stationary). If the exposure time-series is
 3      non-stationary and case exposures are compared with referent exposures systematically selected
 4      from a different period in time, a bias may be introduced into estimates of the measure of
 5      association for the exposure and disease.  These biases are particularly important when
 6      examining the small associations that appear to exist between PM and health outcomes.
 7
 8      8.1.2  Confounding and Effect Modification
 9           A pervasive problem in the analysis of epidemiology data, no matter what design or
10      strategy, is the unique attribution of the health outcome to the nominal causal agent (i.e.,  airborne
11      particles) in this document.  The health outcomes attributed to particles are not specific (for
12      example, mortality in a broad range of ICD-9 categories) and may also be attributable to high or
13      low temperatures, influenza and other diseases, and/or exposure to gaseous criteria air pollutants.
14      Many of the other factors can be measured, directly or by proxies.  Some of these co-variables
15      are confounders, others are effect modifiers.  The distinctions are important.
16           Confounding \$ "... a confusion of effects.  Specifically,  the apparent effect of the
17      exposure of interest is distorted because the effect of an extraneous factor is mistaken for or
18      mixed with the actual exposure effect (which may be null)."  (Rothman and Greenland, 1998,
19      p. 120).  These authors list three criteria for a confounding factor:
20           (1) A confounding factor must be a risk factor for the disease (health effect).
21           (2) A confounding factor must be associated with the exposure under study in the source
22               population (the population at risk from which the cases are derived).
23           (3) A confounding factor must not be affected by the exposure or the disease, i.e.,  it cannot
24               be an intermediate step in the causal path between the exposure and the disease.
25           A causal pathway is one in which members of the population are exposed to putative causal
26      agents that can actually produce the observed health effect. The primary cause may be mediated
27      by secondary causes (possibly proximal to exposure) and may have either a direct effect on
28      exposure or an indirect effect through the secondary causes,  or both, as illustrated below.
29      A non-causal pathway may involve factors that are not associated with the health effect or for
30      which there is no population exposure, so that the factors are not potential confounders.

        April 2002                                 8-4         DRAFT-DO NOT QUOTE OR CITE

-------
 1           The determination of whether a potential confounder is an actual confounder depends on
 2      biological or physical knowledge about its exposure and health effects. Patterns of association in
 3      epidemiology may be helpful in suggesting where to look for this knowledge, but do not replace
 4      it.  Gaseous criteria pollutants (CO, NO2, SO2, O3) are candidates for confounders since: (1) all
 5      of these have adverse health effects, with CO more often identified with cardiovascular effects
 6      and the others with respiratory effects (including symptoms and hospital admissions), as part of
 7      the wide spectrum of cardiopulmonary disease also associated with particles; (2) the gaseous
 8      criteria pollutants may be associated with particles for several reasons, including (a) common
 9      sources, (b) correlated changes in response to wind and weather, and (c) SO2 and NO2 may be
10      precursors to sulfate and nitrate components of ambient particle mixes, while NO2 contributes to
11      the formation of organic aerosols during photochemical transformations.
12           A common source, such as combustion of gasoline in motor vehicles emitting CO, NO2,
13      and primary particles, may play an important role in confounding among these pollutants, as does
14      weather and seasonal effects.  Even though O3 is a secondary pollutant also associated with
15      emission of NO2, it is often less highly associated with particles.  Levels of SO2 in the western
16      U.S. are often quite low, so that secondary formation of particle sulfates plays a much smaller
17      role there, resulting in usually relatively little confounding of SO2 with PM mass concentration in
18      the west.  On the other hand, in the industrial midwest and northeastern states, SO2 and sulfate
19      levels during many of the epidemiology studies were relatively high and highly correlated with
20      fine particle mass concentrations, so that criterion 3 (no causal path leading from confounder to
21      exposure, or exposure to confounder to health effect) may not be  strictly true for SO2 vs sulfate
22      or overall fine particle mass.  If the correlation with PM and SO2  is not too high, it may be
23      possible to estimate some part of their independent effects.  If there is a causal pathway, then it is
24      not clear whether the observed relation of exposure to health effect is a direct effect of the
25      exposure, an indirect effect mediated by the confounder, or a mixture of these.
26           Most extraneous variables fall into the category of effect modifiers.  "Effect-measure
27      modification differs from confounding in several ways. The main difference is that, whereas
28      confounding is a bias that the investigator hopes to prevent or remove from the effect estimate,
29      effect-measure modification is a property of the effect under study ... In epidemiologic analysis
30      one tries to eliminate confounding but one tries to detect and estimate effect-measure
31      modification." (Rothman and Greenland, 1998, p. 254). Examples of effect modifiers in some

        April 2002                                  8-5         DRAFT-DO NOT QUOTE OR CITE

-------
 1      of the studies evaluated in this chapter include environmental variables (such as temperature or
 2      humidity in time-series studies), individual risk factors (such as education, cigarette smoking
 3      status, age in a prospective cohort study), and community factors (such as percent of population
 4      > 65 years old).  It is often possible to stratify the relationship between health outcome and
 5      exposure by one or more of these risk factor variables.
 6           Effect modifiers may be encountered within a single-city time series studies, or across cities
 7      in a two-stage hierarchical model or meta-analysis. We will use the latter case to illustrate some
 8      of the possibilities using a hypothetical case with four cities in  which a co-pollutant of the PM
 9      index is to be evaluated as a possible effect modifier.  In the examples in Figure 8-1, we assume
10      that the co-pollutant has a relatively high positive correlation with the PM index. It is also
11      assumed that the excess relative risk for PM is calculated in a model in which PM is the only air
12      pollutant.  For any given co-pollutant concentration within each city, there is likely to be only a
13      modest range of values of the PM index and the associated excess relative risk, as is suggested by
14      the elliptical figures. The relationship between mortality and PM in Figure 8-la is assumed to be
15      the same and positive in all four cities; thus, with increasing co-pollutant concentration  within
16      each city, the excess relative risk increases because the co-pollutant is strongly correlated with
17      the PM index. However, in the hypothetical 8-la, the co-pollutant is not an effect modifier for
18      PM,  as can be shown by a regression of the estimated mean PM effect on the mean co-pollutant
19      concentration across the four cities.
20           The relationship between PM and mortality in Figure 8-lb is assumed to differ across the
21      four cities,  ranging from strongly negative in City 1 to strongly positive in City 4.  Thus, with
22      increasing co-pollutant concentration within each city, the excess relative risk decreases in City 1
23      and City 2 but increases in City 3 and City 4, because the co-pollutant is strongly correlated with
24      the PM index. In the hypothetical Figure 8-lb, the co-pollutant is an effect modifier for PM, as
25      can be shown by a regression of the estimated mean PM effect  on the mean co-pollutant
26      concentration across the four cities, even though the simple mean of the excess relative  risks
27      across the four cities is nearly zero. A relationship would be found if all within-city effects were
28      positive, or if the across-city ecological regression were negative. Stratification  by  levels of the
29      putative effect modifier is also often useful.
30           Potential confounding (Figure 8-2a)  is more difficult to identify and several statistical
31      methods are available, none of them being completely satisfactory. The ususal methods are:

        April 2002                                  8-6         DRAFT-DO NOT QUOTE OR CITE

-------
   Q_
   .0
   0>
   0)
   LJJ
City 1
City 2
City 3
City 4
                       Co-Pollutant Concentration
   Figure 8-la.  Strong within-city association between PM and mortality, but no
               second-stage association.
                                                        City 3
                                                                      City 4
                                          Co-Pollutant Concentration
                    Co-Pollutant Concentration
 Figure 8-lb.  Within-city association between PM and mortality ranges from negative
             to positive with mean across cities approximately zero, but with strong
             positive second-stage association.
April 2002
                            8-7
                    DRAFT-DO NOT QUOTE OR CITE

-------
                                        Confounder
                            Figure 8-2a
                                                       , Outcomes
                                                           ft
                                                        Exposure
                                                        Outcomes
                                         Modifier
                            Figure 8-2b
                                  Primary Cause (s
                                  Secondary Cause(s)
                            Figure 8-2c
                                                        Exposure
                                                        Outcomes
                                            Exposure
                                                        Outcomes
                                 Secondary Cause
                                  Primary Cause(s)
                                 Secondary Cause
                                           : Exposure
                            Figure 8-2d
        Figure 8-2. (a) Graphical depiction of confounding; (b) Graphical depiction of effect
                    modification; (c) Graphical depiction of a causal agent with a secondary
                    confounder; (d) Graphical depiction of a causal agent and two potential
                    confounders.
1
2
3
4
5
6
7
Within a city:
(A)   Fit both a single-pollutant model and then several multi-pollutants models, and
      determine if including the co-pollutants greatly changes the estimated effect and
      inflates its estimated standard error;
(B)   If the PM index and its co-pollutants are nearly multi-collinear, carry out a factor
      analysis, and determine which gaseous pollutants are most closely associated with
      PM in one or more common factors.
       April 2002
                                                  DRAFT-DO NOT QUOTE OR CITE

-------
 1           Using data from several cities:
 2           (C)   Proceed as in Method A and pool the effect size estimates across cities for single-
 3                 and multi-pollutant models;
 4           (D)   Carry out a hierarchical regression of the PM effects vs. the mean co-pollutant
 5                 concentration and determine if there is a significant relationship;
 6           (E)   First carry out a regression of PM vs. the co-pollutant concentration within each city
 7                 and the regression coefficient of mortality vs. PM for each city. Then fit a second-
 8                 stage model regressing the mortality-PM coefficient vs. the PM-co-pollutant
 9                 coefficient, concluding that the co-pollutant is a confounder if there is a significant
10                 regression coefficient at the second stage (See Figure 8-2c).
11           The disadvantages of the methods are discussed in detail in Section 8.4.  Briefly, the multi-
12      pollutant regression coefficients in method A may be unstable and have greatly inflated standard
13      errors, weakening their interpretation.  In method B, the factors may be sensitive to the choice of
14      co-pollutants and the analysis method, and may be difficult to relate to real-world entities.
15      In method C, as with any meta-analysis, it is necessary to consider the heterogeneity of the
16      within-city effects before pooling them. Several large multi-city studies have revealed
17      unexpected heterogeneity, not fully explained at present.
18           While method D is sometimes interpreted as showing confounding if the regression
19      coefficient is non-zero, this  is an argument for effect modification, not confounding.
20           Method E is sensitive  to the assumptions being made. For example, if PM is the primary
21      cause in Figure 8-2c and the co-pollutant the secondary cause, then the two-stage approach may
22      be valid.  However, if the model is mis-specified and there are two or more secondary causes,
23      some of which may not be identified, then the method may give misleading results.
24           An additional issue of great relevance is whether or not the population in a community time
25      series study or the participants in a prospective cohort study are exposed to measurable levels of
26      the potential confounder, particularly the ambient gaseous co-pollutants. If there is no exposure,
27      then the potential confounder does not satisfy the requirement that it is related to both exposure
28      and outcomes.  This is discussed in Section 8.4 in connection with the role of exposure
29      measurement errors in air pollution epidemiology.
30
31

        April 2002                                 8-9        DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.1.3  Selection of Studies for Review and Ambient PM Increments Used to
 2             Report Risk Estimates
 3           Numerous PM epidemiology papers have been published since the 1996 PM AQCD.
 4      An ongoing medline search has been and is continuing to be conducted in conjunction with other
 5      strategies to identify PM literature pertinent to developing criteria for PM NAAQS. Those
 6      epidemiologic studies that relate measures of ambient PM to human health outcomes are
 7      assessed in this chapter, but occupational exposures studies are not. Some of the criteria used for
 8      selecting relevant literature for consideration here include whether a given study presents:
 9      (1) pertinent ambient PM indices:  e.g., PM10, PM25, PM10_25, etc.; (2) analyses of health effects
10      of specific PM chemical or physical constituents (e.g., metals, sulfates, nitrates or ultrafme
11      particles, etc.); (3) health endpoints not previously extensively researched; (4) multiple pollutant
12      analyses; and/or (5) for long-term effects, mortality displacement information.  The publication
13      of pertinent new studies has been  and is proceeding at a prodigious rate; and the review and
14      evaluation of pertinent literature in this PM AQCD development process is an ongoing process
15      which continues to obtain and assess new evidence.
16           The literature review method is similar to those used by others (e.g., Basu and Samet,
17      2000):  (a) Establish a publication base using Medline and other data bases using a set of key
18      words (particles, air pollution, mortality, morbidity, cause of death, PM, and others); (b) add
19      papers to the publication base by staff review of Current Contents and tables of contents of
20      journals in which relevant papers  are published; and (c) staff requests to scientists known to be
21      active in this field for papers recently accepted for publication.  Efforts have been made to assess
22      here pertinent new studies published mainly through December, 2001, as well as some studies
23      published in early 2002 as acquired (if, in the opinion of staff, such recent new papers provide
24      important inputs towards resolving critical scientific uncertainties).
25           The effect of mortality from exposure to PM or other pollutants in this document is usually
26      expressed as a relative risk or risk rate (RR) relative to a baseline mortality or morbidity rate.
27      The crude mortality rates in 88 cities in 48 contiguous states in the NMMAPS study ranged from
28      about 8 deaths per day per million population in Denver, CO to about 40 per day per million in
29      St. Petersburg, FL. It is likely that age-adjusted rates such as those used in the APHEA 2 study
30      (Katsouyanni et al., 2001) would have shown a smaller range.  As reported in Samet et al.
31      (2000a), there was little association between PM10 effect size and crude mortality rate in the

        April 2002                                 8-10        DRAFT-DO NOT QUOTE OR CITE

-------
 1      continental U.S. cites; however, Katsouyanni et al. (2001) found a negative relation between
 2      PM10-equivalent effect size estimates and age-adjusted mortality rate in 29 European cities.
 3      We plotted the relationship between increased or decreased mortality rate in NMMAPS for
 4      ranges between the 25th and 75 percentiles (results not shown), but there was little apparent new
 5      information in those plots other than the RR.
 6           The PM increments used in this document to convert regression coefficients into
 7      meaningful increments of excess risk are based on data from the U.S. fine particle monitoring
 8      network for 1999 and 2000, the most recent years available.  The difference between the annual
 9      mean and the annual 95th percentile was used to characterize annual variation within each site;
10      and the average across  all sites was used to select an appropriate increment for short-term
11      studies, about 50 //g/m3 for PM10 and 25 //g/m3 for PM2 5 and PM10_2 5, after rounding for ease of
12      calculation.  As there is little experimental evidence about differences in effects of fine (PM25)
13      and coarse (PM10_25) particles, common increments are used for both. The difference between the
14      average of annual mean PM concentrations across all sites and the average of the annual 95th
15      percentiles across all sites was about 20 //g/m3 for PM10 and 10 //g/m3 for PM2 5 and PM10_25,
16      which are values used here for PM increments in long-term studies.
17           Thus, the pollutant  increments utilized here to report Relative Risks (RR's) or Odds Ratio
18      for various health effects are: for PM10, 50 //g/m3; for PM25, 25 //g/m3; for SO4=, 155 nmoles/m3
19      (15 //g/m3);  and, for H+, 75 nmoles/m3 (3.6 //g/m3, if as H2SO4) for short-term (<24 h) exposure
20      studies.  The increments for short-term studies are the same as used in the 1996 PM AQCD,
21      a choice now driven by current data.  In the  1996 PM AQCD, the same increments were used for
22      the long- and short-term exposure studies. However, 20 //g/m3 is the increment used here for
23      PM10 and 10 //g/m3 for PM2 5 and PM10_25 for long-term exposure studies. These estimates
24      derived from new 1999-2000 data are smaller than these used for long-term studies in the 1996
25      PM AQCD.
26           Greater emphasis is placed in text discussions on integrating and interpreting findings from
27      the body of evidence provided by the newer studies (as well as relating them to those reviewed in
28      the 1999 PM AQCD), rather than detailed evaluation of each of the numerous newly available
29      studies.  Particular emphasis is focused in the text on those studies and analyses thought to
30      provide the most pertinent information for U.S. standard setting purposes. For example, North
31      American studies conducted in the U.S. or Canada are generally accorded more text discussion

        April 2002                                 8-11         DRAFT-DO NOT QUOTE OR CITE

-------
 1      than those from other geographic regions; and analyses using gravimetric (mass) measurements
 2      are generally accorded more text attention than those using non-gravimetric ambient PM
 3      measures, e.g., black smoke (BS) or coefficient of haze (COH). Also, more emphasis is placed
 4      on text discussion of new multi-city studies that employ standardized methodological analyses
 5      for evaluating PM effects across several or numerous cities and often provide overall effects
 6      estimates based on combined analyses of information pooled across multiple cities.
 7           In the sections that follow on PM mortality and morbidity effects, key points derived from
 8      the 1996 PM AQCD assessment of then-available information are first concisely highlighted.
 9      Succinct summary tables are included and key information is discussed below in the main text
10      with regard to the most important numerous new studies that have become available since that
11      prior PM AQCD. More detailed information for these and other newly available studies is
12      summarized in tabular form in Appendices 8A and 8B, in which important methodological
13      features and results are presented.  The Appendix tables have a uniform general organization
14      with divisions that include:  (1) information about study location and ambient PM levels,
15      (2) study description of methods employed, (3) results and comments and (4) quantitative
16      outcomes for PM measures.
17
18
19      8.2 MORTALITY EFFECTS OF PARTICIPATE MATTER EXPOSURE
20      8.2.1 Introduction
21           The relationship of PM and other air pollutants to excess mortality has been intensively
22      studied and has been an important issue addressed in previous PM criteria assessments (U.S.
23      Environmental Protection Agency, 1986, 1996a). Mortality is the most severe adverse health
24      endpoint and, in some ways, the easiest to study. Excellent death records are maintained at every
25      level of government in most all nations and are typically made available to researchers.  Also,
26      from a narrowly technical point of view, individual deaths are more amenable to statistical
27      analyses, since individual deaths from natural causes (typically respiratory and cardiovascular
28      diagnoses) are statistically independent, except in rare extremely infectious instances. Individual
29      deaths are also non-recurring events, unlike hospital admissions or respiratory symptoms.
30

        April 2002                                8-12        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Recent findings are evaluated here for the two most important epidemiology designs by
 2      which mortality is studied:  time-series mortality studies (Section 8.2.2) and prospective cohort
 3      studies (Section 8.2.3). The time-series studies mostly assess acute responses to short-term PM
 4      exposure, although some recent work suggests that time-series data sets are also useful to
 5      examine responses to exposures over a longer time scale. Time-series studies use community-
 6      level air pollution measurements to index exposure and  community-level response (i.e., the total
 7      number of deaths each day by age and/or by cause of death).  Prospective cohort studies usefully
 8      complement time-series studies; they use individual health records, with survival lifetimes or
 9      hazard rates adjusted for individual risk factors, and typically evaluate human health impacts of
10      long-term PM exposures indexed by community-level measurements.
11
12      8.2.2 Mortality Effects of Short-Term Particulate Matter Exposure
13      8.2.2.1 Summary of 1996 Particulate Matter Criteria Document Findings and Key Issues
14           The time-series mortality studies reviewed in the 1996 and other past PM AQCD's
15      provided much evidence that ambient PM air pollution is associated with increases in daily
16      mortality. The 1996 PM AQCD assessed about 35 PM-mortality time-series studies published
17      between 1988 and 1996.  Information derived from those studies was consistent with the
18      hypothesis that PM is a causal agent in the short-term mortality impacts of air pollution.
19           The PM10 relative risk estimates derived from short-term PM10 exposure studies reviewed
20      in the 1996 PM AQCD suggested that an increase of 50 //g/m3 in the 24-h average of PM10 is
21      most clearly associated with an increased risk of premature total nonaccidental mortality (total
22      deaths minus those from accident/injury) on the order of relative risk (RR) = 1.025 to 1.05 in the
23      general population or, in other words, 2.5 to 5.0% excess deaths per 50 //g/m3 PM10 increase.
24      Higher relative risks were indicated for the elderly and for those with pre-existing
25      cardiopulmonary conditions. Also, based on the then recently published Schwartz et al. (1996a)
26      analysis of Harvard Six City data, the 1996 PM AQCD found the RR for excess total mortality in
27      relation to 24-h fine particle concentrations to be in the range of RR = 1.026 to 1.055 per
28      25 //g/m3 PM2 5 (i.e., 2.6 to  5.5% excess risk per 25 //g/m3 PM2 5 increment).
29           While numerous studies reported PM-mortality associations, important issues needed to be
30      addressed in interpreting their findings.  The 1996 PM AQCD extensively discussed most critical
31      issues, including: (1) seasonal confounding and effect modification;  (2) confounding by weather;
        April 2002                                8-13        DRAFT-DO NOT QUOTE OR CITE

-------
 1      (3) confounding by co-pollutants; (4) measurement error; (5) functional form and threshold;
 2      (6) harvesting and life shortening; and (7) the role of PM components.  As important issues
 3      related to model specification became further clarified, more studies began to address the most
 4      critical issues, with some having been at least partially resolved, whereas others required still
 5      further investigation.  The next several paragraphs summarize the status of these issues at the
 6      1996 PM AQCD publication time.
 7           One of the most important components in time-series model specification is adjustment for
 8      seasonal cycles and other longer-term temporal trends. Residual over-dispersion and
 9      autocorrelation result from inadequate control for these temporal trends, and not adequately
10      adjusting for them could result in biased RRs.  Modern smoothing methods allow efficient fits of
11      temporal trends and minimize such statistical problems. Thus, most recent studies controlled for
12      seasonal and other temporal trends, and it was unlikely that inadequate control  for such trends
13      seriously biased estimated PM coefficients. Effect modification by  season was examined in
14      several studies.  Season-specific analyses are often not feasible in small-sized studies (due to
15      marginally significant PM effect size), but some studies (e.g., Samet et al., 1996; Moolgavkar
16      and Luebeck, 1996) suggested that estimated PM coefficients varied from season to  season.
17      It was not fully resolved, however, if these results represent real seasonal effect modifications or
18      may be due to varying extent of correlation between PM and co-pollutants or weather variables
19      by season.
20           While most available studies included control for weather variables, some reported
21      sensitivity of PM coefficients to weather model specification, leading some investigators to
22      speculate that inadequate weather model specifications may still have erroneously ascribed
23      residual weather effects to PM. Two PM studies (Samet et al., 1996, 1998; Pope and Kalkstein,
24      1996) involved collaboration with a meteorologist and utilized more elaborate  weather modeling,
25      e.g., use of synoptic weather categories.  These studies found that estimated PM effects were
26      essentially unaffected by the synoptic weather variables and also indicated that the synoptic
27      weather model did not provide better model fits in predicting mortality when compared to other
28      weather model specifications used in previous PM-mortality studies. Thus, these results
29      suggested that the reported PM effects were not explained by weather effects.
30           Many earlier PM studies considered at least one co-pollutant in the mortality regression,
31      and some also examined several co-pollutants. In most cases, when PM indices were significant

        April 2002                                 8-14        DRAFT-DO NOT  QUOTE OR CITE

-------
 1      in single pollutant models, addition of a co-pollutant diminished the PM effect size somewhat,
 2      but did not eliminate the PM associations. When multiple pollutant models were performed by
 3      season, the PM coefficients became less stable, again, possibly due to PM's varying correlation
 4      with co-pollutants among season and/or smaller sample sizes. However, in many studies, PM
 5      indices showed the highest significance (versus gaseous co-pollutants) in single and multiple
 6      pollutant models.  Thus, it was concluded that PM-mortality associations were not seriously
 7      distorted by co-pollutants, but interpretation of the relative significance of each pollutant in
 8      mortality regression as relative causal strength was difficult because of limited quantitative
 9      information on relative exposure measurement/characterization errors among air pollutants.
10           Measurement error can influence the size and significance of air pollution coefficients in
11      time-series regression analyses and is also important in assessing confounding among multiple
12      pollutants, as varying the extent  of such error among the pollutants could also influence the
13      corresponding relative significance.  The 1996 PM AQCD discussed several types of such
14      exposure measurement or characterization errors, including site-to-site variability and site-to-
15      person variability—errors thought to bias the estimated PM coefficients downward in most cases.
16      However, there was not sufficient  quantitative information available to estimate such bias.
17           The 1996 PM AQCD also  reviewed evidence for threshold and various other functional
18      forms of short-term PM mortality associations. Several studies indicated that associations were
19      seen monotonically below the existing PM standards. It was considered difficult, however, to
20      statistically identify a threshold from available data because of low data density at lower ambient
21      PM concentrations, potential influence of measurement error, and adjustments for other
22      covariates. Thus, the use of relative risk (rate ratio) derived from the log-linear Poisson models
23      was considered adequate and appropriate.
24           The extent of prematurity of death (i.e., mortality displacement,  or harvesting) in observed
25      PM-mortality associations has important public health policy implications.  At the time of the
26      1996 PM AQCD review, only a  few studies had investigated this issue. While one of the studies
27      suggested that the extent of such prematurity might be only a few days, this may not be
28      generalizable because this estimate was obtained for identifiable PM episodes. There was not
29      sufficient evidence to suggest the extent of prematurity for non-episodic periods, from which
30      most of the recent PM relative risks were derived. The 1996 PM AQCD concluded:


        April 2002                                 8-15        DRAFT-DO NOT QUOTE OR CITE

-------
 1           "In summary, most available epidemiologic evidence suggests that increased mortality
 2           results from both short-term and long-term ambient PM exposure.  Limitations of available
 3           evidence prevent quantification of years of life lost to such mortality in the population.  Life
 4           shortening, lag time, and latent period of PM-mediated mortality are almost certainly
 5           distributed over long time periods, although these temporal distributions have not been
 6           characterized." (p. 13-45)
 7           Only a limited number of PM-mortality studies analyzed fine  particles and chemically
 8      specific components of PM. The Harvard Six Cities Study (Schwartz et al., 1996a) analyzed
 9      size-fractionated PM (PM2 5, PM10/15, and PM10/15.2 5) and PM chemical components (sulfates and
10      H+).  The results suggested that PM25 was most significantly associated with mortality among the
11      components of PM. While H+ was not significantly associated with mortality in this and an
12      earlier analysis (Dockery et al., 1992), the smaller sample size for H+ than for other PM
13      components made a direct comparison difficult.  The 1996 PM AQCD also noted that mortality
14      associations with BS or COH reported in earlier studies in Europe and the U.S. during the 1950s
15      to 1970s most likely reflected contributions from fine particles, as those PM indices had low 50%
16      cut-off diameters (« 4.5yam).  Furthermore, certain respiratory morbidity studies showed
17      associations between hospital admissions/visits with components of PM in the fine particle
18      range.  Thus, the U.S. EPA 1996 PM AQCD concluded that there was adequate evidence to
19      suggest that fine particles play especially important roles in observed PM mortality effects.
20           Overall, then, the status of key issues raised in the  1996  PM AQCD can be summarized as
21      follows: (1) the observed PM effects are unlikely to be seriously biased by inadequate statistical
22      modeling (e.g., control for seasonality); (2) the observed PM effects are unlikely to be
23      significantly confounded by weather; (3) the observed PM effects may be to some extent
24      confounded or modified by co-pollutants, and such extent may vary from season to season;
25      (4) determining the extent of confounding and effect modification by co-pollutants requires
26      knowledge of relative exposure measurement characterization error among pollutants (there was
27      not sufficient information on this); (5) no clear evidence for any threshold for PM-mortality
28      associations was reported (statistically identifying a threshold  from  existing data was also
29      considered difficult, if not impossible); (6) some limited evidence for harvesting, a few days of
30      life-shortening, was reported for episodic periods (no study was conducted to investigate
31      harvesting in non-episodic U.S. data); (7) only a relatively limited number of studies suggested a

        April 2002                                8-16        DRAFT-DO NOT QUOTE OR CITE

-------
 1      causal role of fine particles in PM-mortality associations, but in the light of historical data,
 2      biological plausibility, and the results from morbidity studies, a greater role for fine particles than
 3      coarse particles was suggested in the 1996 PM AQCD as being likely.  The AQCD concluded:
 4           "The evidence for PM-related effects from epidemiologic studies is fairly strong, with most
 5           studies showing increases in mortality, hospital admissions, respiratory symptoms, and
 6           pulmonary function decrements associated with several PM indices. These epidemiologic
 7           findings cannot be wholly attributed to inappropriate or incorrect statistical methods,
 8           misspecification of concentration-effect models, biases in study design or implementation,
 9           measurement of errors in health endpoint, pollution exposure, weather, or other variables,
10           nor confounding of PM effects with effects of other factors. While the results of the
11           epidemiology studies should be interpreted cautiously, they nonetheless provide ample
12           reason to be concerned that there are detectable human health effects attributable to PM at
13           levels below the current NAAQS." (p. 13-92)
14
15      8.2.2.2  Introduction to Newly Available Information on Short-Term Mortality Effects
16           Since the 1996 PM AQCD, numerous new studies have examined short-term associations
17      between PM indices and mortality. Newly available U.S. and Canadian studies on relationships
18      between short-term PM  exposure and daily mortality are summarized in Table 8-1.  More
19      detailed summaries of these and of other short-term exposure PM-mortality studies from other
20      geographic areas (e.g., Europe, Asia, etc) are described in Appendix Table 8A-1.  Information on
21      study location, study period, levels of PM, outcomes, methods, results, and reported risk
22      estimates and lags is provided  in Table 8A-1. In addition to these summary tables, discussion in
23      the text below highlights findings from several multi-city studies. Discussion of implications of
24      new study results for types of issues identified in foregoing text is mainly deferred to Section 8.4.
25           The summarization of studies in Table 8-1 and 8A-1 (and in other tables) is not meant to
26      imply that all listed studies should be accorded equal weight in the overall interpretive
27      assessment of evidence regarding PM-associated health effects. In general, increasing scientific
28      weight should be accorded to those studies (i.e., those not clearly flawed and which have
29      adequate control for confounding) in proportion to the precision of their estimate  of a health
30      effect.  Small studies and studies with an inadequate exposure gradient generally produce less
31      precise estimates than large studies with an adequate exposure gradient.  Therefore, the range of
32      exposures (e.g., as indicated by the IQR), the size of the study as indexed by the total number of
        April 2002                                 8-17        DRAFT-DO NOT QUOTE OR CITE

-------
            TABLE 8-1. RECENT U.S. AND CANADIAN TIME-SERIES STUDIES OF PM-RELATED DAILY MORTALITY*
to
o
o
to
         Reference
                                          Location(s)
                                                                                 Pollutants in Models
                Comments
Multi-City Mortality Studies in the U.S. and Canada
oo
oo
fe
H
6
o
o
H
O
O
H
W
O
         PM10 studies using NMMAPS data
         Samet et al. (2000a,b,c);
         Dominici et al. (2000a,b);
         Samet (2000)
         Daniels et al. (2000)
         Dominici et al. (2002)
                             88 cities in the 48 contiguous U.S. states
                             plus AK and HI, 1987-1994; mainly 20
                             largest.
                             20 cities in the 48 contiguous U.S. states,
                             1987-1994
                             88 cities in the 48 contiguous U.S. states,
                             1987-1994
                                                                             PM10, O3, CO, NO2, SO2
                                                                             PM10 only
                                                                             PM10 only
         Braga et al. (2000)
                             Five large U.S. cities: Chicago, IL;
                             Detroit, MI; Pittsburgh PA,
                             Minneapolis-St. Paul, MN; Seattle, WA
                                                                             PM,n only
Numerous models; range of PM10 values
depending on city, region, co-pollutants.
Pooled estimates for 88 cities, individual
estimates for 20 largest with co-pollutant
models

Smooth non-parametric spline model for
concentration-response functions. Average
response curve nearly linear.

Smooth non-parametric spline models for PM10
concentration-response functions. Average
response curves are nearly linear in the
industrial Midwest and Northeast regions, and
overall, but non-linear (usually concave) in the
other regions. Possible thresholds in
Southwest, Southeast.

Pooled estimate across cities adjusted for
influenza epidemics.
         Brief summary of new time-series studies on daily mortality since the 1996 Air Quality Criteria Document for Paniculate Matter (U.S. Environmental
         Protection Agency, 1969a).  More complete descriptive summaries are provided in Appendix Table 8A-1.  The endpoint is total daily non-trauma mortality
         unless noted otherwise. Due to the large number of models reported for sensitivity analyses for some of these papers, some evaluating various lags and
         co-pollutant models, some for individual cities and others for estimates pooled across cities, quantitative risk estimates are not presented in this table.
         Specific mortality risk estimates for fine and coarse particle models are shown in Table 8-2. Multiple-pollutant models are discussed in Section 8.4.2.2.
O

-------
                                TABLE 8-1 (cont'd).  RECENT U.S. AND CANADIAN TIME-SERIES STUDIES
                                                      OF PM-RELATED DAILY MORTALITY
to
o
o
to
Reference
             Location(s)
Pollutants in Models
                                                                                                                         Comments
         Multi-City Mortality Studies in the U.S. and Canada (cont'd)
oo

VO
fe
H
6
o
o
H
O
O
H
W
O
         Studies using everyday PM10 data
         Schwartz (2000b)
         Schwartz and Zanobetti
         (2000)
Zanobetti and Schwartz
(2000)


Moolgavkar (2000a)
         Popeetal. (1999a)


         Laden et al. (2000)
                            Same ten U.S. cities as in
                            (Schwartz, 2000a)


                            Same ten U.S. cities as in
                            (Schwartz, 2000a)
Four large U.S. cities: Chicago, IL;
Detroit, MI, Minneapolis-St. Paul, MN;
Pittsburgh, PA

Three large U.S. counties (cities):
Cook City (Chicago), IL; Los Angeles,
CA; Maricopa Cty. (Phoenix), AZ.
                            Ogden, Provo-Orem, and Salt Lake City,
                            UT.

                            Same six cities as in Harvard Six city
                            study, with Harvard air monitors and
                            community daily mortality time series:
                            Boston (Watertown), MA, Harriman-
                            Kingston, TN; Portage-Madison, WI;
                            St. Louis, MO; Steubenville, OH;
                            Topeka, KS.
                                       PM10 only.
                                       PM10 only.
                                                                            PM,n only.
                                                                            PM10 in all three; PM2 5 in
                                                                            Los Angeles. O3, CO, NO2,
                                                                            and SO, in some models.
                                       PM10 only in all three.
                                       Chemically speciated PM2 5,
                                       and factors aligned with
                                       putative sources for each
                                       city identified by specific
                                       chemical elements as tracers.
                          Several pooled estimates across cities evaluated
                          for single day, moving average, and distributed
                          lags.

                          Pooled estimates of concentration-response
                          functions across cities using smooth semi-
                          parametric functions of PM10 with the same
                          span of 0.

                          Pooled estimate of effect size across cities was
                          modified somewhat by race and gender.


                          The results showed little consistency for
                          different time lags and cities, the PM10 or PM2 5
                          effects on CVD mortality were greatly
                          attenuated by including one or more gaseous
                          co-pollutants

                          Positive, significant and similar effects for
                          PM10 on total, CVD, and respiratory mortality

                          Different coefficients in different cities,
                          depending on source type, chemical indicators,
                          and principal factor method. The motor vehicle
                          combustion component was significant, other
                          factors occasionally, but not the crustal element
                          component.
O

-------
to
o
o
to
                                TABLE 8-1 (cont'd). RECENT U.S. AND CANADIAN TIME-SERIES STUDIES
                                                      OF PM-RELATED DAILY MORTALITY
Reference
             Location(s)
    Pollutants in Models
                                                                                                                        Comments
         Multi-City Mortality Studies in the U.S. and Canada
oo
to
o
         Tsaietal. (1999, 2000)
         Clyde et al. (2000)
        Burnett et al. (2000)
Burnett etal. (1998a)
                            Camden, Elizabeth, and Newark, NJ.
                            Phoenix, AZ, May, 1995-March, 1998.
                            Seattle, WA, 1990-1995.
                                       PM2 5, PM15, sulfates.
                                       PM2 5, PM10_2 5 in Phoenix.
                                       PM10, PM2 5, nephelometer,
                                       SO, in Seattle.
                            Eight Canadian cities: Montreal, Ottawa,    PM10, PM2 5, PM10_2 5, SO4,
                            Toronto, Windsor, Calgary, Edmonton,     O3, CO, NO2, SO2
                            Winnipeg, Vancouver
Eleven Canadian cities.  1980-1991.
Main emphasis on O3, CO,
NO2, SO2. PM2 5, PM10.2 5,
SO4 on varying schedules.
                                                                                                         Significant effects of PM25, PM10, and sulfates
                                                                                                         in Newark, Camden at most lags, but not
                                                                                                         Elizabeth.

                                                                                                         PM10_2 5 significant in most of the 25 "best"
                                                                                                         models for Phoenix, PM2 5 in almost none.
                                                                                                         PM2 5 and PM10 in some models for Seattle,
                                                                                                         none in the 5 best.

                                                                                                         Significant effects of PM2 5 and PM10, less so
                                                                                                         for PM10_2 5; particle effects stable, co-pollutant
                                                                                                         effects decreased by particles

                                                                                                         Qualitative indication of effect modification
                                                                                                         of gaseous pollutant effects by particles.
fe
H
6
o
o
H
O
O
H
W
O
        Klemm et al. (2000)


        Schwartz et al. (2002)
                            Same six cities as (Laden et al., 2000)
                            1979-1988.

                            Same six cities as (Laden et al., 2000)
                            1979-1988
                                       PM10, PM2.5, PM10.2.5, S04
                                       PM25, PM10.25, 15 elements in
                                       PM2 5, O3, CO, NO2, SO2
                             Replicated Schwartz et al. (1996a) with
                             additional sensitivity analyses.

                             Five source factors identified, as in (Laden
                             et al., 2000).  Meta-smoothing of
                             non-parametric concentration-mortality curves
                             for PM2 5 and for five source factors.  Total and
                             "traffic" source PM2 5 significantly associated
                             with mortality, nearly linear for PM2 5, steeper
                             slope at low concentrations of traffic particles.
                             No apparent threshold.
O

-------
                                TABLE 8-1 (cont'd).  RECENT U.S. AND CANADIAN TIME-SERIES STUDIES
                                                       OF PM-RELATED DAILY MORTALITY
to
o
o
to
Reference
             Location(s)
                                                                                 Pollutants in Models
                Comments
oo
to
fe
H
6
o
o
H
O
O
H
W
O
         Single-City Mortality Studies in the U.S. and Canada

         Ostro et al. (1999a, 2000)      Coachella Valley (Palm Springs), CA
         Fairley (1999)
         Schwartz etal. (1999)
Schwartz and Zanobetti
(2000)


Lippmann et al. (2000)
         Chock et al. (2000)
                             Santa Clara County (San Jose), CA
                             Spokane, WA
Chicago, IL
                                     Detroit, MI
                             Pittsburgh, PA
                                                                    PM10 in earlier study, PM2 5
                                                                    and PM10_2 5 in later study; O3,
                                                                    CO,NO2
                                       PM10, PM2 5, PM10.2 5, sulfates,
                                       nitrates, O3, CO, NO2.
                                       PM10 only
                                                                             PM,n only
                                       PM10, PM2 5, PM10.2.5,
                                       sulfates, acidity, TSP, O3,
                                       CO, NO2, SO2
                                       PM10,PM25,PM10.25,03,
                                       CO, NO2, SO2
PM2 5 effects significant, PM10 and PM
effects non-significant for total mortality; for
cardiovascular mortality, PM10 and PM10_2 5
significant, PM2 5 not

All significant in one-pollutant models, nitrates
significant in all multi-pollutant models, PM2 5
significant except with particle nitrates.

No association between mortality and high
PM10 concentrations on dust storm days with
high crustal particles.

Larger effects with longer-term PM10 and
mortality moving averages for total, in-hospital,
and out-of-hospital mortality.

Positive but non-significant effects on mortality
for the 1992-1994 data,  but significant effects
for respiratory mortality vs. PM10 or TSP in
1985-1990 data.

Fine and coarse particle data on about 1/3 of
days with PM10. Data split into ages < 75 and
75+, and seasons. Significant effects for PM10,
not for size fractions. Regional sulfate, traffic-
related PM, and biogenic combustion factors
have maximum associations on different lag
days.
O

-------
                                TABLE 8-1 (cont'd). RECENT U.S. AND CANADIAN TIME-SERIES STUDIES
                                                      OF PM-RELATED DAILY MORTALITY
to
o
o
to
Reference
             Location(s)
    Pollutants in Models
                Comments
oo
to
to
H
6
o
*
o
H
O
O
H
W
O
         Single-City Mortality Studies in the U.S. and Canada

         Klemm and Mason (2000)     Atlanta, GA
         Gwynn et al. (2000)
         Schwartz (2000c)
Lipfert et al. (2000a)
         Levy (1998)
         Mar etal. (2000)
                            Buffalo, NY
                            Boston, MA
Philadelphia, PA-Camden, NJ seven-
county area
                            King County (Seattle), WA
                            Phoenix, AZ, near the EPA platform
                            monitor.
                                       PM2 5, PM10_2 5, nitrate,
                                       oxygenated hydrocarbons
                                       (HC), elemental carbon (EC),
                                       organic carbon (OC)

                                       PM10, CoH, H+, SO4, O3, CO,
                                       NO2, SO2.
                                       PM,
PM10, PM2 5, PM10.2 5, sulfates,
acids, metals, O3, CO, NO2,
S02.
                                           (nephelometer), PM10,
                                       CO, SO,.
                                       PM10, PM2 5, PM10.2 5, fine
                                       particle elements, estimated
                                       soil and non-soil PM, EC,
                                       OC, O3, CO, NO2, SO2;
                                       sources by factor scores.
                                                                                                No significant effects due to short time series,
                                                                                                ca. one year. Larger effect and shorter
                                                                                                confidence interval for PM2 5 than for PM10_2 5.
All PM components significantly associated
with total mortality in single-pollutant models,
not gaseous pollutants.

Larger effects with longer-term PM2 5 and
mortality moving averages (span 15 to 60 days)
for total and cause-specific mortality.

Exploration of mortality in different areas
relative to air monitor location. Peak O3 very
significant, greatly reduces PM effects.

PMj associated only with out-of-hospital
ischemic heart disease deaths, total mortality
with neither PM10 nor PM[

Total mortality significantly associated with
NO2, CO, weakly with PM10, PM10.2 5, EC, SO2.
Cardiovascular mortality significantly
associated with PM10, PM2 5, PM10.2 5, EC, OC,
CO, NO2, SO2, source factors.
O

-------
to
o
o
to
                                TABLE 8-1 (cont'd).  RECENT U.S. AND CANADIAN TIME-SERIES STUDIES
                                                     OF PM-RELATED DAILY MORTALITY
Reference
             Location(s)
Pollutants in Models
Comments
         Single-City Mortality Studies in the U.S. and Canada

         Clyde et al. (2000)            Phoenix, AZ
         Smith et al. (2000)
                            Phoenix, AZ (within city and within
                            county), 1995-1997.
                                                                  PM10, PM2 5
                                      PM25,PM10.2
oo
to
Gamble (1998)
Ostro (1995)
Dallas, TX
1990-1994
San Bernar
CA
1980-1986.
                                                                           PM10, O3, CO, NO2, SO2
                                     San Bernardino and Riverside Counties,    PM2 5 estimated from visual
                                                                           range,
                                                                           O3
                         Effect on elderly mortality consistently higher
                         for PM10_2 5 among 25 "best" models.  Estimates
                         combined using Bayesian model averaging.

                         Significant linear relationship with PM10_2 5, not
                         PM2 5. Piecewise linear models with possible
                         PM10.2 5 threshold for elderly mortality 20-25
                                                                                               O3, CO, NO2 significantly associated with
                                                                                               mortality, PM10 and NO2 not associated

                                                                                               Positive, significant PM2 5 effect only in
                                                                                               summer
         Kelsall et al. (1997)
Philadelphia, PA
1974-1988
                                                                  TSP, SO2, NO2, O3, CO
                         TSP, O3, CO, NO2 significant alone, TSP effect
                         reduced when SO, included.
O
O

O
H
O
H
W
O
Moolgavkar and Luebeck      Philadelphia, PA 1973 -1988
(1996)

Murray and Nelson (2000)     Philadelphia, PA, 1973-1990
                                      TSP, O3, NO2, SO2.


                                      TSP only
                         NO2 most significant pollutant, TSP effects
                         stronger in summer and fall.

                         Kalman filtering used to estimate hazard
                         function in a state space model. Both TSP and
                         the product of TSP and average temperature are
                         significant, but not together. Includes estimate
                         of at-risk population.
O

-------
to
o
o
to
                               TABLE 8-1 (cont'd).  RECENT U.S. AND CANADIAN TIME-SERIES STUDIES
                                                     OF PM-RELATED DAILY MORTALITY
Reference
Location(s)
Pollutants in Models
                                                                                                                        Comments
oo
to
fe
H
6
o
o
H
O
O
H
W
O
         Single-City Mortality Studies in the U.S. and Canada

         Neasetal. (1999)
         Schwartz (2000d)
        Burnett etal. (1998b)
                            Philadelphia, PA
                            1973-1980

                            Philadelphia, PA
                            1974-1988
                            Toronto, ON, Canada 1980-1994
Goldberg et al. (2001a,b,c,d)   Montreal, PQ, Canada, 1984-1995
         Ozkaynaketal. (1996)
                            Toronto, ON, Canada 1970-1991
                          TSP only


                          TSP, SO2, humidity-
                          corrected extinction
                          coefficient

                          TSP, CoH, SO4=, CO, NO2,
                          SO2, O3,PM10andPM25
                          estimated from every-sixth-
                          day data and observed daily
                          SO4=, TSP, and CoH

                          PM2 5 and PM10 every sixth
                          day until 1992, daily through
                          1993.  CO, NO2, NO, O3,
                          SO2. Missing PM data
                          estimated from sulfates, CoH,
                          extinction coefficient.
                          TSP, CoH, O3, CO, NO2, SO2
                         Case-crossover study. Significant TSP  effect.


                         No SO2 effect when TSP in model.  TSP
                         significant unless extinction coefficient in
                         model.

                         Significant excess total mortality for PM2 5,
                         PM10, TSP
                         Excess total and cause-specific mortality with
                         most PM indices reported (estimated PM2 5,
                         sulfates, CoH). In the age 65+ age group, total
                         mortality significantly elevated in individuals
                         with prior cancer, acute lower resp. disease,
                         any cardiovascular disease, chronic coronary
                         artery disease, congestive heart failure.

                         Significant association with 0-day lag TSP.
                         Factor analysis identified a factor with high
                         loadings on CoH, CO, and NO2 (traffic
                         presumably) significantly associated with total
                         and most cause-specific deaths.	
O

-------
 1      observations (e.g., days) and total number of events (i.e., total deaths), and the inverse variance
 2      for the principal effect estimate are all important indices useful in determining the likely
 3      precision of health effects estimates and in according relative scientific weight to the findings of
 4      a given study.
 5           As can be seen in Tables 8-1 and 8A-1, with a few exceptions, nearly all of the newly
 6      reported analyses continue to show statistically significant associations between short-term (24 h)
 7      PM exposures indexed by a variety of ambient PM measurements and increases in daily mortality
 8      in numerous U.S. and Canadian cities, as well as elsewhere around the world. Also, the effects
 9      estimates from the newly reported studies are generally consistent with those derived from the
10      earlier 1996 PM AQCD assessment, with the newly reported PM risk estimates generally falling
11      within the range of ca. 1 to 8% increase in excess deaths per 50 //g/m3 PM10 and ca. 2 to 6%
12      increase per 25//g/m3 PM2 5. Several newly available PM epidemiology studies which
13      conducted time-series analyses in multiple cities are of particular interest, as discussed below.
14
15      8.2.2.3 New Multi-City Studies
16           The new multi-city studies are of particular interest here due to their evaluation of a wide
17      range of PM exposures and large numbers of observations holding promise of providing more
18      precise effects estimates than most smaller scale independent studies of single cities.  Another
19      major advantage of the multi-city studies, over meta-analyses for multiple "independent" studies,
20      is the consistency in data handling and model specifications, which eliminates variation due to
21      study design.  Further, unlike regular meta-analysis, they clearly  do not suffer from potential
22      omission of negative studies due to "publication bias". Furthermore, geographic patterns of air
23      pollution effects can be systematically evaluated in multiple-city analyses. Thus, the results from
24      multi-city studies can provide especially valuable evidence regarding the consistency and/or
25      heterogeneity, if any, of PM-health effects relationships across geographic locations.  Also, many
26      of the cities included in these multi-city studies were ones for which no time-series analyses had
27      been previously reported.
28
29
30


        April 2002                                 8-25        DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.2.2.3.1  U.S. Multi-City Studies
 2      U.S. PM10 20-Cities and 90-Cities NMMAPS Analyses
 3           The National Morbidity, Mortality, and Air Pollution Study (NMMAPS) focused on time-
 4      series analyses of PM10 effects on mortality during 1987-1994 in the 90 largest U.S. cities (Samet
 5      et al., 2000a,b), in the 20 largest U.S. cities in more detail (Dominici et al., 2000a), and PM10
 6      effects on emergency hospital admissions in 14 U.S. cities (Samet et al., 2000a,b). These
 7      NMMAPS analyses are marked by extremely sophisticated statistical approaches addressing
 8      issues of measurement error biases, co-pollutant evaluations, regional spatial correlation, and
 9      synthesis of results from multiple cities by hierarchical Bayesian meta-regressions and
10      meta-analyses. These analyses provide extensive new information of much importance in being
11      among that most highly relevant to the setting of U.S. PM standards, because no other study has
12      examined as many U.S. cities in such a consistent manner. NMMAPS used only one consistent
13      PM index (PM10) across all cities (noted PM10 samples were only collected every 6 days in most
14      of the 90 cities); death records were collected in a uniform manner; and demographic variables
15      were uniformly addressed. Both the 20 and 90 cities analyses studies employ multi-stage models
16      (see Table 8-1) in which heterogeneity in individual cities' coefficients in the first stage GAM
17      Poisson models were evaluated in the second stage models with city or region specific
18      explanatory variables.
19           In both the  20 and 90 cities studies, the combined estimates of PM10 coefficients were
20      positively associated with mortality at all the lags examined (0, 1, and 2 day lags), although the
21      1-day lag PM10 resulted in the largest overall combined estimate.  Figure 8-3 shows the estimated
22      percent excess total deaths per 10 //g/m3 PM10 at lag 1 day in the 88 (90 minus Honolulu and
23      Anchorage) largest cities,  as well as (weighted average) combined estimates for U.S. geographic
24      regions depicted  in Figure 8-4.  The majority of the coefficients were positive for the various
25      cities listed along the left axis of Figure 8-3. The estimates for the individual cities were first
26      made independently, without borrowing information from other cities. The cities were then
27      grouped into the 7 regions seen in Figure 8-4 (based on characteristics of the ambient PM mix
28      typical of each region,  as delineated in the 1996 PM AQCD). The bolded segments represent the
29      posterior means and 95% posterior intervals of the pooled regional effects under the more
30      conservative prior A for the heterogeneity across both regions and cities within regions. The
31      solid circles and squares denote, respectively, the overall regional means without and with

        April 2002                                 8-26        DRAFT-DO NOT QUOTE OR CITE

-------
           % Increase in Mortality per 10 |jg/m3 Increase in PM10


seat
sanj

denv


stoc
mode
spok
overall.1
°VeDhoe

sana
elpa
aust
albu
corp
lubb
overall.1
la

s a

rive
?s
overall.1
minn
wich


overall.1
chic
det
clev
pitt
i? 5

cine



akr

tole



overall.1

phil
bait

h t
dc
prov

syra



overall.1

nous
miam
atla
stpe
tamp
memp
orla
jckv
birm
char
nash
tuls
no
batr
Itrk
gmb
knox
?Sf
hunt
overall.1
rail-north
overall
-4 -2 (
I I










•




	 ° j









	 =^0 	


	 6























0 °



	 «
	 e —
fi —
—
	 9 	 * 	 5-
	 e 	
•
D 2 4
I I


0




> 	 o

-e 	
^r-°


n
tlT 	
^=

n
0 °
° n
(?
/P


>"

0 o




-e 	 5-





0 °





— **-2 —
— ^ 	 o

n °
	 e 	 	 	
	 5~e 	






^^^^^^^
^ 	
'- 	 — o 	
	 e 	
i 	
	 e 	
^
-------
                  Northwest
       Southern
       California
                                   Upper
                                  Midwest
Industrial
 Midwest
                                                                        Northeast
                            Southwest   1  Southeast
       Figure 8-4. Map of the United States showing the 90 cities (the 20 cities are circled) and
                 the seven regions considered in the NMMAPS geographic analyses. Regions:
                 Northwestern; Southern California; Southwest; Upper Midwest; Industrial
                 Midwest; Northeast; Southeast.
 1
 2
 3
 4
 5
 9
10
11
borrowing information from other regions, ("overall 1" = the regional mean without other
regions, "overall 2" = with information from other regions).  The triangles and bolded segments
at the bottom of Figure 8-3 display combined estimates of nationwide overall effects of PM10 for
all cities overall, and for all cities minus those in the Northeast (overall-north).
    Note that there  appears to be some regional-specific variation in the overall combined
estimates, shown as "overall 1" and "overall 2" for the two sets of modeling assumptions and
specifications used in analyses combining data from all the cities in a given region.  This can be
discerned more readily in Figure 8-5 (which depicts overall region-specific excess risk estimates
for day 0 and 2 day lags, as well as for lag 1 day). For example, the coefficients for the Northeast
are generally higher than for other regions (the Northeast combined estimate, 4.5% excess total
deaths per 50 //g/m3 increase in PM10, was about twice that for the 90-cities overall).  The  overall
       April 2002
                                      8-28
DRAFT-DO NOT QUOTE OR CITE

-------
                                                               >
CO
E
o 2-

a=
-&0-
ro.£
Ho
•- 0
0) C
w
CO
£ -1-
o '
_c
s£

^ ^ o* ,X £
X X 
^ 0^
i
!
tlf + t
i

i
T- CN O T- (M
O) O> O) D) D)
CO CO CO CO CO
en rn "fo "co Ja
                                                                                       0)
                                                                                       s
                                        0)
                                        §
       Figure 8-5. Percent excess mortality risk (lagged 0, 1, or 2 days) estimated in the
                   NMMAPS 90-City Study to be associated with 10-jUg/m3 increases in PM10
                   concentrations in cities aggregated within U.S. regions shown in Figure 8-2.
 1     national combined estimate (i.e., at lag 1 day, 2.3% excess total deaths per 50 //g/m3 increase in
 1     PM10) for the 90 cities is consistent with the range of estimates reported in the 1996 PM AQCD.
 3           In the 90 cities study, the weighted second-stage regression included five types of county-
 4     specific variables: (1) mean weather and pollution variables; (2) mortality rate (crude mortality
 5     rate); (3) sociodemographic variables (% not graduating from high school and median household
 6     income); (4) urbanization (public transportation); (5) variables related to measurement error
 7     (median of all pair-wise correlations between monitors).  Some of these variables were
 8     apparently correlated (e.g., mean PM10 and NO2, household income and education) so that the
 9     sign of coefficients in the regression changed when correlated variables were included in the
10     model.  Thus, while some of the county-specific variables were statistically significant (e.g.,
1 1     mean NO2 levels), interpreting the role of these county-specific variables may require caution.
12     Regarding the heterogeneity of PM10 coefficients, the investigators concluded that they "did not
13     identify any factor or factors that might explain these differences".
14           Another important finding from Samet and coworkers' analyses was the weak influence of
15     gaseous co-pollutants on the PM10 effect size estimates.  In both the 20 and 90 cities analyses,
16     PM10 coefficients changed little when O3 was added to regression models. Additions of a third
17     pollutant (i.e., PM10 + O3 + another gaseous pollutant) did reduce PM10 coefficients somewhat
       April 2002
8-29
DRAFT-DO NOT QUOTE OR CITE

-------
 1      (e.g., from -2.2 to ~ 1.7 per 50 //g/m3 PM10 at lag 1 day in the combined 90 cities analysis), but
 2      the PM10 coefficients remained statistically significant at p< 0.05. The gaseous pollutants
 3      themselves in single-, two-, and three-pollutant models were less consistently associated with
 4      mortality than PM10. Ozone was not associated with mortality using year-round data; but, in
 5      season-specific analyses, it was associated with mortality negatively in winter and positively in
 6      summer. SO2, NO2, and CO were weakly associated with mortality, but additions of PM10 and
 7      other gaseous pollutants did not always reduce their coefficients, possibly suggesting their
 8      independent effects. As noted in Section 8.1, CO and NO2 from motor vehicles are likely
 9      confounders of PM25 and, thus, of PM10 when it is not dominated by the coarse particle fraction.
10      The investigators concluded that the PM10 effect on mortality "did not appear to be affected by
11      other pollutants in the model".
12
13      U.S. 10-Cities Studies
14           In another set of multi-city analyses, Schwartz (2000a,b), Schwartz and Zanobetti (2000),
15      Zanobetti and Schwartz (2000), Braga et al. (2000), and Braga et al. (2001) analyzed 1987-1995
16      air pollution and mortality data from ten U.S. cities (New Haven, CT; Pittsburgh, PA;
17      Birmingham, AL; Detroit, MI; Canton, OH; Chicago, IL; Minneapolis-St. Paul, MN; Colorado
18      Springs, CO; Spokane, WA; and Seattle, WA.) or subsets (4  or 5 cities) thereof. The selection of
19      these cities was based on the availability of daily (or near daily) PM10 data.  The main results of
20      the study were presented in the Schwartz (2000a) paper and the other studies noted above
21      focused on each of several  specific issues, including:  potential confounding, effect modification,
22      distributed lag, and threshold.  In this section, the results for the Schwartz (2000a) main analyses
23      and that of Braga et al. (2000) on confounding are discussed, and results for analyses of other
24      specific issues are discussed later in appropriate sections. For each  of the 10 cities, daily total
25      (non-accidental) mortality was fitted using a GAM Poisson model adjusting for temperature,
26      dewpoint,  barometric pressure, day-of-week, season, and time.  Deaths stratified by location of
27      death (in or outside hospital) were also examined. The data were also analyzed by season
28      (November through April as heating season). In the second stage, the PM10 coefficients were
29      modeled as a function of city-dependent covariates including co-pollutant to PM10 regression
30      coefficient (to test potential confounding), education, unemployment rate, poverty level, and
31      percent non-white.  Threshold effects were also examined. The inverse variance weighted

        April 2002                                8-30        DRAFT-DO NOT QUOTE OR CITE

-------
 1      averages of the ten cities' estimates were used to combine results. PM10 was significantly
 2      associated with total deaths, and the effect size estimates were the same in summer and winter.
 3      Adjusting for other pollutants did not substantially change the PM10 effect size estimates.  The
 4      socioeconomic variables did not modify the estimates.  The effect size estimates for the deaths
 5      outside hospital were substantially greater than for inside hospital.  The combined percent excess
 6      death estimate for total mortality was 3.4% (95% CI: 2.7-4.1) per 50 //g/m3 increase in PM10, but
 7      was larger for days with PM10 < 50 //g/m3.
 8           Braga et al. (2000) evaluated potential confounding of the reported PM-mortality
 9      associations by effects of respiratory epidemics, using data from a subset of 5 of the 10 cities
10      evaluated by Schwartz (2000a). When adjustments were made for respiratory epidemics, small
11      decreases in PM10 effects were seen  in the cities evaluated.  The overall estimated percent excess
12      deaths per 50 //g/m3 PM10 for the five cities was 4.3% (CI 3.0, 5.6) without control for respiratory
13      epidemics, but slightly decreased to  4.0% (CI 2.6, 5.3) with control for epidemics.
14
15      U.S. 3-Cities Study
16           Moolgavkar (2000a) evaluated associations between short-term measures of major air
17      pollutants and daily deaths in three large U.S. metropolitan areas (Cook Co., IL, encompassing
18      Chicago; Los Angeles Co.,  CA; and Maricopa Co., AZ, encompassing Phoenix) during a 9-year
19      period (1987-1995).  Generalized additive models (GAM) were used in a standard manner to
20      conduct time-series Poisson regression analyses independently for each of the three cities
21      (allowing comparison of results across them not due to methodological differences), but no
22      combined analyses were attempted to derive overall PM effects  estimates. Total non-accidental
23      deaths and cause-specific deaths from cardiovascular disease (CVD), cerebrovascular disease
24      (CrD), and chronic obstructive lung  disease (COPD), and associated conditions were analyzed in
25      relation to 24-h readings for PM, O3, CO,  NO2, SO2 averaged over all monitors in a given county.
26      Daily readings were available for each of  the gaseous pollutants in all three countries, as were
27      PM10 values for Cook County.  However,  PM10 values were only available every sixth day in
28      Maricopa and Los Angeles  Counties; as were PM2 5 values in Los Angeles Co.  PM values were
29      highest in the winter and fall in Los  Angeles Co., in the fall in Maricopa Co., and in summer in
30      Cook Co., whereas the gases (except for O3) were highest in winter in all three counties (O3 was
31      highest in summer in all three). The PM indices were moderately correlated (r = 0.30 to 0.73)

        April 2002                                 8-31         DRAFT-DO NOT QUOTE OR CITE

-------
 1      with CO, NO2, and SO2 in Cook Co. and Los Angeles Co., but poorly correlated (r < 0.22) with
 2      those gases in Maricopa Co. Ozone was very poorly (r < 0.20) or negatively correlated with PM
 3      or the other gases in each location (except for Cook Co., r = 0.36 for O3 vs PM10).  Total
 4      non-accidental, CVD, and COPD deaths were all highest during winter in all three counties, but
 5      CrD deaths were relatively constant from season to season (no season-specific analyses reported).
 6           Controlling for temperature and relative humidity  effects in  separate analyses for each
 7      mortality endpoint for each of the three countries, varying patterns of results were found from
 8      one location to another, as noted in Table 8A-1. In general,  although PM10 in each of the three
 9      counties (and PM25 in Los Angeles) and each of the gaseous pollutants (except O3) were all
10      statistically significantly associated with total non-accidental mortality at one or more lag times
11      (0 to 5 days) in single pollutant models, the PM effect estimates tended to be reduced and non-
12      significant in many of the multi-pollutant (PM plus one  other gas  or PM plus all others) analyses.
13      In contrast, effect estimates for several of the gases (CO, SO2, and NO2) tended to be more  robust
14      than those for PM in multi-pollutant models, with their  estimates remaining statistically
15      significant (although usually somewhat attenuated) at one or more lag times when included in
16      multi-pollutant models with PM10 or PM2 5. Similarly, a somewhat analogous varying pattern of
17      results was observed for the cause-specific mortality analyses (discussed further below in Section
18      8.2.2.5). That is, although PM10 or PM25 were statistically significantly related to CVD and
19      COPD-related (and to CrD only in Maricopa Co., lag 5) mortality in single pollutant models,
20      their coefficients were typically markedly reduced and became non-significant in multi-pollutant
21      analyses with one or more of the gases included in the model. Moolgavkar (2000a) concluded
22      that, while direct effects of individual components of air pollution cannot be ruled out, individual
23      components can best be thought of as indices of the overall air pollution mix; and he noted
24      considerable heterogeneity of air pollution effects across the three geographic areas evaluated.
25      Moolgavkar (2000a) did not calculate any pooled effect estimates possibly because of the
26      heterogeneity seen among the cities studied.
27
28      8.2.2.3.2 Canadian Multi-City Study Analyses
29      Urban Air Pollution Mix and Daily Mortality in 11 Canadian  Cities
30           The number of daily deaths for non-accidental causes during 1980-1991 were obtained for
31      11 Canadian cities and linked to concentrations of ambient gaseous air pollutants using relative

        April 2002                                8-32        DRAFT-DO NOT QUOTE OR CITE

-------
 1      risk regression models for longitudinal count data (Burnett et al., 1998a). The GAM Poisson
 2      models used evaluated daily mortality versus O3, NO2, SO2 and CO (including adjustments for
 3      seasonal cycles, day-of-week effects, and weather effects), but no PM indices were included in
 4      their analyses because daily PM measurements were not available.  However, data were available
 5      for fine and coarse PM mass from dichot samples, and sulfates, on variable schedules somewhat
 6      more frequently than once per six days in Montreal, Toronto, and Windsor (with smaller
 7      numbers in the other cities).  This allowed an ecologic comparison of gaseous pollutant risks by
 8      mean fine particle concentration (their Figure 1). These comparisons suggested a weak negative
 9      confounding of NO2 and SO2 effects with fine particles, and a weak positive confounding of
10      particle effects with O3.
11
12      Eight Largest Canadian  Cities  Study
13           Burnett et al. (2000) analyzed various PM indices (PM10, PM25, PM10_2 5, sulfate, COH, and
14      47 elemental component concentrations for fine and coarse fractions) and gaseous air pollutants
15      (NO2, O3, SO2, and CO) for association with total mortality in the 8 largest Canadian cities:
16      Montreal, Ottawa-Hull,  Toronto, Windsor, Winnipeg, Calgary, Edmonton, and Vancouver. This
17      study differs from (Burnett et al., 1998a), including  fewer cities but more recent years of data
18      (1986-1996 vs. 1980-1991) and detailed  analyses of particle mass components by size and
19      elemental composition.  Each city's mortality, pollution, and weather variables were separately
20      filtered for seasonal trends and day-of-week patterns. The residual  series from all cities were
21      then combined and analyzed in a GAM Poisson model. The weather model was selected from
22      spline-smoothed functions of temperature, relative humidity, and maximum change in barometric
23      pressure within a day and with 0 and 1 day lags, using forward stepwise procedures. Pollution
24      effects were examined at lags 0 through 5 days. To avoid unstable parameter estimates in multi-
25      pollutant models, principal components were also used as predictors in the regression models.
26           Ozone was weakly correlated with  other pollutants, and other pollutants were "moderately"
27      correlated with each other (the highest was r = 0.65  for NO2 and CO). The strongest association
28      with mortality for all pollutants considered were for 0 or 1 day lags. PM2 5 was a stronger
29      predictor of mortality than PM10_25.  The  gaseous pollutant effects estimates were generally
30      reduced by inclusion of PM25 or PM10, but not PM10_2 5, where strength of prediction is measured
31      by the t value  or statistical significance of the excess risk. In addition to the results implicating

        April 2002                                 8-33        DRAFT-DO NOT QUOTE OR CITE

-------
 1      the fine particle fraction (PM2 5) most clearly, other findings on fine particle components were
 2      also of interest. Specifically, sulfate, Fe, Ni, and Zn were most strongly associated with
 3      mortality. The total effect of these four components was greater than that for PM2 5 mass alone,
 4      the authors suggesting that the characteristics of the complex chemical mixture in the fine
 5      fraction may be a better predictor of mortality than the mass index alone.
 6
 7      8.2.2.3.3 European Multi-City APHEA Study Analyses
 8           The Air Pollution and Health: a European Approach (APHEA) project is a multi-center
 9      study of short-term effects of air pollution on mortality and hospital admissions with a wide
10      range of geographic, climatic, sociodemographic, and air quality patterns. The obvious strength
11      of this approach is to be able to evaluate potential  effect modifiers in a consistent manner.
12      It should be noted that PM indices measured in those cities varied.  In APHEAl, the PM indices
13      measured were mostly black smoke (BS), except for: Paris, Lyon (PM13); Bratislava, Cologne,
14      and Milan (TSP); and Barcelnoa (BS and TSP). In APHEA2,  10 out of the 29 cities used actual
15      PM10 measurements; in 11 additional cities, PM10 levels were estimated based on regression
16      models relating collocated PM10 measurements to BS or TSP.  In the remaining 8 cities, only BS
17      measurements were available (14 cities had BS measurements). As discussed below, there have
18      been several papers published that present either a meta-analysis or pooled summary estimates of
19      these multi-city mortality results: (1) Katsouyanni et al. (1997) —  SO2 and PM results from
20      12 cities; (2) Touloumi et al. (1997) — ambient oxidants (O3 and NO2) results from six cities;
21      (3) Zmirou et al. (1998) — cause-specific mortality results from 10 cities (see Section 8.2.2.5);
22      (4) Samoli et al. (2001) — a reanalysis of APHEAl using a different model specification to
23      control for long-term trends and seasonality; and (5) Katsouyanni et al. (2001) — APHEA2, with
24      emphasis on the examination of confounding and effect modification.
25
26      APHEAl Sulfur Dioxide and Particulate Matter Results for 12 Cities
27           The Katsouyanni  et al. (1997) analyses evaluated data from the following cities:  Athens,
28      Barcelona, Bratislava, Cracow, Cologne, Lodz, London, Lyons, Milan, Paris, Poznan,  and
29      Wroclaw. In the western European cities,  an increase of 50 //g/m3 in SO2 or BS was associated
30      with a 3% (95% CI = 2.0, 4.0) increase in daily mortality; and the corresponding figure was 2%
31      (95% CI = 1.0, 3.0) for estimated PM10 (they used conversion: PM10 = TSP*0.55). In the

        April 2002                                8-34        DRAFT-DO NOT QUOTE OR CITE

-------
 1      central/eastern European cities, the increase in mortality associated with a 50 //g/m3 change was
 2      0.8% (CI = -0.1, 2.4) for SO2 and 0.6% (CI = 0.1, 1.1) per 50 //g/m3 change in BS.  Estimates of
 3      cumulative effects of prolonged (two to four days) exposure to air pollutants were comparable to
 4      those for one day effects. The effects of both pollutants (BS, SO2) were stronger during the
 5      summer and were mutually independent. Regarding the contrast between the western and
 6      central/eastern Europe results, the authors speculated that this could be due to: difference in
 7      exposure representativeness; difference in pollution toxicity or mix; difference in proportion of
 8      sensitive sub-population; and model fit for seasonal control. Bobak and Roberts (1997)
 9      commented that the heterogeneity between central/eastern and western Europe could be due to
10      the difference in mean temperature. However, Katsouyanni and Touloumi (1998) noted that,
11      having examined the source of heterogeneity, other factors could apparently explain the
12      difference  in estimates as well as or better than temperature.
13
14      APHEA1  Ambient Oxidants (Ozone and Nitrogen Dioxide) Results for Six Cities
15           Touloumi et al. (1997) reported on additional APHEA data analyses, which evaluated
16      (a) short-term effects of ambient oxidants on daily deaths from all causes (excluding accidents),
17      and (b) impacts on effect estimates for NO2 and O3 of including a PM measure (BS) in
18      multi-pollutant models.  Six cities in central and western Europe provided data on daily deaths
19      and NO2 and/or O3 levels. Poisson autoregressive models allowing for overdispersion were
20      fitted. Significant positive associations were found between daily deaths and both NO2 and O3.
21      Increases of 50 //g/m3 in NO2 (1-hour maximum) or O3 (1-hour maximum) were associated with
22      a 1.3% (95% CI 0.9-1.8) and 2.9% (95% CI 1.0-4.9) increase in the daily mortality, respectively.
23      There was a tendency for larger effects of NO2 in cities with higher levels of BS: when BS was
24      included in the model, the pooled estimate for the O3 effect was only slightly reduced, but the
25      coefficient for NO2 was reduced by half (but remained significant). The authors  speculated that
26      the short-term effects of NO2 on mortality might be confounded by other vehicle-derived
27      pollutants  (e.g., airborne ambient PM indexed by BS measurements). Thus, while this study
28      reports only relative risk levels for NO2 and O3 (but not for BS), it illustrates the importance of
29      confounding of NO2 and PM effects and the relative limited confounding of O3 and PM effects.
30
31

        April 2002                               8-35        DRAFT-DO NOT QUOTE OR CITE

-------
 1      APHEA1: A Sensitivity Analysis for Controlling Long-Term Trends and Seasonally
 2           In order to investigate further the source of the regional heterogeneity of PM effects and to
 3      examine the sensitivity of the RRs, the APHEA1 data were reanalyzed by APHEA investigators
 4      (Samoli et al., 2001).  Unlike previous analysis (i.e., analysis by Katsouyanni et al., 1997) in
 5      which sinusoidal terms for seasonal control and polynomial terms for weather were used, the
 6      investigators this time used a GAM model with smoothing terms for seasonal trend and weather,
 7      which is a more commonly used approach in recent years.  Using this model, the estimated
 8      relative risks for central-eastern cities were larger than those obtained in the previous analysis,
 9      reducing the contrast of estimated PM effects between central-eastern and western European
10      cities. Also, restricting the analysis to days with concentration < 150 ug/m3 further reduced the
11      differences between the western and central-eastern European cities. The authors conclude that
12      part of the heterogeneity in the estimated air pollution effects between western and central -
13      eastern cities in previous publications was caused by the statistical approach and the data range.
14      These results indicate that the apparent regional heterogeneity could be somewhat sensitive to
15      model specification. Since the number of cities used in the APHEA1 study is relatively small
16      (eight western and five central-eastern cities), the apparent regional heterogeneity found in the
17      earlier publications could also be due to chance.  Thus, such heterogeneity may be sensitive to
18      model specification and/or choice of data range.  The combined estimate for 50 //g/m3 increase in
19      PM10 was reported to be 3.3% (95CI: 2.6, 4.1)
20
21      APHEA2: Confounding and Effect Modification Using Extended Data
22           The APHEA2 analyses (Katsouyanni et al. 2001) included more cities  (29 cities) and a
23      more recent study period (variable years in 1990-1997, as compared to 1975-1992 in APHEA1).
24      As with the recent reanalysis of APHEA1 by Samoli et al. (2001), APHEA2 analyses used a
25      GAM Poisson model with a smoother to control for season and trends.  The  analyses put
26      emphasis on effect modification by city-specific factors. Thus, the city-specific coefficients from
27      the first stage of Poisson  regressions were modeled in the second stage regression using city-
28      specific characteristics as explanatory variables.  Inverse-variance weighted pooled estimates
29      (fixed-effects model) were obtained  as part of this model. When substantial heterogeneity was
30      observed, the pooled estimates were obtained using random-effects models.  These city-specific
31      variables included:  (1) air pollution  level and mix, such as average air pollution levels and

        April 2002                                8-36       DRAFT-DO NOT QUOTE OR CITE

-------
 1      PM/NO2 ratio (as an indicator of traffic-generated PM); (2) climatic variables, such as mean
 2      temperature and relative humidity; (3) health status of the population, such as the age-adjusted
 3      mortality rates, the percentage of persons over 65 years of age, and smoking prevalence;
 4      (4) geographic area (three regions: central-eastern, southern, and north-western). The study also
 5      addressed the issue of confounding by simultaneous inclusion of gaseous co-pollutants in city-
 6      specific regressions, and obtaining the pooled PM estimates for each co-pollutant included.
 7           Unlike APHEA1, in which the region (larger PM estimates in western Europe than in
 8      central-eastern Europe) was highlighted as the important factor, APHEA2 found several effect
 9      modifiers.  NO2 (i.e., index of high pollution from traffic) was an important one. The cities with
10      higher NO2 levels showed larger PM effects.  That is, the estimated PM10 risk was approximately
11      4-fold in cities with NO2 levels in the 75th percentile ("high"), as compared to cities with NO2
12      levels in the 25th percentile ("low") of the distribution. The cities with warmer climate showed
13      larger PM effects. The investigators noted that this might be  due to the better estimation of
14      population exposures with outdoor community monitors (because of more open windows). Also,
15      the cities with low standardized mortality rate showed larger PM effects.  The investigators
16      speculated that this may be because a smaller proportion of susceptible people (to air pollution)
17      are available in a population with a large age-standardized mortality rate. Interestingly, in the
18      pooled PM risk estimates from  models with gaseous pollutants, it was also NO2 that affected
19      (reduced) PM risk estimates most. For example, in the fixed-effects models, approximately 50%
20      reductions in both PM10 and BS coefficients were  observed when NO2 was included in the model.
21      SO2 only minimally reduced PM coefficients, whereas O3 actually increased PM coefficients.
22      Thus, in this analysis, NO2 was implicated both as a confounder and an effect modifier. The
23      overall combined estimate for total mortality for PM10 or BS was 3.0% (95CI: 2.0, 4.1).
24
25      8.2.2.3.4 An Examination of Effect Modification Using Past Results
26           Levy et al. (2000) sought  to explain the apparent heterogeneity of PM effects found in past
27      studies. Their analysis is different from the other multi-city studies discussed above in that they
28      analyzed the PM coefficients from past studies,  rather than obtaining city-specific coefficients in
29      a consistent time-series model specification. However, their results are worth mentioning here,
30      as they examined various city-specific covariates that are similar to those examined in other
31      multi-city studies, as well  those that are unique to  their study.  They applied an empirical Bayes

        April 2002                                 8-37        DRAFT-DO NOT QUOTE OR CITE

-------
 1      meta-analysis to 29 PM estimates from 21 published studies; and, in a second stage regression,
 2      they considered city-specific variables such as mortality rate, gaseous pollutants' regression,
 3      coefficients (that is, regressing a gaseous pollutant on PM), PM10 levels, PM2 5/PM10 ratio, central
 4      air conditioning prevalence, and heating/cooling degree days. Among these variables,
 5      PM2 5/PM10 ratio was a significant predictor (larger PM estimates for higher PM2 5/PM10 ratio)
 6      in the 19 U.S. cities data subsets. While sulfate data were not available for all the 19 studies, the
 7      investigators noted that the sulfate/PM10 ratio was highly correlated with both the mortality
 8      (r = 0.84) and with the PM2 5/PM10 ratio in the limited subset of data, indicating that the
 9      sulfate/PM10 ratio could be an even better predictor of regional heterogeneity of PM risk
10      estimates.  It would be interesting to estimate PM25/PM10 ratios or sulfate/PM10 ratios for a larger
11      U.S. dataset (e.g., Samet et al.'s 90  cities study) and examine if Levy et al.'s finding holds for
12      larger geographic coverage. After adjusting for city-specific covariates, Levy et al.'s combined
13      total mortality excess death estimate for the 19 U.S. PM studies was 3.5% (95% CI: 2.7, 4.4) per
14      50 //g/m3 increase in PM10.
15
16      8.2.2.3.5 Comparison of Effects Estimates from Multi-City Studies
17           In summary, based on pooled analyses of data combined across multiple cities, the percent
18      excess (total, non-accidental) deaths estimated per 50 //g/m3 increase in PM10 in the above
19      multi-city studies  were: (1) 2.3% in the 90 largest U.S. cities (4.5% in the Northeast region);
20      (2) 3.4% in 10 U.S. cities; (3) 3.5% in the 8 largest Canadian cities; and (4)  2.0% in western
21      European cities (using PM10 = TSP*0.55) in the original APHEA1; (5) 3.3% in the  reanalysis of
22      APHEA1; (6) 3.0% in APHEA2; and (7) 3.5% in Levy et al.'s analysis of the 19 U.S. studies.
23      These combined estimates are all consistent with the range of PM10 estimates previously reported
24      in the 1996PM AQCD.
25
26      8.2.2.4  The Role of Particulate Matter Components
27           Delineation  of the roles of specific ambient PM components in contributing to associations
28      between short-term PM exposures and mortality requires evaluation of several factors,  e.g., size,
29      chemical composition, surface characteristics, and presence of gaseous co-pollutants.  While
30      possible combinations of interactions among these factors can in theory be limitless, the actual
31      data tend to cover definable ranges  of aerosol characteristics and co-pollutant environments due

        April 2002                                 8-38        DRAFT-DO NOT QUOTE OR CITE

-------
 1      to typical source characteristics (e.g., fine particles tend to be combustion products in most
 2      cities). Newly available studies conducted in the last few years have begun to provide more
 3      extensive information on the issue of PM component roles; their results are discussed below in
 4      relation to three topics:  (1) PM particle size (e.g., PM25 vs. PM10_25); (2) chemical components;
 5      and (3) source oriented evaluations.
 6
 7      8.2.2.4.1  Particulate Matter Particle Size Evaluations
 8           Numerous studies published since the 1996 PM AQCD substantiate associations between
 9      PM25  and increased total mortality.  Consistent with the 1996 PM AQCD findings, effect size
10      estimates from the new  studies generally fall within the range of 2 to 6% excess total mortality
11      per 25 //g/m3 PM2 5, with many being statistically significant at p<0.05.
12           With regard to the relative importance of the fine and coarse fractions of inhalable PM10
13      particles capable of reaching thoracic regions of the respiratory tract, at the time of the 1996 PM
14      AQCD, there was only one acute mortality study (Schwartz et al., 1996a) that examined this
15      issue.  That study suggested that fine particles (PM25), distinctly more so than coarse fraction
16      (PM10_25) particles,  were associated with daily mortality.  A recent study (Klemm et al., 2000), to
17      reconstruct the data and to replicate the original analyses, essentially reproduced the original
18      investigators'  results.
19           Since the 1996 PM AQCD,  several new studies have used size-fractionated PM data to
20      investigate the relative importance of fine (PM25) vs. coarse (PM10_25) fraction particles.
21      Table 8-2 provides  synopses of those studies with regard to the relative importance of the two
22      size fractions, as well as some characteristics of the data. The average levels of PM25 ranged
23      from about 13 to 20 //g/m3 in the U.S. cities, but much higher average levels were measured in
24      Mexico City (27.4 //g/m3) and Santiago, Chile (64.0 //g/m3). As can be seen in Table 8-2, in the
25      northeastern U.S. cities  (Pittsburgh, Philadelphia, and Detroit) and Atlanta, GA, there was more
26      PM25  mass than PM10_25 mass on the average, whereas in the western U.S. (Phoenix, AZ;
27      Coachella Valley, CA; Santa Clara County, CA) the average PM10_2 5 levels were higher than
28      PM2 5  levels. It should be noted that the three Phoenix studies in Table 8-2 use much the same
29      data set, using fine  and coarse particle data from EPA's 1995-1997 platform study. Seasonal
30      differences in PM component levels should also be noted. For example, in Santa Clara County
31      and in Santiago, Chile, the winter PM2 5 levels averaged twice those  during summer.  The

        April 2002                                 8-39        DRAFT-DO NOT QUOTE OR CITE

-------
       TABLE 8-2.  SYNOPSIS OF SHORT-TERM MORTALITY STUDIES THAT
              EXAMINED RELATIVE IMPORTANCE OF PM2 5 AND PM10 2 5
Author, City
  Means (^g/m3); ratio
  ofPM25toPM10;and
  correlation between
  PM2 5 and PM10.2 5.
          Results regarding relative importance of
             PM2 5 vs. PM10_2 5 and comments.
Fairley (1999).
Santa Clara County,
CA
Ostro et al. (2000).
Coachella Valley,
CA
 Clyde et al. (2000).
 Phoenix, AZ
Mar etal. (2000).
Phoenix, AZ
1995-1997.
 Smith et al. (2000).
 Phoenix, AZ
Lippmann et al.
(2000). Detroit, MI
1992-1994.
Lipfert et al. (2000a).
Philadelphia, PA
1992-1995.
PM25 mean= 13;
PM25/PM10 = 0.38;
r=0.51.
PM2 5 (Palm Springs and
Indio, respectively)
mean= 12.7, 16.8;
PM25/PM10 = 0.43,0.35;
r= 0.46, 0.28.

PM25mean = 13.8;
PM25/PM10 = 0.30;
r=0.65.
PM25 (TEOM) mean= 13;
PM25/PM10 = 0.28;
r=0.42.
Not reported, but likely
same as Clyde's or Mar's
data from the same
location.
PM25 mean=18;
PM25/PM10 =0.58;
r=0.42.
PM25 mean=17.3;
PM2J/PM10 =0.72.
Of the various pollutants including PM10, PM25, PM10_25,
sulfates, nitrates, COH, CO, NO2, and O3, strongest
associations were found for ammonium nitrate and PM2 5.
PM2 5 was significantly associated with mortality, but PM10_2 5
was not, separately and together in the model. Sulfate was a
significant predictor of mortality in single pollutant model,
but not when PM2 5 was included simultaneously.  Winter
PM2 5 level is more than twice that in summer.

Total mortality was more significantly associated with PM2 5
than with PM10_2 5. Cardiovascular mortality was associated
with PM10_2.5 more significantly than  with PM2 5, but their
effect size estimates per IQR were similar.
Using Bayesian Model Averaging that incorporates model
selection uncertainty, with 29 covariates (lags 0- to 3-day),
effects of coarse particles (most consistent at lag 1 day) were
found to be stronger than that for fine particles. The
association was for mortality confined to the region where
fine particles (PM2 5) are expected to be uniform.

Total mortality was weakly (p < 0.10) associated with PM10_25.
It was less strongly (p > 0.10) associated with PM25.
Cardiovascular mortality was both significantly associated
with PM2 5 (lags 1, 3, 4) and PM10.2 5 (lag 0).

In linear PM effect model, a statistically significant mortality
association found with PM10_2 5, but not with PM2 5. In models
allowing for a threshold, evidence of a threshold for PM2 5
(in the range of 20-25 ,wg/m3) suggested, but not for PM10_2 5.
Seasonal interaction in the PM10_2 5 effect also reported: the
effect being highest in spring and summer when anthropogenic
concentration of PM10_25 is lowest.

Both PM2 5 and PM10_2 5 were positively associated with
mortality outcomes to a similar extent. Simultaneous inclusion
of PM2 5 and PM10_2 5 also resulted in comparable effect sizes.
Similar patterns were seen in hospital admission outcomes.

The authors conclude that no systematic differences were seen
according to particle size or chemistry. However, when PM2 5
and PM10_2.5 were compared, PM2 5 (at lag 1 or average of lag 0
and 1) was more significantly (with larger attributable risk
estimates) associated with cardiovascular mortality than
PM10.,5.
April 2002
                           8-40
                DRAFT-DO NOT QUOTE OR CITE

-------
   TABLE 8-2 (cont'd). SYNOPSIS OF SHORT-TERM MORTALITY STUDIES THAT
              EXAMINED RELATIVE IMPORTANCE OF PM2 5 AND PM10 2 5
Author, City
  Means Og/m3); ratio
  ofPM25toPM10;and
  correlation between
  PM2 5 and PM10.2 5.
          Results regarding relative importance of
             PM2 5 vs. PM10_2 5 and comments.
Klemm and Mason
(2000). Atlanta, GA
Klemm et al. (2000)
 Chock et al. (2000).
 Pittsburgh, PA
Burnett et al. (2000)
8 Canadian cities
1986-1996

Castillejos et al.
(2000). Mexico City.
1992-1995
 Cifuentes et al.
 (2000).
 Santiago, Chile
 1988-1996.
Anderson et al.
(2001). The west
Midlands
conurbation, UK.
1994-1996.
PM25mean = 19.9;
PM25/PM10 =0.65.
Mean PM2 s ranges from
11.3 in Portage to 29.6 in
Steubenville.  Mean
PM10_2 5 ranges from 6.6 in
Portage to 16.1 in
Steubenville.  Mean
PM2 5/PM10 ranges from
50.1%inTopekato66%
in Kinston-Harriman.

Data distribution not
reported.
PM25/PM10 *  0.67.
PM25 mean=13.3;
PM25/PM10=0.51;
r=0.37.

PM25mean=27.4;
PM25/PM10=0.61;
r=0.52.
PM25 mean=64.0;
PM25/PM10=0.58;
r=0.52.
PM25mean=14.5;
PM2J/PM10 =0.62;
r=0.92.
No significant associations were found for any of the
pollutants examined, possibly due to a relatively short study
period (1-year).  The coefficient and t-ratio were larger for
PM2 5 than for PM10.2 5.

Significant associations between total mortality and PM2 5 in 3
cities and in pooled effect. No significant association with
PM10_2.5 in the replications study for any city.
Seasonal dependence of correlation among pollutants, multi-
collinearity among pollutants, and instability of coefficients
were all emphasized in discussion and conclusion. These
considerations and small size of dataset stratified by age group
and season limit confidence in results finding no consistently
significant associations for any size fraction.

PM2 5 was a stronger predictor of mortality than PM10_2 5. For
chemical species, sulfate ion, nickel, and zinc from the fine
fraction were most strongly associated with mortality.

Both PM2 5 and PM10_2 5 were associated individually with
mortality, but the PM10_2 5 effect size was larger and more
significant. When both were included in the model, the effect
size of PM10_2 5 remained the same but that of PM2 5 was
virtually eliminated.

Results were different for warmer and colder months. PM2 5
was more important than PM10_2 5 in the whole year and in
winter, but not in summer. The mean of PM2 5 was more than
twice higher in winter (82.4 ^g/m3) than in summer (32.8),
whereas the mean of PM10_2 5 was more comparable for winter
(49.9 Mg/m3) and for summer (42.9).
No significant association seen between total mortality and
any of the PM indices in the all year analysis, but PM10 and
PM2 5 were significantly associated with total mortality in the
warm season (April-September).  PM10_2 5 was generally more
weakly associated with mortality outcomes than PM10 or PM2 5
April 2002
                           8-41
               DRAFT-DO NOT QUOTE OR CITE

-------
 1      temporal correlation between PM25 and PM10_25 ranged between 0.30 and 0.65.  Such differences
 2      in ambient PM mix characteristics from season to season or from location to location
 3      complicates assessment of the relative importance of PM25 and PM10_2 5.
 4           To facilitate a quantitative overview of the effect size estimates and their corresponding
 5      uncertainties from these studies, the percent excess risks are plotted in Figure 8-6. These
 6      excluded the Clyde et al. study, in which the model specification did not obtain RRs for PM2 5
 7      and PM10_2 5 separately, and the Smith et al.  study, which did not present linear term RRs for
 8      PM25 and PM10_2 5. Note that, in most of the original studies, the RRs were computed for
 9      comparable distributional features (e.g., inter quartile range, mean, 5th-to-95th percentile,  etc.).
10      However, the increments derived  and their absolute values varied across studies; and therefore,
11      the RRs used in deriving the excess risk estimates delineated in Figure 8-6 were re-computed for
12      consistent increments of 25 //g/m3 for both PM2 5 and PM10_25.  Note also that re-computing the
13      RRs per 25 //g/m3 in some cases changed the relative effect size between PM2 5 and PM10_25, but
14      it did not affect the relative significance.
15           All of the studies found positive associations between both the fine and coarse PM indices
16      and increased mortality risk, with most for PM2 5 and a few for PM10_25 being statistically
17      significant. However, most of the studies did not have large enough sample  sizes to separate out
18      what often appear to be relatively small differences in effect size estimates; but several do show
19      statistical distinctly larger and significant mortality associations with PM2 5 than for non-
20      significant PM10_2 5 effects. For example, the Klemm et al. (2000) recomputation of the Harvard
21      Six Cities time-series study reconfirmed the original Schwartz et al. (1996a) finding of PM2 5
22      being significantly associated with excess mortality, whereas PM10_2 5 was not.  Similar results
23      were obtained by the other multi-city study, i.e., the 8 largest Canadian cities study by Burnett
24      et al. (2000), and by the Atlanta (Klemm and Mason, 2000), Santa Clara (Fairley, 1999), and the
25      Coachella Valley (Ostro et al., 2000) studies.  There were two studies in which the importance of
26      PM2 5 and PM10_2 5 were considered to be similar or, at least, not distinguishable: Philadelphia, PA
27      (Lipfert et al., 2000a) and Detroit, MI (Lippmann et al., 2000).  The three Phoenix studies
28      obtained "mixed"  results, in that the Smith et al. (2000) and Clyde et al. (2000) analyses  (not
29      shown in Figure 8-6) found PM10_2 5 to appear to be more important in explaining mortality than
30      PM2 5, but Mar et al. (2000) found both to be significant (as depicted in Figure 8-6).  Also, the
31      Mexico City analysis by Castillejos et al. (2000) implicated PM10_25 as the apparent more

        April 2002                                8-42        DRAFT-DO NOT QUOTE OR CITE

-------
>
to
o
o
to
 • Q
~
Lag 2 day MA } ^ - } Winter
o
HH
H
W
Figure 8-6. Percent excess risks estimated per 25 Aig/m3 increase in PM2 5 or PM10_2 5 from new studies evaluating both PM2
                                                                                                       -2.5
                 and PM10_2 5 data for multiple years, based on single pollutant (PM only) models. All lags = 1 day, unless indicated

                 otherwise (See Section 8.4.2 for same studies shown here that found different risk estimates in MP models with

                 both fine and coarse particles included).

-------
 1      important fraction of PM10. However, the Santiago, Chile study (Cifuentes et al., 2000) found
 2      significant associations with both fine and coarse fractions and interesting seasonal differences,
 3      as well.  In Chock et al.'s (2000) analysis of Pittsburgh, PA data, the authors emphasized the lack
 4      of significant PM associations; and no specific comments were made regarding the relative
 5      importance of PM25 versus PM10_25.
 6           The Canadian 8-city study (Burnett et al., 2000) is noteworthy for a variety of reasons,
 7      including the use of elemental composition and principal components analyses to provide
 8      additional information about the relative importance of fine and coarse particles. The PM25
 9      effect on mortality is greater than the PM10_2 5 effect for all gaseous-pollutant models in Table 5 of
10      Burnett et al. (2000) and in the principal component model 1 in their Table 8, where both PM
11      size fractions and the four gaseous co-pollutants are used simultaneously. PM component
12      models from this study are discussed further below, in Section 8.2.2.4.2.
13           The Lippmann et al. (2000) results for Detroit are also noteworthy in that additional PM
14      indices were evaluated besides those depicted in Figure 8-6 and the overall results obtained may
15      be helpful in comparing fine- versus coarse-mode PM effects. In analyses of 1985 to 1990 data,
16      PM-mortality relative risks and their statistical significance were generally in descending order:
17      PM10, TSP-SO4=, and TSP-PM10. For the 1992-1994 period, relative risks for equivalent
18      distributional increment (e.g., IQR) were comparable  among PM10, PM25, and PM10_25 for both
19      mortality and hospital admissions categories; and SO4= was more strongly associated with most
20      outcomes than FT.  Consideration of the overall pattern of results led the authors to state that the
21      mass of the smaller size index could explain a substantial portion of the variation in the larger
22      size indices.  In these data, on average, PM25 accounted for 60% of PM10 (up to 80% on some
23      days) and PM10 for 66% of TSP mass. Also, the temporal correlation between TSP and PM2 5
24      was r = 0.63, and for PM25 vs. PM10 r = 0.90, suggesting that much of the apparent larger particle
25      effects may well be mainly driven by temporally covarying smaller PM2 5 particles. The stronger
26      associations for sulfates than FT, suggestive of non-acid fine particle effects, must be caveated by
27      noting the very low FT levels present (often circa non-detection limit).
28           Three research groups have examined the same Phoenix, AZ data set, using different
29      methods. While these groups used somewhat different approaches, there is some consistency
30      among their results in that PM10_2 5 appeared to emerge as possibly the more important predictor
31      of mortality versus PM2 5.  In the Clyde et al. (2000) analysis, PM-mortality associations were

        April 2002                                 8-44        DRAFT-DO NOT QUOTE OR CITE

-------
 1      found only for the geographic area where PM2 5 was considered uniformly distributed, but the
 2      association was with PM10_2 5, not PM2 5. Based on the Bayes Information Criterion, the highly
 3      ranked models consistently included 1-day lagged PM10_25. Smith et al.'s (2000) analyses found
 4      that, based on a linear PM effect, PM10_2 5 was significantly associated with total mortality, but
 5      PM25 was not. In the Mar et al. (2000) analysis, total mortality was significantly associated with
 6      CO and NO2 and weakly (p < 0.1) associated with PM10 and PM10_2 5 (and PM2 5 was more weakly
 7      associated); however, cardiovascular mortality (CVM) was significantly associated with both
 8      PM25 and PM10_2 5 at p<0.05.  CVM was also significantly associated with a motor vehicle source
 9      category with loading of PM2 5, EC, OC, CO, NO2, and some trace metals, as shown by factor
10      analyses discussed below. The PM2 5 in Phoenix is mostly generated from motor vehicles,
11      whereas PM10_25 consists mainly of two types of particles: (a) crustal particles from natural (wind
12      blown dust) and anthropogenic (construction and road dust) processes, and (b) organic particles
13      from natural biogenic processes (endotoxin and molds) and anthropogenic (sewage aeration)
14      processes.
15           The Castillejos et al. (2000) and Cifuentes et al. (2000) analyses also appear to implicate
16      PM10_2 5, as well as PM25, as importantly contributing to mortality in two non-U.S. locations,
17      Mexico City and Santiago, Chile. The  latter study also suggests possible seasonal differences in
18      Santiago, the PM effects in summer being more than double those in winter at that South
19      American location.
20
21      Crustal Particle Effects
22           Since the 1996 PM AQCD, several  studies have yielded interesting new information
23      concerning possible roles of crustal wind-blown particles or crustal particles within the fine
24      particle fraction (i.e., PM25) in contributing to observed PM-mortality effects.
25           Schwartz et al. (1999), for example, investigated the association of coarse particle
26      concentrations with non-accidental deaths in Spokane, Washington, where dust storms elevate
27      coarse particle concentrations. During  the 1990-1997 period, 17 dust storm days were identified.
28      The PM10 levels during those storms averaged 263 //g/m3, compared to 39 //g/m3 for the entire
29      period. The coarse particle domination of PM10 data on those dust storm days was confirmed by
30      a separate measurement of PM10 and PMX during a dust storm in August,  1996: the PM10 level
31      was 187 //g/m3, while PMX was only 9.5 //g/m3.  The deaths on the day of a dust storm were

        April 2002                                 8-45        DRAFT-DO NOT QUOTE OR CITE

-------
 1      contrasted with deaths on control days (n=95 days in the main analysis and 171 days in the
 2      sensitivity analysis), which are defined as the same day of the year in other years when dust
 3      storms did not occur. The relative risk for dust storm exposure was estimated using Poisson
 4      regressions, adjusting for temperature, dewpoint, and day of the week. Various sensitivity
 5      analyses considering different seasonal adjustment, year effects, and lags, were conducted. The
 6      expected relative risk for these storm days with an increment of 221 //g/m3 would be about 1.04,
 7      based on PM10 relative risk from past studies, but the estimated RR for high PM10 days was found
 8      to be only 1.00 (95% CIO.95-1.05) per 50 //g/m3 PM10 change in this study. Schwartz et al.
 9      concluded that there was no evidence to suggest that coarse (presumably crustal) particles were
10      associated with daily mortality.
11           Pope et al. (1999a) investigated PM10-mortality associations in three metropolitan areas
12      (Ogden, Salt Lake City, and Provo/Orem) in Utah's Wasatch Front mountain region during the
13      1985-1995 period.  Although the three metropolitan areas shared common weather patterns,
14      pollution levels and patterns among the three areas were different due to different emission
15      sources. The authors utilized an index of air stagnation (the clearing index which the National
16      Weather Service computes from temperature, moisture and wind) to identify and screen obvious
17      windblown dust days, days clearly identified as with low stagnation index but high PM10.  They
18      found that Salt Lake City experienced substantially more episodes of wind-blown dust.  They
19      therefore conducted Poisson regression of mortality series using both unscreened and screened
20      PM10 data. The effects of screening were most apparent in Salt Lake City results.  Before
21      screening, no significant relationships were observed; after screening, the RRs per 50 //g/m3
22      increase in PM10 for mortality in the three metropolitan areas were 1.12 (95%  CI:  1.045 - 1.20),
23      1.023  (1.00 - 1.047), and 1.019 (0.979 - 1.06) for Ogden, Salt Lake City, and Provo/Orem,
24      respectively. These results suggest that the pollution episodes of wind-blown  (crustal-derived)
25      dusts were less associated with mortality than were the episodes of (presumably) combustion-
26      related particles.
27           Ostro et al. (1999a) analyzed the Coachella Valley, CA data for 1989-1992.  This desert
28      valley, where coarse particles of geologic origin comprise circa 50-60% of annual-average PM10
29      (> 90% during wind episodes throughout the year), includes the cities of Palm Springs and Indio,
30      CA. Total, respiratory, cardiovascular, non-cardiorespiratory and age-over-50 deaths were
31      analyzed.  The correlation between gravimetric and beta-attenuation measurements, separated by

        April 2002                                8-46        DRAFT-DO NOT QUOTE OR CITE

-------
 1      25 miles, was high (r = 0.93); and the beta-attenuation data were used for analysis.  GAM
 2      Poisson models adjusting for temperature, humidity, day-of-week, season, and time were used.
 3      Seasonally stratified analyses were also conducted.  Lags 0 through 3 days (separately) of PM10,
 4      along with moving averages of 3 and 5 days, were evaluated, as were O3, NO2, and CO.
 5      Associations were found between 2- or 3-day lagged PM10 and all mortality categories examined,
 6      except non-cardiorespiratory. Effect size estimates  for total and cardiovascular deaths were
 7      larger for warm season (May through October) than for all year, analogous to the Cifuentes et al.
 8      (2000) findings for Santiago, Chile.  NO2 and CO were statistically significant predictors of
 9      mortality in single pollutant models; but in multi-pollutant models, all gaseous pollutants
10      coefficients were reduced and non-significant, whereas PM10 coefficients remained the same and
11      significant. Ostro et al. (2000) also conducted a follow-up study of the Coachella Valley data for
12      1989-1998, using actual PM2 5 and PM10_25 data for  the last 2.5 years but PM2 5 and PM10_25
13      concentrations estimated for the other, earlier years. PM2 5, CO, and NO2 were significantly
14      associated with all-cause mortality; and PM10 and PM10_2 5 with cardiovascular mortality, but not
15      PM2 5 (possibly due to the low range of concentrations and reduced sample size for PM2 5 data
16      versus PM10 data). Thus, although the cardiovascular mortality results hint at crustal particle
17      effects possibly being important in this desert situation, the ability to discern more clearly the
18      role of fine particles would likely be improved by analyses of more years of actual data for PM2 5.
19           Laden et al.  (2000) analyzed Harvard Six Cities study data and Mar et al. (2000) the
20      Phoenix data to investigate the role of crustal particles in PM2 5 samples on daily mortality.
21      These studies are discussed in more detail below in Section 8.2.2.4.3 on the source-oriented
22      evaluation of PM; and only the basic results regarding crustal particles are mentioned here. The
23      elemental abundance data (from X-ray fluorescence spectroscopy analysis of daily filters) were
24      analyzed to estimate the concentration of crustal particles in PM2 5 using factor analysis. Then
25      the association of mortality with fine crustal mass was estimated using Poisson regression
26      (regressing mortality on factor scores for "crustal factor"), adjusting for time trends and weather.
27      No positive association was found between fine  crustal mass factor and mortality.
28           The above results, overall, mostly suggest that crustal particles (coarse or fine) per se are
29      not likely associated with daily mortality.  However, as noted in the previous section, three
30      analyses of Phoenix, AZ data suggested that PM10_25 may be associated with mortality. The
31      results from one of the three studies (Smith et al., 2000) suggest that coarse particle mortality

        April 2002                                 8-47         DRAFT-DO NOT QUOTE OR CITE

-------
 1      associations are stronger in spring and summer, when the anthropogenic portion of PM10_2 5 is
 2      lowest as determined by factor analysis. However, during spring and summer, biogenic
 3      processes (e.g., wind-blown endotoxins and molds) may contribute more to the PM10_2 5 fraction
 4      in the Phoenix area, clouding any attribution of observed PM10_2 5 effects there to crustal particles,
 5      per se. Disentangling potential contributions of biogenically-derived organic particle
 6      components from those of crustal materials in the PM10_25 fraction in Mexico City and Santiago
 7      poses further interesting challenges.
 8
 9      Ultrafine Particle Effects
10           The Wichmann et al. (2000) study evaluated the attribution of PM effects to specific size
11      fractions, including both the number concentration (NC) and mass concentration (MC) of
12      particles in a given size range.  The study was carried out in the small German city of Erfurt
13      (pop. 200,000) in the former German Democratic Republic, by a team of scientists at the
14      Gessellschaft fur Strahlenforschung (GSF) and Ludwig Maximilian University in Germany.
15      Erfurt was heavily polluted by particles and SO2 in the 1980s, and excess mortality was attributed
16      to high levels of TSP by Spix et al. (1993).  Concentrations of PM and SO2 have markedly
17      dropped since then.  The present study provides a much more detailed look at the health effects
18      of ultrafine particles (diameter < 0.1 //m) than earlier studies, and allows examination of effects
19      related to number counts for fine and ultrafine particles, as well as to their mass.
20           The Mobile Aerosol Spectrometer (MAS), developed by GSF, produces number and mass
21      concentrations in three size classes of ultrafmes (0.01 to 0.1 //m) and three size classes of larger
22      fine particles (0.1 //m to 2.5 //m). The  mass concentration MCO.01-2.5 is well  correlated with
23      gravimetric PM2 5, and the number concentration NCO.01-2.5 is well correlated with total particle
24      counts from a condensation  particle counter (CPC). Mortality data were coded by cause of death,
25      with some discrimination between underlying causes and prevalent conditions of the deceased.
26      Some analyses looked at cardiovascular causes without respiratory, respiratory  without
27      cardiovascular, and both causes together as separate groups. Age was used as a modifying factor,
28      as was weekly data for all  of Germany on influenza and similar diseases. Daily mortality data
29      were fitted using a Poisson Generalized Additive Model (GAM), with adjustments for weather
30      variables, time trends, day of week,  and particle indices. Two types of models were fitted, one


        April 2002                                 8-48        DRAFT-DO NOT QUOTE OR CITE

-------
 1      using the best single-day lag for air pollution and a second using the best polynomial distributed
 2      lag (PDL) model for air pollution.
 3           Winter PM generally had the most significant positive effects on mortality, and fall PM
 4      effects were similar in magnitude, but less significant because of the smaller NC and MC in fall
 5      than in winter. Summer PM effects were consistently lower and not significant. PDL models
 6      generally had larger and more significant PM effects than single-day lag models.  Log-
 7      transformed pollution models occasionally provided better fits than untransformed pollutant
 8      models, particularly for number concentration indices in single-day lag models. However, there
 9      were some nonlinear relationships that could not be adequately described by either parametric
10      model, as shown by use of LOESS models. The results cited  in Table 8-1 and Appendix
11      Table 8A-1 are all for linear PDL models, to facilitate comparison.
12           Mass concentration was most often significantly associated with excess mortality in
13      one-pollutant models, with excess risks for MAS MCO.1-2.5 being about 6.2% (CI1.4, 11.2) per
14      25 //g/m3.  The non-significant estimate from filter PM25 was about 3% (CI -1.7, 7.9) per
15      25 //g/m3.  Filter PM10 estimates were also significant predictors of mortality overall, about 6.6%
16      excess risk per 50 //g/m3 (CI 0.7 to 12.8) in PDL models.
17           Mass concentrations for smaller fine particles were also often significant, with excess risk
18      for MCO.01-1.0 being ca. 5.1% (CI 0.2, 10.2) per 25 //g/m3 in a linear PDL model.  Smaller-size
19      components of MCO.01-1.0 were also significantly associated, or nearly so,  with excess
20      mortality.  The intermodal fraction MCI.0-2.5 was also significant in a PDL logarithmic model,
21      4.7% (CI 1.05, 8.5) per IQR in log concentration.  No results were reported for the effects of
22      ultrafme mass concentrations in classes 0.01-0.03, 0.03-0.05,  or 0.05-0.1 //g/m3.
23           Number concentrations of ultrafme particles were also associated with excess mortality,
24      significantly or nearly so in smaller size classes. The results for linear models are shown in
25      Table 8-3.  The table also shows how much the estimated excess risks are reduced,  sometimes
26      drastically, when co-pollutants (especially SO2 and NO2) are included in a two-pollutant model.
27      Number and mass concentrations of various ultrafme and fine particles in all size ranges are
28      rather well correlated with gaseous co-pollutants except for the intermodal size range MCI.0-2.5.
29      The correlations range from 0.44 to 0.62 with SO2, from 0.58 to 0.66 with NO2, and from 0.53 to
30      0.70 with CO. The mass correlations range from 0.53 to 0.62 with SO2, from 0.48 to 0.60 with
31      NO2, and from 0.56 to 0.62 with CO.  The large decreases in excess risk for number

        April 2002                                 8-49        DRAFT-DO NOT QUOTE OR CITE

-------
               TABLE 8-3. EXCESS TOTAL MORTALITY RISKS ESTIMATED TO BE
          ASSOCIATED WITH VARIOUS AMBIENT PARTICLE SIZE-RELATED INDICES
PM Index
NCO.01-0.03
NCO.03-0.05
NCO.05-0.1
NCO.01-2.5
NCO.01-0.1




MCO.01-2.5

Co-Pollutant
None
None
None
None
None
SO2
NO2
CO
MCO.01-2.5
None
SO2

Excess Risk, %
3.00a
3.80a
4.00a
6.891b
8.238b
4.758b
0.739b
3.594b
4.123b
6.194C
2.014C
Single-Pollutant Models
Lower 95% CL
-0.342
0.021
-0.307
0.662
0.252
-0.451
-3.951
-2.312
-1.437
1.409
-2.304

Upper 95% CL
6.455
7.722
8.493
13.504
16.86
10.239
5.658
9.856
9.996
11.205
6.523
         "Risks estimates for mortality associated with number concentrations (NC) in specified ranges. At actual
         interquartile range, respectively 8888, 2524, and 1525 particles/cm3.
         bAt standard increment 25,000 particles/cm3; winter IQR is 22,211 particles/cm3, annual IQR is 12,690 particles/cm3.
         cAt standard increment 25 ^g/m3.
         Source: Based on Wichman et al. (2000), as calculated by U.S. EPA.
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
concentration, particularly when NO2 is a co-pollutant with NCO.01-0.1. clearly involves a more
complex structure than simple correlation. The large decrease in excess risk when SO2 is a
co-pollutant with MCO.01-2.5 is not readily explained, and it is discussed in some detail in
Wichmann et al. (2000).
     SO2 is a strong predictor of excess mortality in this study; and its estimated effect is little
changed when different particle indicators are included in a two-pollutant model.  The authors
noted:  ". . .the [LOESS] smoothed dose response curve showed most of the association at the
left end, below 15 //g/m3, a level at which effects were considered biologically implausible. . ."
Replacement of sulfur-rich surface coal has reduced mean SO2 levels in Erfurt from 456 //g/m3
in 1988 to 16.8 //g/m3 during 1995 to 1998 and to 6 //g/m3 in 1998.  The estimated concentration-
response functions for SO2 are very different in these time periods, comparing Spix et al. (1993)
        April 2002
                                           8-50
DRAFT-DO NOT QUOTE OR CITE

-------
 1      with Wichmann et al. (2000) results.  Wichmann et al. concluded "These inconsistent results for
 2      SO2 strongly suggested that SO2 was not the causal agent but an indicator for something else."
 3      The authors offered no specific suggestions as to what the "something else" might be, but they
 4      did finally conclude that their studies from Germany strongly supported particulate air pollution
 5      as more relevant than SO2 to observed mortality impacts.
 6           The authors also found that ultrafme particles, NO2 and CO form a group of pollutants
 7      strongly identified with motor vehicle traffic. Immediate and delayed effects seemed to be
 8      independent in two-pollutant models, with single-day lags of 0 to 1 days and 4 to 5 days giving
 9      'best fits' to data.  The delayed effect of ultrafme particles was stronger than that for NO2 or CO.
10           Another finding of interest is that the excess risk in Erfurt is larger and more significantly
11      associated with ages < 70 years than with  older ages.  This is consistent for PDL models for
12      NCO. 01-0.1. MCO.01-2.5. andPM10.  None of the single lag day models were significant.
13           Examination of prevalent disease categories found larger and more significant risks for
14      respiratory disease mortality than for cardiovascular mortality in almost all models. Combined
15      cardiovascular or combined respiratory diseases were generally the next highest category. Other
16      natural causes (i.e., neither respiratory nor cardiovascular) almost always had the lowest risk.
17
18      8.2.2.4.2  Chemical Components
19           Ten new studies from the U.S. and Canada examined specific chemical components of PM.
20      Table 8-4 shows the  chemical components examined in these studies, the mean concentrations
21      for Coefficient of Haze (COH), sulfate,  and H+, as well as the list of those that were found to be
22      associated with increased mortality.  There are several chemical components of PM whose
23      associations with mortality can be compared across studies, including COH,  sulfate, and H+.
24
25      Coefficient of Haze, Elemental Carbon, and Organic Carbon
26           COH is highly  correlated with elemental carbon (EC) and is often considered as a good PM
27      index for motor  vehicle sources (especially diesel), although other combustion processes such as
28      space heating likely also contribute to COH levels. Several studies (Table 8-4) examined COH;
29      and, in most cases, positive and significant associations with mortality outcomes were reported.
30      In terms of relative significance of COH in comparison to other PM components, COH was not
31      the clearly most significant PM component in any of these  studies. The average level of COH in

        April 2002                                8-51        DRAFT-DO NOT QUOTE OR CITE

-------
 TABLE 8-4. SUMMARY OF PARTICULATE MATTER CHEMICAL COMPONENTS
                      ANALYZED IN RECENT STUDIES
Mean COH Mean SCV
Author, City (1000ft) (Mg/m3)
Burnett et al. 0.42 9.2
(1998b)
Toronto, Canada.
1980-1994.
Burnett etal. 0.26 2.6
(2000).
8 largest Canadian
cities
1986-1996.
Fairley (1999). 0.5 1.8
Santa Clara County,
CA
Other PM
Mean H+ components
(nmol/m3) analyzed
TSP, estimated
PM10 and PM2 5
PM10, PM2 5,
PM10_2 5, and
47 trace elements
PM10, PM2 5,
PM10_2 5, and
nitrate
PM components
associated with mortality.
Comments.
TSP, COH, sulfate,
estimated PM10 and PM2 5.
However, CO together
with TSP explained most
of the association.
PM10, PM2 5, COH,
sulfate, Zn, Ni, and Fe
significantly associated
with total mortality.
COH, sulfate, nitrate,
PM10, and PM2 5 were
associated with mortality.
                                                        PM2 5 and nitrate most
                                                        significant.
Gwynn et al. 0.2
(2000).
Buffalo, NY
1988-1990
Lipfert et al. 0.28
(2000a).
Philadelphia, PA
1992-1995
Lippmann et al.
(2000). Detroit, MI
1992-1994
Klemm and Mason
(2000). Atlanta, GA
1998-1999

5.9 36.4 PM10
5.1 8.0 Nephelometry,
NH4+, TSP, PM10
PM2 5, and PM10.2 5
5.2 8.8 PM10PM25,and
PM10.2.5
5.2 0 Nitrate, EC, OC,
oxygenated HC,
PM10, PM25, and
PM10.2.5

Sulfate, H+, PM10, and
COH were associated with
total mortality. COH was
least significant predictor.
Essentially all PM
components were
associated with mortality.
PM10, PM2 5, and PM10.2 5
were more strongly
associated with mortality
outcomes than sulfate or
"Interim" results based on
one year of data.
No statistically significant
associations for any
pollutants. Those with
t-ratio of at least 1.0 were:
H+, PM10, and PM2 5,
April 2002
8-52
DRAFT-DO NOT QUOTE OR CITE

-------
             TABLE 8-4 (cont'd).  SUMMARY OF PARTICIPATE MATTER CHEMICAL
                           COMPONENTS ANALYZED IN RECENT STUDIES
Author, City
Mean COH Mean SO4~
(1000ft) (Mg/m3)
MeanH+
(nmol/m3)
Other PM
components
analyzed
PM components
associated with mortality.
Comments.
        Mar et al. (2000).
        Phoenix, AZ
        1995-1997
        Tsai et al. (2000).
        Newark, Elizabeth,
        and Camden, NJ
        1981-1983
             12.7
        Hoek et al. (2000).
        The Netherlands
        1986-1994

        Goldberg et al.
        (2000). Montreal,
        Quebec, Canada.
        1984-1993.
        Anderson et al.
        (2001). The west
        Midlands
        conurbation, UK.
        1994-1996.
              3.8
           (median)
0.24
              3.7
                                   S, Zn, Pb,
                                   soil-corrected K,
                                   reconstructed soil,
                                   EC, OC, TC,
                                   PM10, PM2 5, and
                                   PM,,
PM15, PM2 5,
sulfates
cyclohexane-
solubles (CX),
dichloromethane-
solubles (DCM),
and acetone-
solubles (ACE).

PM10, BS, and
nitrate
Predicted PM2 5,
and extinction
coefficient (visual-
range derived).
PM10, PM2 5,
PM10.25, andBS.
S, Pb, and soil were
negatively associated with
total mortality. PM10 and
PM10_2 5 were positively
associated with total
mortality. Soil-corrected
K, non-soil PM2 5, EC,
OC, TC, PM10, PM2 5, and
PM10_2 5 were associated
with cardiovascular
mortality.

PM15, PM2 5, sulfate, CX
and ACE were
significantly associated
with total and/or
cardiovascular mortality
in Newark and/or
Camden.
Sulfate, nitrate, and BS
were more consistently
associated with total
mortality than PM10.
COH, predicted PM2 5,
and sulfate were
associated with various
mortality outcomes
(mostly elderly and
stronger associations in
summer).
Significant associations
between all-cause
mortality with PM indices
(except PM10_2 5) were
seen only in warm season.
1      these studies ranged from 0.2 (Buffalo, NY) to 0.5 (Santa Clara County, CA) 1000 linear feet.

2      The correlations between COH and NO2 or CO in these studies (8 largest Canadian cities; Santa

3      Clara County, CA; and Buffalo, NY) were moderately high (r ~ 0.7 to 0.8), suggesting a likely
       April 2002
                       8-53
  DRAFT-DO NOT QUOTE OR CITE

-------
 1      motor vehicle contribution. Some of the inconsistencies in the results across cities may be in
 2      part, due to the differences in COH levels.  For example, in Buffalo, NY (where COH was
 3      lowest), no significant association was found for any pollutant, possibly due to small sample size
 4      (« 1 year of data). However, both EC and OC were significant predictors of cardiovascular
 5      mortality in the Phoenix study, with their effect sizes per IQR being comparable to those for
 6      PM10,  PM2 5, and PM10_2 5; there, EC and OC represented major mass fractions of PM25 (11% and
 7      38%, respectively) and correlated highly with PM25 (r = 0.84 and 0.89, respectively).  They were
 8      also highly correlated with CO and NO2 (r ~ 0.8 to 0.9), indicating their associations with an
 9      "automobile" factor.  Thus, the COH and EC/OC results from the Mar et al. (2000) study suggest
10      that PM components from motor vehicle sources are likely associated with mortality.
11          In a recent study in Montreal, Quebec, by Goldberg et al. (2000), COH appeared to be
12      correlated with some of the mortality outcomes more strongly than other PM indices such as the
13      visual-range derived extinction coefficient (considered to be a good indicator of sulfate).
14      However, the main focus of the study was the role of cardio-respiratory risk factors for air
15      pollution, and the investigators warned against comparing the relative strength of associations
16      among PM indices, pointing out complications such as likely error involved in the visual range
17      measurements. Also, the estimated PM25 values were predicted from other PM indices,
18      including COH and extinction coefficient, making it difficult to compare straightforwardly the
19      relative importance of PM indices.
20
21      Sulfate and Hydrogen Ion
22           Sulfate and H+, markers of acidic components of PM, have been hypothesized to be
23      especially harmful components of PM (Lippmann and Thurston, 1996). The newly available
24      studies that examined sulfate are shown in Table 8-4; four of them also analyzed H+ data. The
25      sulfate concentrations ranged from 1.8 //g/m3 (Santa Clara County, CA) to 12.7 //g/m3 (three NJ
26      cities). Aside from the west versus east coast contrast, the higher levels observed in Toronto and
27      the three NJ cities are likely due to their study period coverage of the early 1980's, when sulfate
28      levels  were higher.  Sulfate explained 25 to 30% of PM25 mass in eastern U.S. and Canadian
29      cities,  but it was only 14% of PM2 5 mass in Santa Clara County, CA.  The mean H+level in the
30      Buffalo, NY study (36.4 nmol/m3) was much higher than the levels in Philadelphia, Detroit, or
31      Atlanta, in part because the Buffalo study covered the 1988 summer when summer-haze episodes

        April 2002                                 8-54        DRAFT-DO NOT QUOTE OR CITE

-------
 1      occurred. The H+ levels measured in the other three cities were low, especially in Atlanta, GA
 2      (where the mean concentration was reported to be 0.0 //g/m3). Even the mean H+ concentration
 3      for Detroit, MI (the H+ was actually measured in Windsor, a Canadian city a few miles from
 4      downtown Detroit), 8.8 nmol/m3, was low compared to the reported detection limit of
 5      15.1 nmol/m3 (Brook et al., 1997) for the measurement system used in the study. Note that the
 6      corresponding detection limit for sulfate was 3.6 nmol/m3 (or 0.34 //g/m3) and the mean sulfate
 7      level for Detroit was 54 nmol/m3 (or  5.2 //g/m3), so that the signal-to-noise ratio is expected to be
 8      higher for sulfate than for H+.  Thus, the ambient levels and possible relative measurement errors
 9      for these data should be considered in interpreting the results of the  studies listed in Table 8-4.
10           Sulfate was a statistically significant (at p< 0.05) predictor of mortality, at least in single
11      pollutant models, in: Toronto, CN; the 8 largest Canadian cities; Santa Clara County, CA;
12      Buffalo, NY; Philadelphia, PA; Newark, NJ; Camden, NJ; and Montreal, Quebec, but not in
13      Detroit, MI, Elizabeth, NJ, or Atlanta, GA. However, it should be noted that the relative
14      significance across the cities is influenced by the sample size (both the daily mean death counts
15      and number of days available), as well as the range of sulfate levels, and therefore should be
16      interpreted with caution. Figure 8-7 shows the excess risks (± 95%  CI) estimated per 5 //g/m3
17      increase in 24-h sulfate reported in these studies, compared to the earlier Six Cities Study result.
18      The largest estimate was seen for Santa Clara County, CA, but the wide confidence band
19      (possibly due to the small variance of the sulfate, since its levels were low) should be taken into
20      account.  Also, in the Santa Clara County analysis, the sulfate effect was eliminated once PM2 5
21      was included in the model, perhaps being indicative of sulfate mainly serving as a surrogate for
22      fine particles in general there.  In any case, more weight should be accorded to estimates from
23      other studies with narrower confidence bands.  In the other studies, the effect size estimates
24      mostly ranged from about 1 to 4% per 5  //g/m3 increase in 24-h sulfate.
25           The relative significance of sulfate and FT compared to other PM components varied from
26      city to city, as seen in Table 8-4.  Because each study included different combinations of
27      co-pollutants that had different extents of correlation with sulfate and because multiple mortality
28      outcomes were analyzed, it is difficult to assess the overall importance of sulfate across the
29      available studies. However, it can generally be seen that the associations were stronger in cities
30      where the sulfate and H+ levels were relatively high. For example, the Gwynn et al., 2000
31      finding for Buffalo, NY data that H+ and sulfate were most significantly associated with total

        April 2002                                 8-55        DRAFT-DO NOT QUOTE OR CITE

-------
                           Percent                (total mortality, unless otherwise noted)
                                           per 5 |jg/m3 increase in sulfate
        Schwartz et al. (1996)
                  Six Cities"

         Burnett et al. (1998)
            Toronto, Canada "

         Burnett et al. (2000)
                 8 Largest -
            Canadian Cities
              Fairiey (1999) _
            Santa Clara, Co. "

         Gwynn et al. (2000)
                Buffalo, NY "

         Klemm et al. (2000) _
                Atlanta, GA "

         Lipfert et al. (2000a) _
            Philadelphia, PA "

        Lippman et al. (2000)
                 Detroit, Ml "

            Tsai et a!. (2000)
                3 NJ Cities
                                 _2
                                  I
0
I
2
i
4
 i
6
i
8
i
10
 i
                   -•—— Newwark
                   I            Camden
                                                           Elizabeth
       Figure 8-7.  Excess risks estimated for sulfate per 5 Aig/m3 increase from the studies in
                   which both PM2 5 and PM10_2 5 data were available.
1      mortality may be in part due to the high acid aerosol levels in that data. Also, the fact that the

2      Lippmann et al. (2000) finding for Detroit, MI data on H+ and sulfate being less significantly

3      associated with mortality than the size-fractionated PM mass indices may be due to acidic

4      aerosols levels being mostly below the detection limit in that data.  In this case, it appears that the

5      Detroit PM components show mortality effects even without much acidic input.

6           In summary, assessment of new study results for individual chemical components of PM

7      suggest that an array of PM components (mainly fine particle constituents) were associated with

8      mortality outcomes, including: COH, EC, OC, sulfate, H+, and nitrate. The discrepancies seen

9      with regard to the relative significance of these PM components across studies may be in part due
       April 2002
      8-56
         DRAFT-DO NOT QUOTE OR CITE

-------
 1      to the difference in their concentrations.  This issue is further discussed below as part of the
 2      assessment of new studies involving source-oriented evaluation of PM components.
 3
 4      8.2.2.4.3 Source-Oriented Evaluations
 5           Several new studies have conducted source-oriented evaluation of PM components.
 6      In these studies, daily concentrations of PM components (i.e., trace elements) and gaseous
 7      co-pollutants were analyzed using factor analysis to estimate daily concentrations due to
 8      underlying source types (e.g., motor vehicle emissions, soil, etc.), which are weighted linear
 9      combinations of associated individual variables. The mortality outcomes were then regressed on
10      those factors (factor scores) to estimate the impact of source types, rather than just individual
11      variables.  These studies differ in terms of: specific objectives/focus, the size fractions from
12      which trace elements were extracted, and the way factor analysis was used (e.g., rotation). The
13      main findings from these  studies regarding the source-types identified (or suggested) and their
14      associations with mortality outcomes are summarized in Table 8-5.
15           The Laden et al. (2000) analysis of Harvard Six Cities data for 1979-1988 aimed to identify
16      distinct source-related fractions of PM25  and to examine each fraction's association with
17      mortality.  Fifteen elements in the fine fraction samples were routinely found above their
18      detection limits and included in the data analyses.  For each of the six cities, up to 5 common
19      factors were identified from among the 15 elements, using specific rotation factor analysis.
20      Using the Procrustes rotation (a type of oblique rotation), the projection of the single tracer for
21      each factor was maximized. This specification of the tracer element was based on:
22      (1) knowledge from previous source apportionment research; (2) the condition that regression of
23      total fine mass on that element must result in a positive coefficient; and (3) identifications of
24      additional local source factors that positively contributed to total fine mass regression.  Three
25      source factors were identified in all six cities: (1) a soil and crustal material factor with Si as a
26      tracer;  (2) a motor vehicle exhaust factor with Pb as a tracer; and, (3) a coal  combustion factor
27      with Se as a tracer.  City-specific analyses also identified a fuel combustion  factor (V), a salt
28      factor (Cl), and selected metal factors (Ni, Zn, or Mn).  For each city, a GAM Poisson regression
29      model, adjusting for trend/season, day-of-week, and smooth function of temperature/dewpoint,
30      was used to estimate impacts of each source type (using absolute factor scores) simultaneously.
31      Summary estimates across cities were obtained by combining the city-specific estimates, using

        April 2002                                 8-57       DRAFT-DO NOT QUOTE OR CITE

-------
           TABLE 8-5.  SUMMARY OF SOURCE-ORIENTED EVALUATIONS OF
               PARTICULATE MATTER COMPONENTS IN RECENT STUDIES
  Author, City
Source types identified (or suggested) and associated
variables
   Source types associated with mortality.
   Comments.
  Laden et. al., (2000)
  Harvard Six Cities
  1979-1988
  Mar et al. (2000).
  Phoenix, AZ
  1995-1997
Soil and crustal material'.  Si

Motor vehicle emissions'. Pb

Coal combustion'.  Se

Fuel oil combustion'. V

Salt:  Cl

Note: the trace elements are from PM2 5 samples



PM2 5 (fromDFPSS) trace elements:

Motor vehicle emissions and re-suspended road dust:
Mn, Fe, Zn, Pb,  OC, EC, CO, and NO2

Soil:  Al, Si, and Fe

Vegetative burning:  OC, and Kg (soil-corrected
potassium)

LocalSO2 sources: SO2

Regional sulfate: S
   Strongest increase in daily mortality associated
   with mobile source factor. Coal combustion factor
   was positively associated with mortality in all
   metropolitan areas, with exception of Topeka.
   Crustal factor from fine particles not associated
   with mortality.  Coal and mobile sources account
   for majority of fine particles in each city.
   PM, „ factors results:  Soil factor and local SO2
   factor were negatively associated with total
   mortality. Regional sulfate was positively
   associated with total mortality on the same day,
   but negatively associated on the lag 3 day. Motor
   vehicle factor, vegetative burning factor, and
   regional sulfate factor were significantly positively
   associated with cardiovascular mortality.
                           PMU_2S (from dichot)  trace elements:

                           Soil:  Al, Si, K, Ca, Mn, Fe, Sr, and Rb

                           A source of coarse fraction metals: Zn, Pb, and Cu

                           A marine influence:  Cl
                                                     analyzed for associations with mortality because of
                                                     small sample size (every-3ri day samples from
                                                     June 1996).
  Tsai et al. (2000).
  Newark, Elizabeth, and
  Camden, NJ.
  1981-1983.
Motor vehicle emissions: Pb, CO

Geological (Soil): Mn, Fe

Oil burning: V, Ni

Industrial:  Zn, Cu, Cd (separately)

Sulfate/'secondary aerosol:  sulfate

Note: the trace elements are from PM15 samples
   Oil burning, industry, secondary aerosol, and
   motor vehicles factors were associated with
   mortality.
  Ozkaynak et al. (1996).     Motor vehicle emissions: CO, COH, and NO2
  Toronto, Canada.
                                                     Motor vehicle factor was a significant predictor
                                                     for total, cancer, cardiovascular, respiratory, and
                                                     pneumonia deaths.	
April 2002
                                 8-58
DRAFT-DO NOT QUOTE OR CITE

-------
 1      inverse variance weights. The identified factors and their tracers are listed in Table 8-5.  The
 2      results from mortality regression analysis including these factors indicated that the strongest
 3      increase in daily mortality was associated with the mobile source factor.  Also, the coal
 4      combustion factor was positively associated with mortality in all metropolitan areas, except for
 5      Topeka. Lastly, S, Ni, and Pb were specific elements individually associated with mortality, but
 6      the crustal factor from fine particles was not.
 7           Mar et al. (2000) analyzed PM10, PM10_25, two measurements of PM25, and various
 8      sub-components of PM25 for their associations with total (non-accidental) and cardiovascular
 9      deaths in Phoenix, AZ during 1995-1997, using both individual PM components and factor
10      analysis-derived factor scores. GAM Poisson models were used, adjusting for season,
11      temperature, and relative humidity.  The evaluated air pollution variables included: O3, SO2,
12      NO2, CO, TEOM PM10, TEOM PM25, TEOM PM10.25, DFPSS PM25, S,  Zn, Pb, soil, soil-
13      corrected K (Ks), nonsoil PM, OC, EC,  and TC. Lags 0 to 4 days were evaluated.  As earlier
14      noted, individual PM component results indicated that PM10_2 5 was more significantly associated
15      with total mortality than PM2 5, although both TEOM PM2 5 and PM10_2 5 were significantly
16      associated with cardiovascular mortality. A factor analysis conducted on the chemical
17      components of DFPSS PM25 (Al, Si,  S,  Ca, Fe, Zn, Mn, Pb, Br, Ks, OC,  and EC) identified
18      factors for:  motor vehicle emissions/re-suspended road  dust; soil; vegetative burning; local SO2
19      sources; and regional sulfate (see Table  8-5). The results of mortality regression with these
20      factors suggested that the soil factor and local SO2 factor were negatively associated with total
21      mortality.  Regional sulfate was  positively associated with total mortality on the same day, but
22      negatively associated on the lag  3 day. The motor vehicle factor, vegetative burning factor, and
23      regional sulfate factor were each significantly positively associated with  cardiovascular mortality.
24      The authors also analyzed elements from dichot PM10_2 5 samples, and identified soil, a source of
25      coarse fraction metals (industry), and marine influence factors.  However, these factors were not
26      analyzed for their associations with mortality outcomes due to the short measurement period
27      (starting in June 1996 with every-3rd-day sampling).
28           It should be noted here that the Smith et al. (2000) analysis of Phoenix data also included
29      factor analysis on the elements from the coarse fraction and identified essentially the same
30      factors ("a source of coarse fraction metals" factor in Mar et al.'s study was called "the
31      anthropogenic elements" in Smith et al.'s study).  While Smith et al. did  not relate these factors

        April 2002                                 8-59         DRAFT-DO NOT QUOTE OR CITE

-------
 1      to mortality (due to a small sample size), they did show that the anthropogenic elements were
 2      low in summer and spring, when the PM10_2 5 effect was largest.  These results suggest that the
 3      PM10_2 5 effects were not necessarily due to anthropogenic components of the coarse particles,
 4      with biogenically-generated coarse particles perhaps being key during the warmer months (as
 5      noted earlier above).
 6           Tsai et al. (2000) conducted an exploratory analysis of mortality in relation to specific PM
 7      source types for three New Jersey cities (Camden, Newark, and Elizabeth) using factor analysis -
 8      Poisson regression techniques.  During the three-year study period (1981-1983), extensive
 9      chemical speciation data were available, including nine trace elements, sulfate,  and particulate
10      organic matter.  Total (excluding accidents and homicides), cardiovascular, and respiratory
11      mortality were analyzed. Tsai et al. first conducted a factor analysis of trace elements and
12      sulfate, identifying major source types:  motor vehicle (Pb, CO); geological (Mn, Fe); oil burning
13      (V, Ni); industrial (Zn, Cu); and sulfate/secondary aerosols (sulfate).  In addition to Poisson
14      regression of mortality on these factors, they also used an alternative approach in which the
15      inhalable particle mass (IPM, D50 < 15 //m) was first regressed on the factor scores of each of the
16      source types to apportion the PM mass;  and then the estimated daily PM mass for each source
17      type was included in Poisson regression, so that RR could be calculated per mass concentration
18      basis for each PM source type.  They found that oil burning (V, Ni), various industrial sources
19      (Zn, Cd), motor vehicle (Pb, CO), and the secondary aerosols, as well as the individual PM
20      indices IPM, FPM (D50 < 3.5 //m), and sulfates, were all associated with total and/or
21      cardiorespiratory mortality in Newark and Camden, but not in Elizabeth. In Camden, the RRs for
22      the source-oriented PM were higher (~ 1.10) than those for individual PM indices (« 1.02).
23           Ozkaynak et al. (1996) analyzed 21 years of mortality and air pollution data in Toronto,
24      Canada. In addition to the usual simultaneous inclusion of multiple pollutants in mortality
25      regressions, they also conducted a factor analysis of all the air pollution and weather variables,
26      including TSP, SO2, COH, NO2, O3,  CO, relative humidity and temperature. The factor with the
27      largest variance contribution (~ 50%) had the highest factor loadings for CO, COH, and NO2,
28      which they considered to be representative of motor vehicle emissions, since this pollution
29      grouping was also consistent with the emission inventory information for that city.  They then
30      regressed mortality on the factor scores (a linear combination of standardized scores for the
31      covariates), after filtering out seasonal cycles and adjusting for temperature and day-of-week

        April 2002                                 8-60        DRAFT-DO NOT QUOTE OR CITE

-------
 1      effects.  The estimated impacts on mortality from motor vehicle pollution ranged from 1 to 6%,
 2      depending on the outcomes.
 3           In summary, these studies suggest that a number of source-types are associated with
 4      mortality, including motor vehicle emissions, coal combustion, oil burning, and vegetative
 5      burning. The crustal factor from fine particles was not associated with mortality in the Harvard
 6      Six Cities data.  In Phoenix data, where coarse particles were reported to be associated with
 7      mortality, the associations between the factors related to coarse particles (soil, marine influence,
 8      and anthropogenic elements) and mortality could not be evaluated due to the  small sample size.
 9      However, the soil (i.e., crustal) factor from fine particles in the Phoenix data was negatively
10      associated with mortality. Thus, although some unresolved issues remain (mainly due to the lack
11      of sufficient data), the source-oriented evaluation approach, using factor analysis, thus far seems
12      to implicate fine particles of anthropogenic origin as being most important (versus crustal
13      particles of geologic origin) in contributing to observed increased mortality risks.
14
15      8.2.2.5  New Assessments of Cause-Specific Mortality
16           Consistent with similar findings described in the 1996 PM AQCD, most of the newly
17      available studies summarized in Tables 8-1 and 8A-1 that examined non-accidental total,
18      circulatory, and respiratory mortality categories (e.g., Samet et al., 2000a,b; Dominici et al.,
19      2000a; Moolgavkar, 2000a; Gwynn et al., 2000; Lippmann et al., 2000; Ostro et al., 1999a;
20      Schwartz, 2000c) found significant PM associations with both cardiovascular and/or respiratory-
21      cause mortality.  Several (e.g., Ostro et al., 1998; Fairley, 1999; Gwynn et al., 2000; Borja-
22      Aburto et al., 1997; Wordley et al., 1997; Borja-Aburto et al., 1998; Prescott  et al., 1998;)
23      reported estimated PM effects that were generally higher for respiratory deaths than for
24      circulatory or total deaths. Once again, the NMMAPS results for U.S. cities are among those of
25      particular note here due to the large study size and the combined, pooled estimates derived for
26      various U.S. regions.
27           The Samet et al. (2000a,b) NMMAPS 90-cities analyses not only examined all-cause
28      mortality (excluding accidents), but also evaluated cardiovascular, respiratory, and other
29      remaining causes of deaths.  Results were presented for all-cause, cardio-respiratory, and "other"
30      mortality for lag 0, 1, and 2 days. The investigators commented that, compared to the result for
31      cardio-respiratory deaths showing 3.5% (CI 1.0, 5.9) increase per 50 //g/m3 PM10, there was less

        April 2002                                  8-61         DRAFT-DO NOT QUOTE OR CITE

-------
 1      evidence for non-cardio-respiratory deaths. However, the estimates for "other" mortality, though
 2      half those for cardio-respiratory mortality, were nevertheless positive, with fairly high posterior
 3      probability (e.g., 0.84 at lag 0 day) that the overall effects were greater than 0 (estimated percent
 4      excess "other" deaths being ~ 1.3 per 50 //g/m3 PM10 at lag 0). Dominici et al. (2000a) evaluated
 5      the 20 largest U. S. cities, a subset of the cities included in Samet et al.'s NMMAPS analyses.
 6      The pattern of PM10 effects on cardiovascular and respiratory mortality was similar to that
 7      discussed earlier for total mortality, with lag day 1 showing the largest estimates. In this case,
 8      the PM10 effect in these analyses was smaller and weaker for "other" causes. Regional model
 9      results suggested that PM10 effects in the western U.S. were larger than in the eastern or southern
10      U.S. The PM coefficients were little affected by including gaseous pollutants in the model.
11           The Lippmann et al. (2000) analyses of cause-specific mortality in Detroit also evaluated
12      such mortality at various lags (0-3 days) in relation to several PM indices (PM10, PM25, PM10_2 5,
13      sulfate, H+) and various gaseous pollutants (O3, SO2, NO2 and CO), with appropriate adjustment
14      for season, temperature, relative humidity, etc.  Significant effects for both cardiovascular and
15      respiratory mortality were more consistently found for the first three PM indices than for H+ or
16      sulfate.  Effect size estimates tended to be highest for lag 1 day. It is notable here that, in the
17      Lippmann et al. (2000) analysis of Detroit mortality data, the  "other" mortality category also
18      showed statistically significant effect size estimates.  The authors noted, however, that the
19      "other" (non-circulatory and non-respiratory) mortality showed seasonal cycles and apparent
20      influenza peaks, suggesting that this series may have also been influenced by respiratory
21      contributing causes.
22           Another U.S. study, that of Moolgavkar (2000a),  evaluated possible PM effects on cause-
23      specific mortality across a broad range of lag (0-5 days) times. Moolgavkar reported that in
24      Poisson regression GAM analyses, controlling for temperature and relative humidity, varying
25      patterns of results were obtained for PM indices in evaluations of daily deaths related to
26      cardiovascular disease (CVD), cerebrovascular disease (CrD), and chronic obstructive lung
27      disease (COPD) in three large U.S. metropolitan areas. In Cook County (Chicago area), the
28      association of PM10 with CVD mortality was statistically significant at a lag of 3 days based on a
29      single-pollutant analysis and remained significantly associated with CVD deaths with a 3-day lag
30      in two pollutant models including one or another of CO, NO2, SO2, or O3. In joint analyses with
31      both O3 and SO2, however, the PM10 association became markedly reduced and non-significant.

        April 2002                                8-62         DRAFT-DO NOT QUOTE OR CITE

-------
 1      Also, in Los Angeles single-pollutant analyses, PM10 and PM2 5 were significantly associated with
 2      CVD mortality with lags of 2 and 1 days, respectively; but their coefficients were not robust to
 3      inclusion of one or more gaseous pollutants. In Maricopa Co., AZ, PM10 coefficients were large
 4      for several lags and significantly associated with CVD mortality lagged 1 day, as were each of
 5      the gaseous pollutants tested (except O3) at several different lag times; and PM10 coefficients
 6      seemed to be robust in 2-pollutant models including PM10 and NO2. As for cerebrovascular
 7      disease, Moolgavkar (2000) reported that there was little evidence of association for PM with
 8      CrD deaths at any lag in any of the three counties analyzed.  With regard to COPD deaths, PM10
 9      was significantly associated with COPD mortality (lag 2 days) in Cook County.
10           Zmirou et al. (1998) presented cause-specific mortality analyses results for 10 of the
11      12 APHEA European cities (APHEA1). Using Poisson autoregressive models adjusting for
12      trend, season, influenza epidemics, and weather, each pollutant's relative risk was estimated for
13      each city and "meta-analyses" of city-specific estimates were conducted.  The pooled excess risk
14      estimates for cardiovascular mortality were 1.0% (0.3, 1.7) per 25 //g/m3 increase in BS and 2.0%
15      (0.5, 3.0) per 50 //g/m3 increase in SO2 in western European cities. The pooled risk estimates for
16      respiratory mortality in the same cities were: 2.0% (0.8, 3.2) and 2.5% (1.5, 3.4) for BS and SO2,
17      respectively. Also of note, Wichmann et al. (2000) found significant associations of elevated
18      cardiovascular and respiratory disease mortality with various fine (and ultrafme) particle indices
19      evaluated in Erfurt, Germany.  "Other" natural causes (neither cardio- or respiratory-related)
20      almost always had the lowest risk in those models evaluating cause-specific mortality.
21           Seeking unique cause-specificity of effects associated with various pollutants has been
22      difficult because the "cause specific" categories examined are typically rather broad  (usually
23      cardiovascular and respiratory) and overlap; also cardiovascular and respiratory conditions tend
24      to occur together.  Examinations of more specific cardiovascular and respiratory sub-categories
25      may be necessary to test hypotheses about any specific mechanisms, but smaller sample sizes for
26      more specific sub-categories may make a meaningful analysis difficult. The study by Rossi et al.
27      (1999), however, examined associations between TSP and detailed cardio-vascular and
28      respiratory cause-specific mortality in Milan, Italy for a 9-year period (1980-1989).  They found
29      significant associations for respiratory infections (11% increase per 100 //g/m3 increase in TSP;
30      95%CI: 5, 17) and for heart failure (7%;  95%CI:  3, 11), both on the same day TSP.  The
31      associations with myocardial infarction (10%; 95%CI: 3,18) and COPD (12%; 95%CI:  6, 17)

        April 2002                                 8-63        DRAFT-DO NOT QUOTE OR CITE

-------
 1      were found for the average of 3 and 4 day TSP levels. They noted the difference in lags between
 2      temperature effects (i.e., cold temp, at lag 1 day for respiratory infections; hot temp, at lag 0 for
 3      heart failure and myocardial infarction) and air pollution (TSP) effects. The immediate hot
 4      temperature effects and the lagged cold temperature effects for total and cardiovascular mortality
 5      have been reported in past studies (e.g., Philadelphia, Chicago), but investigations of the
 6      differences in lags of PM effects for specific cardiovascular or respiratory categories have rarely
 7      been conducted in time-series mortality studies.
 8           In the Hoek et al. (2001) study of the whole population of the Netherlands, the large sample
 9      size (mean daily total deaths -330,  or more than twice that of Los Angeles County) allowed
10      examination of specific cardiovascular cause of deaths.  Deaths due to heart failure,  arrhythmia,
11      cerebrovascular causes, and thrombocytic causes were more strongly (-2.5 to 4 times larger
12      relative risks) associated with air pollution than the overall cardiovascular deaths. The
13      investigators concluded that specific cardiovascular causes (such as heart failure) were more
14      strongly associated with air pollution than total cardiovascular mortality, but noted that the
15      largest contribution to the association between air pollution and cardiovascular mortality was
16      from ischemic heart disease (about half of all cardiovascular deaths).
17           An HEI report on an epidemiologic study conducted by Goldberg et al. (2000) in Montreal,
18      Canada also provides interesting new information regarding types of medical conditions putting
19      susceptible individuals at increased risk for PM-associated mortality effects; and it highlights the
20      potential importance of evaluating "contributing causes" in cause-specific mortality analyses.
21      First, the immediate causes of death, as listed on death certificates, were evaluated in relation to
22      various ambient PM indices (TSP, PM10, PM2 5, COH, sulfates, extinction coefficients) lagged for
23      0 to 4 days, with results reported emphasizing effects at 3 day lags for three main PM measures
24      (COH, sulfate, estimated PM2 5).  Significant associations were observed between all three
25      measures and total nonaccidental deaths, respiratory diseases, and diabetes, with an approximate
26      2% increase in excess nonaccidental mortality being observed per 9.5 //g/m3 interquartile
27      increase in 3-day mean estimated PM25 exposure.
28           When underlying clinical conditions identified in decedents' medical records were then
29      evaluated in relation to ambient PM measures, all three measures (COH, sulfate, estimated PM2 5)
30      were associated with acute lower respiratory disease, congestive heart failure, and any
31      cardiovascular disease. Estimated PM2 5 and COH were also reported to be associated with

        April 2002                                 8-64        DRAFT-DO NOT QUOTE OR CITE

-------
 1      chronic coronary artery disease, any coronary artery disease, and cancer; whereas, sulfate was
 2      associated with acute and chronic upper respiratory disease.  None of the three PM measures
 3      were related to airways disease, acute coronary artery disease, or hypertension. These results
 4      both tend to support previous findings identifying individuals with preexisting cardiopulmonary
 5      diseases as being at increased risk for ambient PM effects and appear to implicate another risk
 6      factor,  diabetes (which typically also involves cardiovascular complications as it progresses), as a
 7      possible susceptibility condition putting individuals at increased risk for ambient PM effects.
 8           Two recent studies (Gouveia and Fletcher, 2000; Concei9ao et al., 2001), both using data
 9      from Sao Paulo, Brazil, examined child mortality (age under 5 years).  The study periods for
10      these studies did not overlap (1991-1993 for Concei9ao and Fletcher study;  1994-1997 for
11      Concei9ao). Although Gouveia and Fletcher found significant associations between air pollution
12      and elderly mortality, they did not find statistically significant associations between air pollution
13      and child respiratory mortality (PM10 coefficient was negative and not significant).  In the
14      Concei9ao et al. (2001) analysis, significant associations were found between child  respiratory
15      mortality and CO, SO2, and PM10 in single pollutant models, and coefficients for CO and SO2
16      remained significant in the multiple-pollutant (apparently all pollutants together) model. The
17      reported PM10 coefficient in the single pollutant model corresponds to percent excess respiratory
18      death of 7.1% (95% CI: 1.1, 13.7) per 50 Mg/m3 increase in PM10.  However, it should be noted
19      that the average daily respiratory mortality counts for these studies were relatively small
20      (~2.4/day). With the modest length of observations (3 years for Gouveia and Fletcher study, and
21      4 years for Concei9ao  et al.'s study), the statistical  power of the data were likely less than
22      desirable.  Thus, there have not been enough data to elucidate the  range of short-term PM effects
23      on child (respiratory) mortality.
24           Overall, then, the above assessment of newly available information provides interesting
25      additional new information (beyond that in the 1996 PM AQCD) with regard to cause-specific
26      mortality related to ambient PM. That is, a growing number of studies continue to report
27      increased cardiovascular- and respiratory-related mortality risks as being significantly associated
28      with ambient PM measures at one or another varying lag times.  When specific subcategory of
29      cardiovascular disease was examined in a large population (The Netherlands study by Hoek
30      et al.),  some of the subcategories such as heart failure were more strongly associated with PM
31      and other pollutants than total cardiovascular mortality. Largest effects estimates are most

        April 2002                                 8-65        DRAFT-DO NOT QUOTE OR CITE

-------
 1      usually reported for 0-1 day lags (with some studies also now noting a second peak at 3-4 day
 2      lags). A few of the newer studies also report associations of PM metrics with "other" (i.e.,
 3      non-cardiorespiratory) causes, as well. However, at least some of these "other" associations may
 4      also be due to seasonal cycles that include relationships to peaks in influenza epidemics that may
 5      imply respiratory complications as a "contributing" cause to the "other" deaths.  Or, the "other"
 6      category may include sufficient numbers of deaths due to diabetes or other diseases which may
 7      also involve cardiovascular complications as contributing causes.  Varying degrees of robustness
 8      of PM effects are seen in the newer studies, as typified by estimates in multiple pollutant models
 9      containing gaseous co-pollutants; many show little effect of gaseous pollutant inclusion on
10      estimated PM effect sizes, some show larger reductions in PM effects to non-significant levels
11      upon such inclusion, and a growing number also report significant associations of cardiovascular
12      and respiratory effects with one or more gaseous co-pollutants.  Thus, the newer studies both
13      further substantiate PM effects on cardiovascular- and respiratory-related mortality, while also
14      pointing toward possible significant contributions of gaseous pollutants to such cause-specific
15      mortality, as well. The magnitudes of the PM effect size estimates are consistent with the range
16      of estimates derived from the few earlier available studies assessed in the 1996 PM AQCD.
17
18      8.2.2.6  Salient Points Derived from Summarization of Studies  of Short-Term Particulate
19              Matter Exposure Effects on Mortality
20           The most salient key points to be extracted from the above discussion of newly available
21      information on short-term PM exposures relationships to mortality can be summarized as follow:
22
23      PM10 effects estimates. Since the 1996 PM AQCD, thus far, there have been more than 80 new
24      time-series  PM-mortality analyses published. Estimated mortality relative risks in these studies
25      are generally positive, statistically  significant, and consistent with the previously reported
26      PM-mortality associations. Of particular importance are several studies which evaluated
27      multiple cities using consistent data analytical approaches.  The NMMAPS analyses for the
28      largest 90 U.S. cities (Samet et al., 2000a,b), which are thought to probably provide the most
29      precise estimates for PM10 effects applicable to the U.S., derived a combined nationwide excess
30      risk estimate of about 2.3% increase in total (non-accidental) mortality per 50 //g/m3 increase in
31      PM10. The  other multi-city analyses, as well as various single city analyses, also obtained PM10

        April 2002                                  8-66        DRAFT-DO NOT QUOTE OR CITE

-------
 1      effect sizes generally in the range of 1.5 to 8.5% per 50 //g/m3 increase in PM10, consistent with
 2      the range of statistically significant estimates given in the 1996 PM AQCD. However, more
 3      geographic heterogeneity is evident among the newer multi-city study results than was the case
 4      among the fewer studies assessed in the 1996 PM AQCD. In particular, in the NMMAPS
 5      analysis of the 90 largest U.S. cities data, the risk estimates varied by U.S. geographic region,
 6      with the estimate for the Northeast being the largest (4.6% per 50 //g/m3 PM10 increase).  The
 7      observed heterogeneity in the estimated PM risks across cities/regions could not be explained
 8      with the city-specific explanatory variables, such as the mean levels of pollution and weather,
 9      mortality rate, sociodemographic variables (e.g., median household income), urbanization, or
10      variables related to measurement error. Notable apparent heterogeneity was also seen among
11      effects estimates for PM (and SO2) indices in the multi-city APHEA studies conducted in
12      European cities.  In APHEA2, they found that several city-specific characteristics, such as NO2
13      levels and warm climate, were important effect modifiers. The issue of heterogeneity of effects
14      estimates is discussed further below in  Section 8.4.
15
16      Confounding and effect modification by other pollutants. Numerous new short-term PM
17      exposure studies not only  continue to report significant associations between various PM indices
18      and mortality, but also between gaseous pollutants (O3, SO2, NO2, and CO)  and mortality as well.
19      In most of these studies, simultaneous inclusions of gaseous pollutants in the regression models
20      did not meaningfully affect the PM effect size estimates. This was the case for the NMMAPS 90
21      cities study with regard to the overall combined U.S. regional and nationwide risk estimates
22      derived for that  study.  The issue of confounding is discussed further in Section 8.4.
23
24      Fine and coarse particle effects. Newly available studies provide generally statistically
25      significant PM2 5 associations with mortality, with effect size estimates falling in the range
26      reported in the 1996 PM AQCD. New results from Germany appear to implicate both ultrafine
27      (nuclei-mode) and accumulation-mode fractions of urban ambient fine PM as being important
28      contributors to increased mortality risks.  As to the relative importance of fine and coarse
29      particles, in the  1996 PM AQCD there was only one acute mortality study that examined this
30      issue. In that study, the authors suggested that fine particles (PM25), but not coarse particles
31      (PM10_2 5), were associated with daily mortality. Now, more than ten studies have analyzed both

        April 2002                                8-67        DRAFT-DO NOT QUOTE OR CITE

-------
 1      PM25 and PM10_2 5 for their associations with mortality. While the results from some of these new
 2      studies (e.g., Santa Clara County, CA analysis [Fairley, 1999] and the largest 8 Canadian cities
 3      analysis [Burnett et al., 2000]) did suggest that PM2 5 was more important than PM10_2 5 in
 4      predicting mortality fluctuations, other studies (e.g., Phoenix, AZ analyses  [Clyde et al., 2000;
 5      Mar et al., 2000; Smith et al., 2000]; Mexico City and Santiago, Chile studies [Castillejos et al.,
 6      2000; Cifuentes et al., 2000]) suggest that PM10_2 5 may also be important in at least some
 7      locations. Seasonal dependence of size-related PM component effects observed in some of the
 8      studies complicates interpretations.
 9
10      Chemical components ofPM. Several new studies have examined the role of specific chemical
11      components of PM. The studies conducted in U.S. and Canadian cities showed mortality
12      associations with specific fine particle components of PM including H+, sulfate,  nitrate, as well
13      as COH, but their relative importance varied from city to city, likely depending on their levels
14      (e.g., no clear associations in those cities where H+ and sulfate levels were very low, i.e., circa
15      non-detection limits).  The results  of several studies that investigated the role of crustal particles,
16      although somewhat mixed, do not appear overall to support associations between crustal particles
17      and mortality (see also the discussion of source-oriented evaluations presented below).
18
19      Source-oriented evaluations. Several studies conducted source-oriented evaluations of PM
20      components using factor analysis.  The results from these studies generally indicate that several
21      combustion-related source-types are likely associated with mortality, including:  motor vehicle
22      emissions; coal combustion; oil burning; and vegetative burning. The crustal factor from fine
23      particles was not associated with mortality in the Harvard Six Cities data, and the soil (i.e.,
24      crustal) factor from fine particles in the Phoenix data was negatively associated with mortality.
25      Thus, the source-oriented evaluations seem to implicate fine particles of anthropogenic origin as
26      being most important as contributing to increased mortality and generally do not support
27      increased mortality risks being related to short-term exposures to crustal materials in U.S.
28      ambient environments examined to date.
29
30      Cause-specific mortality. Findings for new results concerning cause-specific mortality comport
31      well with those for total (non-accidental) mortality, the former showing generally larger effect

        April 2002                                 8-68         DRAFT-DO NOT QUOTE OR CITE

-------
 1      size estimates for cardiovascular, respiratory, and/or combined cardiorespiratory excess risks
 2      than for total mortality risks.  An analysis of specific cardiovascular causes in a large population
 3      (The Netherlands) suggested the specific causes of deaths such as heart failure was more strongly
 4      associated with PM (and other pollutants) than total  cardiovascular mortality.
 5
 6      Lags.  In general, maximum effect sizes for total mortality appear to be obtained with 0-1 day
 7      lags, with some studies finding a second peak for 3-4 days lags. There is also some evidence
 8      that, if effects distributed over multiple lag days are considered, the effect size may be larger than
 9      for any single maximum effect size lag day. Lags are discussed further in Section 8.4.
10
11      Threshold. Few new short-term mortality studies explicitly address the issue of thresholds.  One
12      study that analyzed Phoenix, AZ data (Smith et al., 2000) did report some limited evidence
13      suggestive of a possible threshold for PM25 there.  However, several different analyses of larger
14      PM10 data  sets across multiple cities (Dominici, et al., 2002; Daniels et al., 2000) generally
15      provide little or no support to indicate a threshold for PM10 mortality effects.  Threshold issues
16      are discussed further in Section 8.4.
17
18      8.2.3  Mortality Effects of Long-Term Exposure to Ambient Particulate
19             Matter
20      8.2.3.1  Studies Published Prior to the 1996 Particulate Matter Criteria Document
21      8.2.3.1.1  Aggregate Population Cross-Sectional Chronic Exposure Studies
22           Mortality effects associated with chronic, long-term exposure to ambient PM have been
23      assessed in cross-sectional studies and, more recently, in prospective cohort studies.  A number
24      of older cross-sectional studies from the 1970s provided indications of increased mortality
25      associated with chronic (annual average) exposures to ambient PM, especially with respect to
26      fine mass or sulfate (SO4=) concentrations. However, questions unresolved at that time regarding
27      the adequacy of statistical adjustments for other potentially important covariates (e.g., cigarette
28      smoking, economic status, etc.) across cities tended to limit the degree of confidence that was
29      placed by the 1996 PM AQCD (U.S. Environmental Protection Agency, 1996a) on such purely
30      "ecological" studies or on quantitative estimates of PM  effects derived from these studies.
31      Evidence comparing the  toxicities of specific PM  components was relatively limited. The sulfate

        April 2002                                8-69        DRAFT-DO NOT QUOTE OR  CITE

-------
 1      and acid components had already been discussed in detail in the previous PM AQCD (U.S.
 2      Environmental Protection Agency, 1986).
 3
 4      8.2.3.1.2  Semi-Individual (Prospective Cohort) Chronic Exposure Studies
 5           Semi-individual cohort studies using subject-specific information about relevant covariates
 6      (such as cigarette smoking, occupation, etc.) have provided more certain findings of long-term
 7      PM exposure effects than past purely "ecological studies" (Kiinzli and Tager, 1997). At the same
 8      time, these better designed cohort studies have largely confirmed the magnitude of PM effect
 9      estimates from past cross-sectional study results.
10           Prospective cohort semi-individual studies of mortality associated with chronic exposures
11      to air pollution of outdoor origins have yielded especially valuable insights into the adverse
12      health effects of long-term PM exposures. The extensive Harvard Six-Cities Study (Dockery
13      et al., 1993) and the American Cancer Society (ACS) Study (Pope et al., 1995) agreed in their
14      findings of statistically significant positive associations between fine particles and excess
15      mortality, although the ACS study did not evaluate the possible contributions of other air
16      pollutants. Neither study considered multi-pollutant models, although the Six-City study did
17      examine various gaseous and particulate matter indices (including total particles, PM25, SO4=, H+,
18      SO2, and ozone), finding that sulfate and PM25 fine particles were best associated with mortality.
19      The excess RR estimates for total mortality in the Six-Cities study (and 95 percent confidence
20      intervals, CI) per increments in PM indicator levels were: Excess RR=15% (CI=6.1%, 32%) for
21      20 //g/m3 PM10; excess RR=11.4% (CI=4.3%, 23% )for 10 //g/m3 PM25; and excess RR=13.4%
22      (CI=5.1%, 29%) for 5 //g/m3 SO4=.  The estimates for total mortality derived from the ACS study
23      were excess RR=6.5% (CI=3.5%, 9.7%) for 10 //g/m3 PM25 and excess RR 3.5% (CI=1.9%,
24      5.1%) for 5 //g/m3 SO4=.  The  ACS pollutant RR estimates were smaller than those from the
25      Six-Cities study, although their 95% confidence intervals overlap. In some cases in these studies,
26      the life-long cumulative exposure of the study cohorts included distinctly higher past PM
27      exposures, especially in cities with historically higher PM levels (e.g., Steubenville, OH); but
28      more current PM measurements were used to estimate the chronic PM exposures.  In the ACS
29      study, the pollutant exposure estimates were based on concentrations at the  start of the  study
30      (during 1979-1983). Also, the average age of the ACS cohort was 56, which could overestimate
31      the pollutant RR estimates and perhaps underestimate the life-shortening associated with PM

        April 2002                                8-70        DRAFT-DO NOT  QUOTE OR CITE

-------
 1      associated mortality.  Still, although caution must be exercised regarding the use of the reported
 2      quantitative risk estimates, the Six-Cities and ACS semi-individual studies provided consistent
 3      evidence of a significant mortality association with long-term exposure to PM of ambient origins.
 4           In contrast to the Six-Cities and ACS studies, early results reported by Abbey et al. (1991)
 5      and Abbey et al. (1995a) from the Adventist Health Study on Smog (AHSMOG) found no
 6      significant mortality effects of previous PM exposure in a relatively young cohort of California
 7      nonsmokers. However, these analyses  used TSP as the PM exposure metric, rather than more
 8      health relevant PM metrics such as PM10 or PM25, included fewer subjects than the ACS study,
 9      and considered a shorter follow-up time than the Six-Cities study (ten years vs.  15 years for the
10      Six-Cities study).  Moreover, the AHSMOG study included only non-smokers, indicated by the
11      Six-Cities Study as having lower pollutant RR's than smokers, suggesting that a longer follow-up
12      time than considered in the past (10 years) might be required to have sufficient power to detect
13      significant pollution effects than is required in studies that include smokers (such as the
14      Six-Cities and ACS studies).  Thus, greater emphasis has been placed thus far on the Six-Cities
15      and ACS studies.
16           Overall, the previously available  chronic PM exposure studies collectively indicated that
17      increases in mortality are associated with long-term exposure to ambient airborne particles.
18      Also, effect size estimates for total mortality associated with chronic PM exposure indices
19      appeared to be much larger than those reported from daily mortality PM studies. This  suggested
20      that a major fraction of the reported mortality relative risk estimates associated with chronic PM
21      exposure likely reflects cumulative PM impacts above and beyond those exerted by the sum of
22      acute exposure events (i.e., assuming that the latter are fully additive over time). The 1996 PM
23      AQCD (Chapter  12) reached several conclusions concerning four key questions about  the
24      prospective cohort studies, as noted below:
25
26           (1) Have potentially important confounding variables been omitted?
27                 "While it is not likely that the prospective cohort studies have overlooked plausible
28           confounding factors that can account for the large effects attributed to air pollution, there
29           may be some further adjustments in the estimated magnitude of these effects as individual
30           and community risk factors are included in the analyses." These include individual
31           variables such  as education, occupational exposure to dust and fumes, and physical activity,

        April 2002                                8-71        DRAFT-DO NOT QUOTE OR CITE

-------
 1           as well as ecological (community) variables such as regional location, migration, and
 2           income distribution.  Further refinement of the effects of smoking status may also prove
 3           useful."
 4
 5           (2)  Can the most important pollutant species be identified?
 6                 "The issue of confounding with co-pollutants has not been resolved for the
 7           prospective cohort studies . . . Analytical strategies that could have allowed greater
 8           separation of air pollutant effects have not yet been applied to the prospective cohort
 9           studies." The ability to separate the effects of different pollutants, each measured as a long-
10           term average on a community basis, was clearly most limited in the Six Cities study.  The
11           ACS study offered a much larger number of cities, but did not examine differences
12           attributable to the spatial and temporal differences in the mix of particles and gaseous
13           pollutants across the cities.  The AHSMOG study constructed time- and location-dependent
14           pollution metrics for most of its participants that might have allowed such analyses, but no
15           results were reported.
16
17           (3)  Can the time scales for long-term exposure effects be evaluated?
18                 "Careful review of the published studies indicated a lack of attention to this issue.
19           Long-term mortality studies have the potential to infer temporal relationships based on
20           characterization of changes in pollution levels over time. This potential was greater in the
21           Six Cities and AHSMOG studies because of the greater length of the historical air pollution
22           data for the cohort [and the availability of air pollution data throughout the study].  The
23           chronic exposure studies, taken together, suggest that there may be increases in mortality in
24           disease categories that are consistent with long-term exposure to airborne particles, and that
25           at least some fraction of these deaths are likely to occur between acute exposure episodes.
26           If this interpretation is correct, then at least some individuals may experience some years of
27           reduction of life as a consequence of PM exposure."
28
29           (4)  Is it possible to identify pollutant thresholds that might be helpful in health
30           assessments?


        April 2002                                8-72         DRAFT-DO NOT QUOTE OR CITE

-------
 1                 "Model specification searches for thresholds have not been reported for prospective
 2           cohort studies. . . .  Measurement error in pollution variables also complicates the search
 3           for potential threshold effects. . . .  The problems that complicate threshold detection in the
 4           population-based studies have a somewhat different character for the long-term studies."
 5
 6      8.2.3.2 Prospective Cohort Analyses of Chronic Particulate Matter Exposure Mortality
 7             Effects Published Since the 1996 Particulate Matter Air Quality Criteria Document
 8           Considerable progress has been made towards addressing further the above issues.
 9      For example, extensive reanalyses (Krewski et al., 2000) of the Six-Cities and ACS Studies,
10      conducted under sponsorship by the Health Effects Institute (HEI), indicate that the published
11      findings of the original investigators (Dockery et al., 1993; Pope et al., 1995) are based on
12      substantially valid data sets and statistical analyses. The HEI reanalysis project has demonstrated
13      that small corrections in input data have very little effect on the findings and that alternative
14      model specifications further substantiate the robustness of the originally reported findings.
15      In addition, some of the above key questions have been further investigated by Krewski et al.
16      (2000) via sensitivity analyses (in effect, new analyses) for the Six City and ACS studies data
17      sets, including consideration of a much wider range of confounding variables.  Newly published
18      analyses of ACS data for more extended time periods (Pope et al., 2002) further substantiate
19      original findings and, also, provide much clearer, stronger evidence for ambient PM exposure
20      relationships with increased lung cancer risk.  Recently published analyses of AHSMOG data
21      (Abbey et al.,  1999;  Beeson et al., 1998) also extend the ASHMOG findings and show some
22      analytic outcomes different from earlier analyses reported out from the study.  Results from the
23      Veterans' Administration- Washington University (hereafter called "VA") prospective cohort
24      study are now available (Lipfert et al., 2000).  Still other, additional new studies suggestive of
25      possible effects  of sub-chronic PM exposures on infant mortality (Woodruff et al., 1997;  Bobak
26      and Leon,  1998; Lipfert, 2000; Chen et al., 2002) are also discussed below.
27
28      8.2.3.2.1  Health Effects Institute Reanalyses of the Six-Cities  and ACS Studies
29           The overall objective of the HEI "Particle Epidemiology Reanalysis Project" was to
30      conduct a rigorous and independent assessment of the findings of the Six Cities (Dockery et al.,
31      1993) and  ACS  (Pope et al., 1995) Studies of air pollution and mortality.  The following

        April 2002                                8-73        DRAFT-DO NOT QUOTE OR CITE

-------
 1      description of approach, key results, and conclusions is largely extracted from the Executive
 2      Summary of the HEI final report (Krewski et al., 2000). The HEI-sponsored reanalysis effort
 3      was approached in two steps:
 4           •Parti: Replication and Validation. The Reanalysis Team sought to test: (a) if the
 5             original studies could be replicated via a quality assurance audit of a sample of the
 6             original data and; (b) if the original numeric results could be validated.
 7           • Part II: Sensitivity Analyses.  The Reanalysis Team tested the robustness of the original
 8             analyses to alternate risk models and analytic approaches.
 9           The Part I audit of the study population data for both the Six Cities and ACS Studies and of
10      the air quality data in the Six Cities Study revealed the  data to be of generally high quality with
11      few exceptions. In both studies, a few errors were found in the data coding for and exclusion of
12      certain subjects; when those subjects were included in the analyses, they did not materially
13      change the  results from those originally reported.  Because the air quality data used in the ACS
14      Study could not be audited, a separate air quality database was constructed for the sensitivity
15      analyses in Part n.
16           The Reanalysis Team was able to replicate the original results for both studies using the
17      same data and statistical methods as used by the Original Investigators. The Reanalysis Team
18      confirmed the original point estimates, as  shown in  Table 8-6. For the Six Cities Study, they
19      reported the relative risk of mortality from all causes associated with an increase in fine particles
20      of 20.0 //g/m3 as 1.28, the same  as the 1.28 per 20 //g/m3 reported by the Original Investigators.
21      For the ACS Study, the relative risk of all-cause mortality associated with a 20 //g/m3 increase in
22      fine particles was 1.19 in the reanalysis, close to the original 1.14 value.
23           The Part II sensitivity analysis applied an array of different models and variables to
24      determine whether the original results would remain robust to different analytic assumptions and
25      model specifications.  The Reanalysis Team first applied the standard Cox model used by the
26      Original Investigators and included variables in the  model  for which data were available from
27      both original studies, but had not been used in the published analyses (e.g. physical activity, lung
28      function, marital status). The Reanalysis Team also designed models to include interactions
29      between variables. None of these alternative models produced results that materially altered the
30      original findings.
31

        April 2002                                 8-74        DRAFT-DO NOT QUOTE OR CITE

-------
         TABLE 8-6. COMPARISON OF SIX CITIES AND AMERICAN CANCER SOCIETY
           (ACS) STUDY FINDINGS FROM ORIGINAL INVESTIGATORS AND HEALTH
                                 EFFECTS INSTITUTE REANALYSIS
Type of Health
Effect & Location
Original Investigators'
Six City1'
Six City11
ACS Study
HEI reanalysis Phase I
Indicator
Findings
PM25
PM15/10
PM2S
: Replication
Mortality Risk per Increment in PMa
Total mortality
Excess Relative Risk (95% CI)
13% (4.4%, 23%)
18% (6%, 32%)
6.8% (3.4%, 10%)

Cardiopulmonary mortality
Excess Relative Risk (95% CI)
17% (5.8%, 42%)
e
11.8% (6.8%, 17%)

Six City Reanalysis4
                     PM15
ACS Study Reanalysis4   PM2S
                     PM15 (dichot)
                     PM15 (SSI)
                                                  11.3% (3%, 23%)
                                                  18% (6%, 34%)
                                                9.1% (3.9%, 14.5%)
                                                   4% (1%, 7%)
                                                  2% (-1%, 4%)
                          18.7% (6.3%, 33%)
                           20% (2%, 41%)
                          15.3% (9.1%, 21%)
                            7% (2%, 12%)
                            6% (3%, 9%)
         "Estimates calculated on the basis of differences between the most-polluted and least-polluted cities, scaled to
         increments of 20 /^g/m3 increase for PM10, and 10 /ug/m3 increments for PM15 and PM2 s.
         bDockery et al. (1993).
         cPope et al. (1995).
         dKrewski et al. (2000).
         'Data presented only by smoking subgroup.
 1           Next, for both the Six Cities and ACS Studies, the Reanalysis Team investigated the
 2      possible effects of fine particles and sulfate on a range of potentially susceptible subgroups of the
 3      population. These analyses did not find differences in PM-mortality associations among
 4      subgroups based on various personal characteristics (e.g., including gender, smoking status,
 5      exposure to occupational dusts and fumes, and marital status). However, estimated effects of
 6      fine particles did vary with educational level; the association between an increase in fine particles
 7      and mortality tended to be higher for individuals without a high school education than for those
 8      with more education. The Reanalysis Team postulated that this finding could be attributable to
 9      some unidentified socioeconomic effect modifier. The authors concluded, "The Reanalysis
10      Team found little evidence that questionnaire variables had led to confounding in either study,
11      thereby strengthening the conclusion that the observed association between fine particle air
12      pollution and mortality was not the result of a critical covariate that had been neglected by the
13      Original Investigators." (Krewski et al., 2000, pp. 219-220).
        April 2002
8-75
                                                      DRAFT-DO NOT QUOTE OR CITE

-------
 1           In the ACS study, the Reanalysis Team tested whether the relationship between ambient
 2      concentrations and mortality was linear. They found some indications of both linear and
 3      nonlinear relationships, depending upon the analytic technique used, suggesting that the shapes
 4      of the concentration-response relationships warrant additional research in the future.
 5           One of the criticisms of both original studies has been that neither analyzed the effects of
 6      change in pollutant levels over time. In the Six Cities Study, for which such data were available,
 7      the Reanalysis Team tested whether effect estimates changed when certain key risk factors
 8      (smoking, body mass index, and air pollution) were allowed to vary over time. In general, the
 9      reanalysis results did not change when smoking and body mass index were allowed to vary over
10      time.  The Reanalysis Team did find for the Six Cities Study, however, that when the general
11      decline in fine particle levels over the monitoring period was included as a time-dependent
12      variable, the association between fine particles and all-cause mortality was reduced (Excess
13      RR = 10.4%,  [1.5%, 20%]). This would be expected, since the most polluted cities would be
14      expected to have the greatest decline as pollution controls were applied. Despite this adjustment,
15      the PM25 effect estimate continued to be positive and statistically significant.
16           To test the validity of the original ACS air quality data, the Reanalysis Team constructed
17      and applied its own air quality dataset from available historical data. In particular, sulfate levels
18      with  and without adjustment were  found to differ by about 10% for the Six Cities Study. Both the
19      original ACS Study air quality data and the newly constructed dataset contained  sulfate levels
20      inflated by approximately 50% due to artifactual sulfate. For the Six Cities Study, the relative
21      risks of mortality were essentially unchanged with adjusted or unadjusted sulfate. For the ACS
22      Study, adjusting for artifactual sulfate resulted in slightly higher relative risks of mortality from
23      all causes and cardiopulmonary disease compared with unadjusted data, while the relative risk of
24      mortality from lung cancer was lower after the data had been adjusted.  Thus, the Reanalysis
25      Team found essentially the same results as the original Harvard Six-Cities and ACS studies, even
26      after using independently developed pollution datasets and after adjusting for sulfate artifact.
27           Because of the limited statistical power to conduct most model specification sensitivity
28      analyses for the Six Cities Study, the Reanalysis Team conducted the majority of its sensitivity
29      analyses using only the ACS Study dataset that considered 151 cities. When a range of city-level
30      (ecologic) variables (e.g., population change, measures of income, maximum temperature,
31      number of hospital beds, water hardness) were included in the analyses, the results generally did

        April 2002                                 8-76        DRAFT-DO NOT QUOTE OR CITE

-------
 1      not change.  The only exception was that associations with fine particles and sulfate were
 2      reduced when city-level measures of population change or SO2 were included in the model.
 3           A maj or product of the Reanalysis Proj ect is the determination that both pollutant variables
 4      and mortality appear to be spatially correlated in the ACS Study dataset. If not identified and
 5      modeled correctly, spatial correlation could cause substantial errors in both the regression
 6      coefficients and their standard errors. The Reanalysis Team identified several methods for
 7      addressing this, each of which resulted in some reduction in the estimated regression coefficients.
 8      The full implications and interpretations of spatial correlations in these analyses have not been
 9      resolved, and were noted to be an important subject for future research.
10           When the Reanalysis Team sought to take into account both the underlying variation from
11      city to city (random effects) and variation from the spatial correlation between cities, associations
12      were still found between mortality and sulfates or fine particles. Results of various models, using
13      alternative methods to address spatial autocorrelation  and including different ecologic covariates,
14      found fine particle-mortality associations that ranged from 1.11 to 1.29 (RR reported by original
15      investigators was 1.17) per 24.5 //g/m3 increase in PM25. With the exception of SO2,
16      consideration of other pollutants in these models did not alter the associations found with
17      sulfates.  The authors reported associations that were stronger for SO2 than for sulfate, which
18      may indicate that the sulfate with artifact was "picking up" some of the SO2 association, perhaps
19      because the artifact is in part proportional to the prevailing SO2 concentration (Coutant, 1977).
20      It should be recognized that the Reanalysis Team did not use data adjusted for artifactual sulfate
21      for most alternative analyses. When they did use adjusted sulfate data, relative risks of mortality
22      from all causes and cardiopulmonary disease increased. This result suggests that more analyses
23      with adjusted sulfate might result in somewhat higher relative risks  associated with sulfate.  The
24      Reanalysis Team concluded: "it suggests that uncontrolled spatial autocorrelation accounts for
25      24% to 64% of the observed relation. Nonetheless, all our models continued to show an
26      association between elevated risks of mortality and exposure to airborne sulfate" (Krewski et al.,
27      2000, p. 230).
28           In summary, the reanalyses generally confirmed the original investigator's findings of
29      associations between mortality and long-term exposure to PM, while recognizing that increased
30      mortality may be attributable to  more than one ambient air pollution component.  Regarding the
31      validity of the published Harvard Six-Cities and ACS Studies, the HEI Reanalysis Report

        April 2002                                 8-77         DRAFT-DO NOT QUOTE OR CITE

-------
 1      concluded that: "Overall, the reanalyses assured the quality of the original data, replicated the
 2      original results, and tested those results against alternative risk models and analytic approaches
 3      without substantively altering the original findings of an association between indicators of
 4      particulate matter air pollution and mortality."
 5
 6      8.2.3.2.2  The Extension of the ACS Study
 1           A very recent publication by Pope et al. (2002) extends the analyses (Pope et al., 1995) and
 8      reanalyses (Krewski et al., 2000) of the ACS CPS-II cohort to an additional eight years of foilow-
 9      up.  The new study has a number of advantages over the previous analyses, in that it: (a) doubles
10      the follow-up time from eight years to sixteen years, and triples the number of deaths;
11      (b) expands the ambient air pollution data substantially, including two recent years of fine
12      particle data, and adds data on gaseous co-pollutants; (c) improves statistical adjustments for
13      occupational exposure; (d) incorporates data on dietary covariates believed to be important
14      factors in mortality, including total fat consumption, and consumption of vegetables, citrus fruit,
15      and high-fiber grains; and (e) uses recent developments in non-parametric spatial smoothing and
16      random effects statistical models as input to the Cox proportional hazards model.  Each
17      participant was identified with a specific metropolitan area, and mean pollutant concentrations
18      were calculated for all metropolitan areas with ambient air monitors in the one to two years prior
19      to enrollment. Ambient pollution during the follow-up period was extracted from the AIRS data
20      base.  Averages of daily averages of the gaseous pollutants were used except for ozone, where
21      the average daily 1-hour maximum was calculated for the whole year and for the typical peak
22      ozone quarter (July, August, September). Mean sulfate concentrations for 1990 were calculated
23      from archived filters using quartz filters, virtually eliminating the historical sulfate artifact
24      leading to overestimation of sulfate concentrations.
25           The Krewski et al. (2000), Burnett et al. (200la), and Pope et al. (2002) studies were
26      concerned that survival times of participants in nearby locations might not be independent of
27      each other, due to missing, unmeasured or mis-measured risk factors or their surrogates that may
28      be spatially correlated with air pollution, thus violating an important assumption of the Cox
29      proportional hazards model. Model fitting proceeded in two stages, the first of which was an
30      adjusted relative risk model with a standard Cox proportional hazards model including
31      individual-specific covariates and indicator variables for each metropolitan area, but not air

        April 2002                                 8-78        DRAFT-DO NOT QUOTE OR CITE

-------
 1      pollutants.  In the second stage, the adjusted log(relative risks) were fitted to fine particle
 2      concentrations or other air pollutants by a random effects linear regression model.
 3           Models were estimated separately for each of four mortality (total, cardiopulmonary, lung
 4      cancer, and causes other than cardiopulmonary or lung cancer deaths) endpoints for the entire
 5      follow-up period, and for fine particles in three time periods (1979-1983, 1999-2000, and the
 6      average of the mean concentrations in these two periods). The results are shown in Table 8-7.
 7      Figures 8-8, 8-9, and 8-10 show the results displayed in Figures 2, 3, and 5 in Pope et al. (2002).
 8      Figure 8-8  shows that a smooth non-parametric model can be reasonably approximated by a
 9      linear model for all-cause mortality, cardiopulmonary mortality, and other mortality; but the
10      log(relative risk) model for lung cancer appears to be non-linear, with a steep linear slope up to
11      an annual mean concentration of about 13 //g/m3 and a flatter linear slope at fine particle
12      concentrations > 13 //g/m3.
13
14
            TABLE 8-7. SUMMARY OF RESULTS FROM THE EXTENDED ACS STUDY*
                                 PM2 5, average over      PM2 5, average over     PM2 5, average over all
         Cause of death               1979-1983               1999-2000             seven years
         All causes                4.1%  (0.8,7.5%)        5.9% (2.0,9.9%)        6.2% (1.6, 11.0%)
         Cardiopulmonary         5.9%  (1.5, 10.5%)       7.9% (2.3, 14.0%)       9.3% (3.3, 15.8%)
         Lung cancer             8.2%  (1.1, 15.8%)       12.7% (4.1,21.9%)      13.5%  (4.4,23.4%)
         Other	0.8%  (-3.0, 4.8%)       0.9% (-3.4, 5.5%)       0.5% (-4.8, 6.1%)
         'Adjusted mortality excess risk ratios (95% confidence limits) per 10 ,wg/m3 PM2 5 by cause of death associated
         with each of the multi-year averages of fine particle concentrations. The multi-year average concentrations are
         used as predictors of cause-specific mortality for all of the 16 years (1982-1998) of the ACS follow-up study.
         The excess risk ratios are obtained from the baseline random effects Cox proportional hazards models adjusted
         for age, gender, race, smoking, education, marital status, BMI, alcohol consumption, occupational dust
         exposure, and diet. Based on Table  2 in Pope et al. (2002) and more precise data from authors (G. Thurston,
         personal communication, March 13. 2002).
 1           Figure 4 in Pope et al. (2000) shows results of the stratified first-stage models:  ages
 2      < 60 and > 69 yr are marginally significant for total mortality; ages > 70 are significant for
 3      cardiopulmonary mortality; and ages 60-69 for lung cancer mortality.  Men are at significantly

        April 2002                                  8-79        DRAFT-DO NOT QUOTE OR CITE

-------
        o
 0.2 -
 0.1 -


-0.1 -


-0.3 -
-0.4 -
                               All Cause
0.2 -
^ 0.1 -
en
& nn
0, °-0 '
in -0.1 -
Q)
I**'
0
-1 -0.3 -
-0.4 -
^t** '
.. . -- ' ** ^-^--*"""""'"""\ — •
	 •^ .-••*•
y1"
x1"
(X*
Cardiopulmonary

                 10
                            15
                                        20
                                                               10
                                                                          15
                                                                                      20
          0.2 :
       ^ 0.1 -
       "
-------
                  Standard Cox Models
     1.10-
     1.15-   £
             3 «5   1
62
~ to
 Random
  Effects
 (with all
covariates)
                                      Random Effects (with
                                      all covariates) plus
                                      Spatial Smoothing
                                       s£
                                       o
                                                                    CD
 <> <

1 -L 1


1 ?
I i


f ]




> 4




5
5
.
<
>


[ ^




^
^




c
c
- £/
>
4


.
> 4




^ <




; 05 TO ^ ^
i §1 S?
3 -I > I 0.


> ^ >


_ _

, ^ 0

. .


— 1.05-
1.00 -
n 95 _

- 	 M- 	 f-f 	 f 	 f 	 f 	 :

>
4

 Figure 8-9.  Relative risk of total and cause-specific mortality at 10 Aig/m3 PM2 5 (mean
             of 1979-1983) of alternative statistical models. The standard Cox models
             are built up in a sequential stepwise manner from the baseline model
             stratified by age, gender, and race by adding additional covariates. The
             random effects model allows for additional city-to-city variation, and the
             spatial smoothing models show the effects of increasingly aggressive
             adjustment for spatial correlation.  Based on Pope et al. (2002).
April 2002
                8-81
  DRAFT-DO NOT QUOTE OR CITE

-------
            Fine       Inhalable, coarse, and    Sulfate
            Particles   total suspended Particlesparticles
                   Gaseous Pollutants
            CO
                                  T-  CO   1.15-
3
ro 1.10-
O
-1 1.05-
^ H f\r\ _
I .UU "
0.95 -
0.90 -
£• 1.25-
C 1.20-
o
E 1.15-
3 1.10-
Q.
.° 1.05-
H -i nn _
i; 1 .UU
O 0.95 -
0.90 -
1.25 -
fe 120-
£ 1.15-
o 1.10-
g> 1.05-
D -i nn -
• 1 .UU
0.95 -
0.90 -
1.25 -
1.20 -
0) 1.15-
+= 1.10-
o
= 1.05-
**• 1.00-
0.95 -
0.90 -
0
c
r-
x».
^
^
Q



^
4







4






4










4

D C
1 C
^ 2
»" ^
. Q




^ <








4





> 4










^.....^

» -g
•> 6
^' -^.
r s
. c


J
4
>






4

^




4
>










^...^

J C
1 s
E" 2
D



^


^ 4




»

4




>




4






>
4

} OC
3 I--
-^ ---
E^ S
L Q




^

k





4
K







4

>






'.....^
>

3 Q, 00 OC
L "s S2- C
£ i Q. Q
: S w c/
. Q. 1- 1-




^

^ T ^
* 1 ]




^

j
^
-L -1 .




>

j
j
4




^

3 CO T-
co o c
^ ~^ s°, 5
3 CO O* C
- 1- CO 0



4

T i
> 1
[ I






T
0 I
. 1



-p
4. <
1

k
r • •



4
^ 	 - 	

r i
) C
3 U

r

>
n
L <







k .
.
<





>•

4





>
L 	 ]

00
? ^>
? CNI C
3 00 OC
)" O" C
3 CO 2





^ ^L
]
^














..


4




ii1

cc
c
r o
) CC
) C
: z






[ J
M







4









»
4




H

i
i I
r c
: c





r









' J
^ 1








4
>




».--:

oo ^
en T-
J- CN «
) 00 c
) O ^
) O C







^ 4







f ^

1







fr




4i

-N ^-
3 §
3 °

)" C






r 4,
[]







4









>
4




•<
>

u
3~ OC
^ c
g o
3 OC
.X S.
f c





4
>







4
>









4
>




.
>

j
3
5
J
3
*S
)"





^








^










>




>- 	

 Figure 8-10.  Relative risk of total and cause-specific mortality for particle metrics and
              gaseous pollutants over different averaging periods (years 1979-2000 in
              parentheses). Based on Pope et al. (2002).
April 2002
8-82
DRAFT-DO NOT QUOTE OR CITE

-------
 1      smoothing can increase the magnitude of the RR and increase its significance by reducing the
 2      width of the confidence intervals in the "50%-span" and "lowest variance" smoothing methods.
 3      For lung cancer mortality, spatial smoothing very slightly decreases the magnitude of the RR but
 4      also increases its significance by reducing the width of the confidence intervals in the "50%-
 5      span" and "lowest variance" smoothing methods.
 6           Figure 8-10 shows a statistically significant relationship between fine particles and total,
 7      cardiopulmonary, and lung cancer mortality whatever averaging span was used for PM2 5, and
 8      slightly larger for the average concentration of the 1979-1983 and 1999-2000 intervals. PM15 for
 9      1979-1983 is significantly associated with cardiopulmonary mortality and marginally with total
10      mortality, whereas 1987-1996 PM15 is not quite significantly associated with cardiopulmonary
11      mortality only. Coarse particles and TSP are not significantly associated with any endpoint, but
12      are positively associated with cardiopulmonary mortality.  Sulfate particles are very significantly
13      associated with all endpoints including mortality from all other causes, but only marginally for
14      lung cancer mortality using 1990 filters.
15           Figure 8-10 shows a highly significant relationship between SO2 and all endpoints
16      including mortality from other causes, although weaker for lung cancer mortality. Ozone (using
17      only the third quarter for 1982-1998) shows a marginally significant relationship with
18      cardiopulmonary mortality, but not the year-round average. The other criteria pollutants, CO and
19      NO2, are not significantly and positively related to any mortality endpoint, unlike the findings for
20      acute mortality studies.
21           This paper is noteworthy because it shows that the general pattern of findings in the first
22      eight years of the study (Pope et al., 1995; Krewski et al., 2000) could be reasonably extrapolated
23      to the patterns that remain present with twice the length of time on study and three times the
24      number of deaths.  As shown later in  Table 8-12 (Pg.  8-94), the excess relative risk estimate
25      (95% CI) per 10 //g/m3 PM25 for total mortality in the original ACS study (Pope et al., 1995) was
26      6.6% (3.6, 9.9%); in the ACS reanalysis (Krewski et al., 2000, Table 20, Full Model) it was 6.6%
27      (3.6, 9.9%); and, in the extended ACS data set (Pope et al., 2002), it was 4.1% (0.8, 7.5%) using
28      the 1979-1983 data and 6.2% (1.6,  11%) using the average  of the 1979-1983 and 1999-2000 data.
29      The excess relative risk estimate (95% CI) per 10 //g/m3 PM25 for cardiopulmonary mortality in
30      the original ACS study (Pope et al., 1995) was 11.6% (6.6,  16.7%); in the ACS reanalysis
31      (Krewski et al., 2000, Table 20, Full Model), it was 10.6%  (5.9, 15.4%); and, in the extended

        April 2002                                8-83        DRAFT-DO NOT QUOTE OR CITE

-------
 1      ACS data set (Pope et al., 2002), it was 5.9% (1.5, 10.5%) using the 1979-1983 data and 9.3%
 2      (3.3, 15.8%) using the average of the 1979-1983 and 1999-2000 data. Thus, the additional data
 3      and statistical analyses in (Pope et al., 2000) yield somewhat smaller estimates than in the
 4      original study (Pope et al., 1995), but similar estimates to the reanalysis of the original ACS data
 5      set (Krewski et al., 2000).
 6           The authors draw the following conclusions:
 7      (1)  The apparent association between fine particle pollution and mortality persists with longer
 8          follow-up as the participants in the cohort grow older and more of them die.
 9      (2)  The estimated fine particle effect on cardiopulmonary mortality and cancer mortality was
10          relatively stable, even after adjustment for smoking status, although the estimated effect was
11          larger and more significant for never-smokers vs. former or current smokers.  The estimates
12          were relatively robust against inclusion  of many additional covariates: education, marital
13          status, BMI, alcohol consumption, occupational exposure, and dietary factors. However, as
14          the authors note, the data on individual risk factors was collected only at the time of
15          enrollment and has not been updated, so that changes in these factors since 1982 could
16          introduce risk factor exposure mis-classification, with a loss of precision in the estimates
17          and might limit the ability to characterize time dependency of effects.
18      (3)  Additional  assessments for potential spatial or regional differences not controlled in the
19          first-stage model were evaluated. If there  are unmeasured or inadequately modeled risk
20          factors that are different across locations or spatially clustered, then PM risk estimates may
21          be biased. If the clustering is independent or random or independent across areas, then
22          adding a random-effects component to the Cox proportional hazards model can deal with the
23          problem. However, if location is associated with air pollution, then the spatial correlation
24          may be evaluated using non-parametric  smoothing methods. No significant spatial auto-
25          correlation  was found after controlling for fine particles. Even after adjusting for spatial
26          correlation, the estimated PM2 5 effects were significant and persisted for cardiopulmonary
27          mortality and lung cancer mortality and  were borderline significant for total mortality, but
28          with much wider confidence intervals after spatial smoothing.
29      (4)  Elevated total, cardiopulmonary, and lung cancer mortality risks were associated with fine
30          particles, but other mortality was not. PM10 for 1987-1996 and PM15 for 1979-1983 were
31          just significantly associated with cardiopulmonary mortality only, but PM10_2 5 and TSP were

        April 2002                                 8-84        DRAFT-DO NOT QUOTE OR CITE

-------
 1          not associated with total or any cause-specific mortality.  All endpoints were very
 2          significantly associated with sulfates, except lung cancer with 1990 sulfate data. All
 3          endpoints were very significantly associated with SO2 using 1980 data, with total and other
 4          mortality using the 1982-1998 data, but cardiopulmonary and lung cancer mortality had only
 5          a borderline significant association with the 1982-1998 SO2 data. None of the other gaseous
 6          pollutants had a significant positive association with any endpoint, except for a borderline
 7          association of third-quarter ozone to cardiopulmonary mortality. In summary, neither coarse
 8          thoracic particles nor TSP were significantly associated with mortality, nor were NO2 and
 9          CO on a long-term exposure basis.  (It should be noted, however, that additional analyses
10          may yet be useful.  The data would allow segmentation of mortality into smaller periods
11          rather than the whole 16 year duration of the mortality follow up, for example from 1982
12          through 1989 and from 1990 through 1998. In this way,  it may be possible to evaluate  any
13          changes in PM mortality rate over time.)
14      (5)  The concentration-response curves estimated using non-parametric smoothers were all
15          monotonic and (except for lung cancer) nearly linear. However, the shape of the curve may
16          become non-linear at much higher concentrations.
17      (6)  The excess risk from PM2 5 exposure is much smaller than that estimated for cigarette
18          smoking for current smokers in the same cohort (Pope et al., 1995), RR = 2.07 for total
19          mortality, RR = 2.28 for cardiopulmonary mortality, and RR = 9.73 for lung cancer
20          mortality. In the more polluted areas of the United States, the relative risk for substantial
21          obesity (a known risk factor for cardiopulmonary mortality) is larger than that for PM2 5, but
22          the relative risk from being moderately overweight is somewhat smaller.
23
24      8.2.3.2.3  AHSMOGAnalyses
25          The Adventist Health Study of Smog (AHSMOG) represents a third major U.S. prospective
26      cohort study of chronic PM exposure-mortality effects. In 1977, the study enrolled 6,338
27      non-smoking non-Hispanic white Seventh Day Adventist residents of California, ages 27 to
28      95 years.  The participants had resided for at least 10 years within 5 miles (8 km) of their then-
29      current residence locations, either within the three major California air basins (San Diego,
30      Los Angeles, or San Francisco) or else were part of a random 10% sample of Adventist Health
31      Study participants residing elsewhere in  California.  The study has been extensively described

        April 2002                                 8-85         DRAFT-DO NOT QUOTE OR CITE

-------
 1      and initial results reported elsewhere (Hodgkin et al., 1984; Abbey et al., 1991; Mills et al.,
 2      1991).  In the latest AHSMOG analyses (Abbey et al., 1999), mortality status of the subjects after
 3      ca. 15-years of follow-up (1977-1992) was determined by various tracing methods, finding 1,628
 4      deaths (989 female, 639 male) in the cohort.  This is a 50% percent increase in the follow-up
 5      period vs. previous AHSMOG reports, which increases the power of the latest analyses over past
 6      published ones.  Of 1,575 deaths from all natural (non-external) causes, 1,029 were
 7      cardiopulmonary, 135 were non-malignant respiratory (ICD9 codes 460-529), and 30 were lung
 8      cancer (ICD9 code  162) deaths. Abbey et al. (1999) also created another death category,
 9      contributing respiratory causes (CRC).  CRC included any mention of nonmalignant respiratory
10      disease as either an underlying or a "contributing cause" on the death certificate. Numerous
11      analyses were done for the CRC category, due to the large numbers and relative specificity of
12      respiratory causes as a factor in  the deaths. Education was used to index socio-economic status,
13      rather than income. Physical activity and occupational exposure to dust were also used as
14      covariates.
15          Cox proportional hazard models adjusted for a variety of covariates, or stratified by sex,
16      were used. The "time" variable used in most of the models was survival time from date of
17      enrollment, except that age on study was used for lung cancer effects due to the expected lack of
18      short-term effects.  A large number of covariate adjustments were evaluated, yielding results for
19      all non-external mortality as shown in Table 8-8 and described by Abbey et al. (1999).
20      Essentially no statistically significant PM related effects were observed for either  males or
21      females, except RR = 1.08 for males in relation to 30 days per year with PM10 > 100 //g/m3.
22          An analogous pattern of results was found for cause-specific mortality analyses of the
23      AHSMOG data. That is, positive and statistically significant effects on cardiopulmonary deaths
24      were found in models that included both sexes and adjustment for age, pack-years of smoking,
25      and body-mass index (BMI) (RR = 1.14, 95% CI1.03-1.56 for 30 day/yr > 100 //g/m3 PM10).
26      Subsets of the cohort had elevated risks: (a) former smokers had higher RR's than never-
27      smokers (RR for PM10 exceedances for never-smokers was marginally significant by itself);
28      (b) subjects with low intake  of anti-oxidant vitamins A, C, E had significantly elevated risk of
29      response to PM10, whereas those with adequate intake did not (suggesting that dietary factors or,
30      possibly, other socio-economic or life style factors for which they are a surrogate may be
31      important covariates); and (c) there also appeared to be a gradient of PM10 risk with respect to

        April 2002                                 8-86        DRAFT-DO NOT QUOTE OR CITE

-------
           TABLE 8-8. RELATIVE RISK OF MORTALITY FROM ALL NONEXTERNAL
                CAUSES, BY SEX AND AIR POLLUTANT, FOR AN ALTERNATIVE
                          COVARIATE MODEL IN THE ASHMOG STUDY

Pollution Index
PM10>100, d/yr.
PM10 mean
SO4 mean
O3>100 ppb, h/yr.
SO2 mean

Pollution Incr.
30 days/yr.
20 Mg/m3
S^g/m3
551 h/yr. (IQR)
3.72 (IQR)

RR
0.958
0.950
0.901
0.90
1.00
Females
LCL
0.899
0.873
0.785
0.80
0.91

UCL
1.021
1.033
1.034
1.02
1.10

RR
1.082
1.091
1.086
1.140
1.05
Males
LCL
1.008
0.985
0.918
0.98
0.94

UCL
1.162
1.212
2.284
1.32
1.18
        LCL = Lower 95% confidence limit
        Source: Abbey etal. (1999).
UCL = Upper 95% confidence limit
 1     time spent outdoors, with those who had spent at least 16 h/wk outside at greater risk from PM10
 2     exceedances. The extent to which time spent outdoors is a surrogate for other variables or is a
 3     modifying factor reflecting temporal variation in exposure to ambient air pollution is not certain.
 4     For example, if the males spent much more time outdoors than females, outdoor exposure time
 5     could be confounded with gender. When the cardiopulmonary analyses are broken down by
 6     gender (Table 8-9), the RR's for female deaths were generally smaller than that for males,
 7     although none of the risks for PM indices or gaseous pollutants were statistically significant.
 8         The AHSMOG cancer analyses showed a confusing array of results for lung cancer mortality
 9     (Table 8-10). For example, RR's  for lung cancer deaths were statistically significant for males
10     for PM10 and O3 metrics, but not for females. In contrast, such cancer deaths were significant for
11     mean NO2 only for females (but not for males), but lung cancer metrics for mean SO2 were
12     significant for both males and females.  This pattern is not readily interpretable, but is reasonably
13     attributable to the very small numbers of cancer-related deaths (18 for females and 12 for males),
14     resulting in wide RR confidence intervals and very imprecise effects estimates.
15         The analyses reported by Abbey et al. (1999) attempted to separate PM10 effects from those
16     of other pollutants by use of two-pollutant models, but no quantitative findings from such models
17     were reported. Abbey et al. mentioned that the PM10 coefficient for CRC remained stable or
       April 2002
8-87
DRAFT-DO NOT QUOTE OR CITE

-------
   TABLE 8-9. RELATIVE RISK OF MORTALITY FROM CARDIOPULMONARY
        CAUSES, BY SEX AND AIR POLLUTANT, FOR AN ALTERNATIVE
                             COVARIATE MODEL

Pollution Index
PM10>100, d/yr.
PM10 mean
SO4 mean
O3>100 ppb, h/yr.
O3 mean
SO2 mean

Pollution Incr.
30 days/yr.
20 Mg/m3
5 A(g/m3
551 h/yr. (IQR)
10 ppb
3.72 (IQR)

RR
0.929
0.933
0.950
0.88
0.975
1.02
Females
LCL
0.857
0.836
0.793
0.76
0.865
0.90

UCL
1.007
1.042
1.138
1.02
1.099
1.15

RR
1.062
1.082
1.006
1.06
1.066
1.01
Males
LCL
0.971
0.943
0.926
0.87
0.920
0.86

UCL
1.162
1.212
1.086
1.29
1.236
1.18
 LCL = Lower 95% confidence limit

 Source: Abbey etal. (1999).
UCL = Upper 95% confidence limit
  TABLE 8-10. RELATIVE RISK OF MORTALITY FROM LUNG CANCER BY AIR
  POLLUTANT AND BY GENDER FOR AN ALTERNATIVE COVARIATE MODEL

Pollution
Index
PM10>100,d/yr.
PM10 mean
NO2 mean
O3>100 ppb,
h/yr



O3 mean
SO2 mean



Pollution
Incr.
30 days/yr.
20 ,wg/m3
19.78 (IQR)
551 h/yr
(IQR)



10 ppb
3.72 (IQR)



Smoking
Category
All1
All
All
All
never
smoker
past
smoker
All
All
never
smokers

RR
1.055
1.267
2.81
1.39



0.805
3.01
2.99

Females
LCL
0.657
0.652
1.15
0.53



0.436
1.88
1.66


UCL
1.695
2.463
6.89
3.67



1.486
4.84
5.40


RR
1.831
2.736
1.82
4.19
6.94

4.25
1.853
1.99


Males
LCL
1.281
1.455
0.93
1.81
1.12

1.50
0.994
1.24



UCL
2.617
5.147
3.57
9.69
43.08

12.07
3.453
3.20


 'All = both never smokers and past smokers.
 LCL = Lower 95% confidence limit

 Source: Abbey etal. (1999).
UCL = Upper 95% confidence limit
April 2002
           DRAFT-DO NOT QUOTE OR CITE

-------
 1      increased when other pollutants were added to the model. Lung cancer mortality models for
 2      males evaluated co-pollutant effects in detail and indicated that NO2 was non-significant in all
 3      two-pollutant models but the other pollutant coefficients were stable.  The PM10 and O3 effects
 4      remained stable when SO2 was added, suggesting possible independent effects, but PM10 and O3
 5      effects were hard to separate because these pollutants were highly correlated in this study.
 6      Again, however, the very small number of lung cancer observations and  likely great imprecision
 7      of reported effects estimates markedly diminish the credibility of these results.
 8          Other analyses, by Beeson et al. (1998), evaluated essentially the same data as in Abbey
 9      et al. (1999), but focused on lung cancer incidence (1977-1992).  There were only 20 female and
10      16 male lung cancer cases among the 6,338 subjects. Exposure metrics were constructed to be
11      specifically relevant to cancer, being the annual average of monthly exposure indices from
12      January, 1973 through the following months, but ending 3 years before date of diagnosis of the
13      case (i.e., representing a 3-year lag between exposure and diagnosis of lung cancer). The
14      covariates in the Cox proportional hazards model were pack-years of smoking and education, and
15      the time variable was attained age.  Many additional covariates were evaluated for inclusion, but
16      only 'current use of alcohol' met criteria for inclusion in the final model.  Pollutants evaluated
17      were PM10, SO2, NO2, and O3. No interaction terms with the pollutants proved to be significant,
18      including outdoor exposure times. The RR estimates for male lung cancer cases were:
19      (a) positive and statistically significant for all PM10 indicators; (b) positive and predominantly
20      significant for O3 indicators, except for  mean O3, number of O3 exceedances > 60 ppb, and in
21      former smokers; (c) positive and significant for mean SO2, except when restricted to proximate
22      monitors; and (d) positive but not significant for mean NO2.  When analyses are restricted to use
23      of air quality data within 32 km of the residences of subj ects, the RR over the IQR of 24 //g/m3 in
24      the full data set is 5.21 (or RR=1.989 for 10 //g/m3). The female RR's were all much smaller
25      than for males, not being statistically significant for any indicator of PM10 or O3, but being
26      significant for mean SO2.
27          The AHSMOG investigators also attempted to compare  effects of fine vs. coarse particles
28      (McDonnell et al, 2000). For AHSMOG participants living near an airport (n=3,769), daily
29      PM2 5 concentrations were estimated from airport visibility using previously-described methods
30      (Abbey et al, 1995b).  Table 8-11 shows the results of this analysis for the male subset near
31      airports (n=1266). Given the smaller numbers of subjects in  these subset analyses, it is not

        April 2002                                8-89       DRAFT-DO NOT QUOTE OR CITE

-------
             TABLE 8-11. COMPARISON OF EXCESS RELATIVE RISKS FOR THREE
             PARTICLE METRICS IN THE MALE SUBSET OF THE AHSMOG STUDY
PM Metric
Underlying Cause of Mortality
All causes


Any contributing nonmalignant
respiratory cause


Lung cancer


PM25
8.4.% (-2.1,
—
9.3% (-3. 8,
22.6% (-2.9,
—
19.8% (-8.8,
39.1% (-21,
—
35.7% (-28,

21%)

24%)
55%)

58%)
146%)

157%)
PM10.2.5
—
5.2% (-8.2, 21%)
-1.0% (-16, 17%)
—
19.6% (-12, 64%)
6.2% (-27, 54%)
—
25. 9% (-3 8, 156%)
7.2% (-52, 137%)
PM10
9.9% (-4. 1,26%)
30.5% (-4.8, 140%)
5 1.2% (-30, 224%)
 1     necessarily surprising that no pollutants are statistically significant in these regressions.  It is
 2     important, however, to caveat that the PM2 5 exposures were estimated from visibility
 3     measurements (increasing exposure measurement error), and a very uneven and clustered
 4     distribution of exposures was presented by the authors. Also, the PM10_2 5 values were calculated
 5     from the differencing of PM10 and PM25, likely contributing to additional measurement error for
 6     the coarse particle (PM10_2 5) variable used in the analyses.
 7
 8     8.2.3.2.4  The EPRI- Washington University Veterans' Cohort Mortality Study
 9         Lipfert et al. (2000b) reported preliminary results from new large-scale mortality analyses
10     using a prospective cohort of up to 70,000 men assembled by the U.S. Veterans Administration
11     (VA) in the mid 1970s. While much smaller than the ACS cohort, this study group shares the
12     similarity that it was not originally formed to study air pollution, but was later linked to air
13     pollution data collected separately, much of it subsequent to the start of the study. The AHSMOG
14     and Six City studies were designed as prospective studies to evaluate long-term effects of air
15     pollution and had concurrent air pollution measurements.  The ACS study was also a prospective
16     study, with air pollution data at about the approximate time of enrollment but not subsequently
       April 2002
8-90
DRAFT-DO NOT QUOTE OR CITE

-------
 1      (Pope et al., 1995). The extended ACS data incorporated much more air pollution data,
 2      including TSP data back to the 1960s and more recent fine particle data.  The PM25 data set was
 3      smaller than the TSP data set and similar to the ACS data.
 4          The study cohort was male, middle-aged (51 ± 12 years) and included a larger proportion of
 5      African-Americans (35%) than the U.S. population as a whole and a large percentage of current
 6      or former smokers (81%). The cohort was selected at the time of recruitment as being mildly to
 7      moderately hypertensive, with screening diastolic blood pressure (DBF) in the range 90 to
 8      114 mm Hg (mean 96, about 7 mm more than the U.S. population average) and average  systolic
 9      blood pressure (SBP) of 148 mm Hg.  The subjects had all been healthy enough to be in  the U.S.
10      armed forces at one time. A comparison of their pre-existing health  status at time of study
11      recruitment vs. the initial health status of the other cohorts would be of interest. The study that
12      led to the development of this clinical  cohort (Veterans Administration Cooperative Study Group
13      on Antihypertensive Agents, 1970; 1967) was a "landmark" VA cooperative study demonstrating
14      that anti-hypertensive treatment markedly decreased morbidity and mortality (Perry et al., 1982).
15      The clinical cohort itself involved actual clinical rather than research settings.  Some differences
16      between the VA cohort and other prospective cohorts are noted below.
17          Pollutant levels of the county of residence at the time of entry into the study were used for
18      analyses versus levels at the VA hospital area. Contextual socioeconomic variables were also
19      assembled at the ZIP-code and county levels. The ZIP-code level  variables were average
20      education, income, and racial mix.  County-level variables included altitude, average annual
21      heating-degree days, percentage Hispanic, and socioeconomic indices. Census tract variables
22      included poverty rate and racial mix. County-wide air pollution variables included TSP, PM10,
23      PM25, PM15, PM15_25, SO4, O3, CO, and NO2 levels at each of the 32 VA clinics where veterans
24      were enrolled. In addition to considering average exposures over the entire period, three
25      sequential mortality follow-up periods (1976-81, 1982-88, 1989-96) were also considered
26      separately in statistical analyses, which evaluated relationships of  mortality in  each of those
27      periods to air pollution in different preceding, concurrent, or subsequent periods (i.e., up to 1975,
28      1975-81, 1982-88, and 1989-86, for TSP in the first three periods, PM10 for the last, and  NO2,
29      95 percentile O3, and 95 percentile CO for all four periods). Mortality in the above-noted periods
30      was also evaluated in relation to SO4 in each of the same four periods noted for NO2, O3, and CO,
31      and to PM25, PM15, and PM15.25 in 1979-81 and 1982-84.

        April 2002                               8-91       DRAFT-DO NOT QUOTE OR CITE

-------
 1          The use of diastolic and systolic blood pressure in the reported regression results may
 2      require further evaluation. The VA Cohort participants were recruited on the basis of initial
 3      diastolic blood pressure (DBF) of 90 to 114 mm Hg.
 4          The participants in the VA Cohort clearly formed an "at-risk" population, and the results by
 5      Vasan et al. (2001) make more plausible the hypothesis in (Lipfert et al., 2000b, p. 62) that
 6      ". . .the relatively high fraction of mortality within this cohort may have depleted it of susceptible
 7      individuals in the later periods of follow-up." The role of DBF and SBP as predictors in
 8      regression models in the VA Cohort may be considered as closer to the endpoint (mortality) than
 9      as a more distal behavioral, environmental, or contextual predictor of mortality such as air
10      pollution, temperature, smoking behavior, BMI, etc.  The author (F. Lipfert, personal
11      communication, March 28, 2002) notes that personal-level variables tend to interact only with
12      each other, as do county-level variables with little correlation across spatial scales.
13          The estimated mean risk of cigarette smoking in this cohort (1.43 Relative Risk) is also
14      smaller than that of the Six City cohort (RR = 1.59) and the ACS cohort (RR = 2.07 for a current
15      smoker).  Some possible differences include the higher proportion of former or current smokers
16      in this cohort (81%) vs. 51% in the ACS study and 42 to 53% in the Six City study.  A possibly
17      more important factor may be the difference in education levels, with only 12% of the ACS
18      participants having less than a  high school education vs 28% of the Six City cohort and not
19      reported for the VA Cohort (although the Armed Services do have enlistment standards). The
20      education differences may be associated with smoking behavior. Also, the large number of
21      interaction terms in the model  may account for part of the difference.
22          The preliminary screening models used proportional hazards regression models (Miller
23      et al., 1994) to identify age, SBP, DBF, body mass index (BMI, nonlinear), age and race
24      interaction terms, and present or  former smoking as baseline predictors, with one or two
25      pollution variables added. In the final model using 233 terms (of which 162 were interactions of
26      categorized SBP, DBF, and BMI variables with age), the most significant non-pollution variables
27      were SBP, DBF, BMI, and their interactions with age, smoking status, average ZIP education,
28      race, poverty, height, and a clinic-specific effect. Lipfert et al. (2000b) noted that the risk of
29      current cigarette smoking (1.43) that they found was lower than reported in other studies. The
30      most consistently positive effects were found for O3 and NO2 exposures in the immediately
31      preceding years. This study used peak O3 rather than mean O3 as in some other cohort studies.

        April 2002                               8-92         DRAFT-DO NOT QUOTE OR CITE

-------
 1      This may account for the higher O3 and NO2 effects here.  While the PM analyses considering
 2      segmented (shorter) time periods gave differing results (including significantly negative mortality
 3      coefficients for some PM metrics), when methods consistent with the past studies were used (i.e.,
 4      many year average PM concentrations), similar results were reported, with the authors finding
 5      that "(t)ne single-mortality-period responses without ecological variables are qualitatively similar
 6      to what has been reported before (SO4= > PM25 > PM15)".  With ecological variables included,
 7      the only significant PM effect was that of TSP up to 1981 on 1976-81  mortality.  It might be
 8      instructive to evaluate more parsimonious regression models with fewer ecological covariates
 9      and interaction terms. It is noteworthy that estimated PM effects appear to be smaller in the later
10      years of the study rather than in the earlier years.  This may also be due to cohort depletion.
11
12      8.2.3.2.5 Relationship ofAHSMOG, Six Cities, ACS and VA Study Findings
13          The results of the more recent AHSMOG mortality analyses (Abbey et al., 1999; McDonnell
14      et al., 2000) are compared here with findings from the earlier Six Cities study (Dockery et al.,
15      1993), the ACS study (Pope et al., 1995), the HEI reanalyses of the latter two studies, the
16      extension of the ACS study (Pope et al., 2002), and the VA study (Lipfert et al., 2000).
17      Table 8-12 compares the estimated RR for total, cardiopulmonary, and cancer mortality among
18      the studies. The number of subjects in these studies varies greatly (8,111 subjects in the
19      Six-Cities Study; 295,223 subjects in the 50 fine  particle (PM25) cities and 552,138 subjects in
20      the 151 sulfate cities of the ACS Study; 6,338 in  the AHSMOG Study; 70,000 in the VA study);
21      and this may partially account for differences among their results.
22          As shown in Table 8-12, the Six Cities study found significant associations with all PM
23      indicators.  In the Krewski  et al. (2000) reanalysis of the ACS study data, larger associations
24      were found for both PM2 5 and PM15 (excess relative risks of 6.6% for 10 //g/m3 PM2 5 and 4% for
25      20 //g/m3 increments  in annual PM15, respectively), although both associations were significant.
26      Most recently, McDonnell et al. (2000) reported evidence from the AHSMOG analyses
27      suggestive of somewhat stronger associations with fine particles than coarse particles,  though the
28      associations were only reported for males and none reached statistical  significance.
29          Overall, the results most recently reported for the AHSMOG study (Abbey et al.,  1999;
30      McDonnell et al., 2000) do not find consistent, statistically significant associations between
31      mortality and long-term PM exposure, though the authors conclude that some evidence was

        April 2002                                8-93        DRAFT-DO NOT QUOTE OR CITE

-------
    TABLE 8-12. COMPARISON OF EXCESS RELATIVE RISKS OF LONG-TERM
  MORTALITY IN THE HARVARD SIX CITIES, ACS, AHSMOG, AND VA STUDIES
Study
Six City3
Six City
New4
ACS5
ACS6
New
ACS
New
ACS
New
ACS
New
ACS
Extend.7
ACS
Extend.
ACS
Extend.
AHSMOG8
AHSMOG9


AHSMOG10
VA12
PM1
PM25
PM25

PM25
PM25

PM15.,5

PM10/15
Dichot
PM10/15 SSI

PM25
1979-83
PM25
1999-000
PM2 5 Avg.

PM10/15
30+ days
PM10/15
>100
PM25
PM25
Total
Ex. RR2
13%
14%

6.6%
7.0%

0.3%

4%

2%

4.1%

5.9%

6.2%

2%
NA


9.3%u
-10.0%
Mortality
95% CI
(4.2, 23%)
(5.3, 23%)

(3.6, 9.9%)
(4.0, 10%)

(-0.9, 1.8%)

(1.0, 9%)

(-1.0,4%)

(0.8, 7.5%)

(2.0, 9.9%)

(1.6, 11%)

(-5, 9%)
NA


(-3.8,24%)
(-15,-4.6%)
Cardiopulmonary
Mortality
Ex.
RR
18%
19%

11.6%
12.0%

0.3%

7%

6%

5.9%

7.9%

9.3%

1%
14%


20%9

95% CI
(5.8, 32%)
(6.3, 33%)

(6.6, 17%)
(7.4, 17%)

(-1.5%, 2.4%)

(3, 12%)

(3, 9%)

(1.5, 10%

(2.3, 14%)

(3.3, 16%)

(-8, 10%)
(3, 26%)


(-9, 55%)

Lung Cancer Mortality
Ex. RR
18%
21%

1.2%
0.8%

-0.9%

0.4%

-0.8%

8.2%

12.7%

13.5%

174%9
NA


36%

95% CI
(-11,57%)
(-8.4,60%)

(-8,7, 12%)
(-8.7, 11%)

(-5.5%, 3.8%)

(-4.0, 5%)

(-4.4, 3%)

(1.1, 16%)

(4.1,22%)

(4.4, 23%)

(45, 415%)
NA


(-28, 157%)

 'Increments are 10 ^g/m3 for PM2 5 and 20 ^g/m3 for PM10/15.
 2Ex.RR (excess relative risk, percent) = 100 * (RR -1) where the RR has been converted from the
 highest-to-lowest range to the standard increment A (10 or 20) by the equation.
     RR = exp(log(RR for range) x A/range).
 3From (Dockery et al., 1993; Krewski et al., 2000, Part II, Table 21a), original model.
 4From (Krewski et al., 2000), Part II, Table 21c.
 5From (Krewski et al., 2000), Part II, Table 25a.
 Trom (Krewski et al., 2000), Part II, Table 25c.
 7From (Pope et al., 2002).
 8From (Abbey et al., 1999), pooled estimate for males and females.
 9For males only; no significant excess risk for females with contributing respiratory causes.
 10From (McDonnell et al., 2000), using two-pollutant (fine and coarse particle) models.
 "Males only.
 12Males only, exposure period 1979-81, mortality 1982-88 from Table 7 (Lipfert et al., 2000b).
April 2002
8-94
DRAFT-DO NOT QUOTE OR CITE

-------
 1      suggestive of associations with fine particles.  Also, the VA study (Lipfert et al., 2000) found no
 2      association with PM2 5. Nevertheless, the lack of consistent findings in the AHSMOG study and
 3      negative results of the VA study, do not cast doubt on the findings of the Six Cities and ACS
 4      studies; both of the late studies had larger study populations, were based on measured PM data
 5      (in contrast with AHSMOG PM estimates based on TSP or visibility measurements), and have
 6      been validated through exhaustive reanalysis.  When considering the results of these four studies,
 7      including the reanalyses results for the Six Cities and ACS studies and the results of the ACS
 8      study extension, it can be concluded that there is substantial evidence for a positive association
 9      between long-term exposure to PM (especially fine particles) and mortality.
10           There is no obvious statistically significant relationship between PM effect sizes, gender,
11      and smoking status across these studies. The AHSMOG analyses show no significant
12      relationships between PM10 and total mortality or cardiovascular mortality for either sex, and
13      only for male lung cancer incidence and lung cancer deaths in a predominantly non-smoking
14      sample.  The ACS results, in contrast, show similar and significant associations with total
15      mortality for both "never smokers" and "ever  smokers", although the ACS cohort may include a
16      substantial number of long-term former smokers with much lower risk than current smokers.
17      The Six Cities study cohort shows the strongest evidence of a higher PM effect in current
18      smokers than in non-smokers, with female former smokers having a higher risk than male former
19      smokers.  This study suggests  that smoking status may be viewed as an "effect modifier" for
20      ambient PM, just as smoking may be a health  effect modifier for ambient O3 (Cassino et al.,
21      1999).
22           When the ACS study results are compared with the AHSMOG study results for SO4=
23      (PM10_2 5 and PM10 were not considered in the ACS study, but were evaluated in ACS reanalyses
24      [Krewski et al., 2000; Pope et al, 2002]), the total mortality effect sizes per 15 //g/m3 SO4= for the
25      males in the AHSMOG population are seen to fall between the Six-Cities and the ACS effect
26      estimates for males: RR=1.28  for AHSMOG male participants; RR=1.61 for Six-Cities Study
27      male non-smokers; and RR=1.10 for never smoker males in the ACS study.  The AHSMOG
28      study 95% confidence intervals encompass both of those other studies' sulfate RR's.
29
30


        April 2002                                8-95       DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.2.3.3 Studies by Particulate Matter Size-Fraction and Composition
 2      8.2.3.3.1  Six Cites, ACS, andAHSMOG Study Results
 3           Ambient PM consists of a mixture that may vary in composition over time and from place
 4      to place.  This should logically affect the relative toxicity of PM indexed by mass at different
 5      times or locations. Some semi-individual chronic exposure studies have investigated relative
 6      roles of various PM components in contributing to observed air pollution associations with
 7      mortality.  However, only a limited number of the chronic exposure studies have included direct
 8      measurements of chemical-specific constituents of the PM mixes indexed by mass measurements
 9      used in their analyses.
10           As shown in Table 8-13, the Harvard Six-Cities study (Dockery et al.,  1993) results
11      indicated that the PM2 5 and SO4= RR associations (as indicated by their respective 95% CFs and
12      t-statistics) were more consistent than those for the coarser mass components.  However, the
13      effects of sulfate and non-sulfate PM2 5 are indicated to be quite similar.  Acid aerosol (H+)
14      exposure was also considered by Dockery et al. (1993), but only less than one  year of
15      measurements collected near the end of the follow-up period were available in most cities; so, the
16      Six-Cities results were much less conclusive for the acidic component of PM than for the other
17      PM metrics measured over many years during the study.  The Six-Cities study also yielded total
18      mortality RR estimates for the reported range across those cities of PM2 5 and SO4= levels that,
19      although not statistically different, were  roughly double the analogous RR's for the TSP-PM15
20      and PM15.2 5 mass components.
21           Table 8-14 presents comparative PM25 and SO4= results from the ACS study, indicating that
22      both had substantial, statistically significant effects on all-cause and cardiopulmonary mortality.
23      On the other hand, the RR for lung cancer was notably larger (and substantially more significant)
24      for SO4= than PM2 5 (not significant).
25           The most recent AHSMOG study analysis reported by Abbey et al. (1999) used PM10 as its
26      PM mass index, finding some significant associations with total and by-cause  mortality, even
27      after controlling for potentially confounding factors (including other pollutants). This analysis
28      also considered SO4= as a PM index for all health outcomes studied except lung cancer, but SO4=
29      was not as strongly associated as PM10 with mortality and was not statistically significant for any
30      mortality category.
31

        April 2002                                8-96        DRAFT-DO NOT QUOTE OR CITE

-------
             TABLE 8-13. COMPARISON OF ESTIMATED RELATIVE RISKS FOR
              ALL-CAUSE MORTALITY IN SIX U.S. CITIES ASSOCIATED WITH
               THE REPORTED INTER-CITY RANGE OF CONCENTRATIONS
                      OF VARIOUS PARTICULATE MATTER METRICS
PM Species
S04=
PM25 - SO4=
PM25
PM15.2.5
TSP-PM,,
Concentration Range
(^g/m3)
8.5
8.4
18.6
9.7
27.5
Relative Risk
Estimate
1.29
1.24
1.27
1.19
1.12
RR
95% CI
(1.06-1.56)
(1.16-1.32)
(1.06-1.51)
(0.91-1.55)
(0.88-1.43)
Relative Risk
t-Statistic
3.67
8.79
3.73
1.81
1.31
       Source: Dockery et al. (1993); U.S. Environmental Protection Agency (1996a).
            TABLE 8-14.  COMPARISON OF REPORTED SO4= AND PM2 5 RELATIVE
               RISKS FOR VARIOUS MORTALITY CAUSES IN THE AMERICAN
                              CANCER SOCIETY (ACS) STUDY
Mortality Cause
All Cause
Cardiopulmonary
Lung Cancer
SO4=
(Range = 19.9 //g/m3)
Relative
Risk
1.15
1.26
1.35
RR
95% CI
(1.09-1.22)
(1.15-1.37)
(1.11-1.66)
RR
t-Statistic
4.85
5.18
2.92
PM25
(Range = 24.5 ^g
Relative
Risk
1.17
1.31
1.03
RR
95% CI
(1.09-1.26)
(1.17-1.46)
(0.80-1.33)
/m3)
RR
t-Statistic
4.24
4.79
0.38
       Source: Pope etal. (1995).
1          Also, very extensive results were reported in Lipfert et al. (2000b) for various components:
2     TSP, PM10, PM2 5, PM15_2 5, PM15, SO4=. There were no significant positive effects for any
3     exposure period concurrent or preceding the mortality period for any PM component, but there
4     was for O3.
5          Results from the Harvard Six Cities, the ACS, and the AHSMOG studies are compared in
6     Table 8-15 (for total mortality) and Table 8-16 (for cause-specific mortality). Results for the VA
7     study are not shown in Tables 8-15 and 8-16 for two reasons. First of all, the cohort is male and

      April 2002                             8-97       DRAFT-DO NOT QUOTE OR CITE

-------
             TABLE 8-15.  COMPARISON OF TOTAL MORTALITY RELATIVE RISK
        ESTIMATES AND T-STATISTICS FOR PARTICULATE MATTER COMPONENTS
                          IN THREE PROSPECTIVE COHORT STUDIES
PM Index
PM10 (50 Mg/m3)


PM2 5 (25 Mg/m3)



S04= (15 Mg/m3)




Days/yr. with
PM10>100 Mg/m3
(30 days)
PMio-2.5 (25 Mg/m3)

Study
Six Cities

AHSMOG
Six Cities

ACS (50 cities)

Six Cities

ACS (151 cities)

AHSMOG
AHSMOG
Six Cities

Subgroup
All
Male Nonsmoker
Male Nonsmoker
All
Male Nonsmoker
All
Male Nonsmoker
All
Male Nonsmoker
All
Male Nonsmoker
Male Nonsmoker
Male Nonsmoker
All
Male Nonsmoker
Relative Risk
1.504a; 1.530b
1.280a
1.242
1.364a; 1.379b
1.207a
1.174
1.245
1.504a; 1.567b
1.359
1.111
1.104
1.279
1.082
1.814a; 1.560b
1.434a
t Statistic
2.94a; 3.27b
0.81a
1.616
2.94a; 3.73b
0.81a
4.35
1.96
2.94a; 3.67b
0.81a
5.107
1.586
0.960
2.183
2.94a'c 1.816b
0.81a
        "Method 1 compares Portage vs. Steubenville (Table 3, Dockery et al., 1993).
        bMethod 2 is based on ecologic regression models (Table 12-18, U.S. Environmental Protection Agency, 1996a).
        "Method 1 not recommended for PM10_2 5 analysis, due to high concentration in Topeka.
1     largely current or former smokers (81%), thus not comparable to the total or male non-smoker
2     populations.  Secondly, there is a wide variety of exposure periods and mortality periods.
3          Estimates for Six Cities parameters were calculated in two ways: (1) mortality RR for the
4     most versus least polluted city in Table 3 of Dockery et al. (1993) adjusted to standard
5     increments; and (2) ecological regression fits in Table 12-18 of U.S. Environmental Protection
6     Agency (1996a). The Six Cities study of eastern  and mid-western U.S. cities suggests a strong
7     and highly significant relationship for fine particles and sulfates, a slightly weaker but still highly
      April 2002
8-98
DRAFT-DO NOT QUOTE OR CITE

-------
        TABLE 8-16. COMPARISON OF CARDIOPULMONARY MORTALITY RELATIVE
              RISK ESTIMATES AND T-STATISTICS FOR PARTICULATE MATTER
                  COMPONENTS IN THREE PROSPECTIVE COHORT STUDIES
PM Index
PM10 (50 //g/m3)


PM25 (25 //g/m3)



S04= (15 ^g/m3)





Days/yr. with
PM10>100 (30 days)

PM10.2.5 (25 ^g/m3)
Study
Six Cities
AHSMOG

Six Cities
ACS (50 cities)


Six Cities
ACS (151 cities)


AHSMOG

AHSMOG

Six Cities
Subgroup
All
Male Nonsmoker
Male Non-CRCc
All
All
Male
Male Nonsmoker
All
All
Male
Male Nonsmoker
Male Nonsmoker
Male Non.-CRCc
Male Nonsmoker
Male Non.-CRCc
All
Relative Risk
1.744a
1.219
1.537
1.527a
1.317
1.245
1.245
1.743a
1.190
1.147
1.205
1.279
1.219
1.082
1.188
2.251s
t Statistic
2.94a
1.120
2.369
2.94a
4.699
3.061
1.466
2.94a
5.470
3.412
2.233
0.072
0.357
1.310
2.370
2.94a'b
       "Method 1 compares Portage vs. Steubenville (Table 3, Dockery et al., 1993).
       bMethod 1 not recommended for PM10.2.5 analysis due to high concentration in Topeka.
       "Male non. - CRC = AHSMOG subjects who died of any contributing non-malignant respiratory cause.
1
2
3
4
5
significant relationship to PM10, and a marginal relationship to PM10_2 5.  The ACS study looked at
a broader spatial representation of cities, and found a stronger statistically significant relationship
to PM2 5 than to sulfate (no other pollutants were examined). The AHSMOG study at California
sites (where sulfate levels are typically low) found significant effects in males for PM10
100 //g/m3 exceedances and a marginal effect of mean PM10, but no PM effects for females or
with sulfates.  On balance, the overall results shown in Tables 8-15 and  8-16 suggest statistically
      April 2002
                                         8-99
DRAFT-DO NOT QUOTE OR CITE

-------
 1      significant relationships between long-term exposures to PM10, PM2 5, and/or sulfates and excess
 2      total and cause-specific cardiopulmonary mortality.
 3           Overall, the semi-individual long-term PM exposure studies conducted to-date collectively
 4      confirm earlier cross-sectional study indications that the fine mass component of PM10 (and
 5      usually especially its sulfate constituent) are more strongly correlated with mortality than is the
 6      coarse PM10_25 component. However, the greater precision of PM2 5 population exposure
 7      measurement (both analytical and spatial) relative to PM10_2 5 makes conclusions regarding their
 8      relative contributions to observed PM10-related associations less certain than if the effect of their
 9      relative errors of measurement could be addressed.
10           Single-pollutant results about PM components are informative, as shown in Table 8-15 for
11      total mortality and in  Table 8-16 for cardiopulmonary causes.  The t-statistics are compared for
12      studies where appropriate: mean PM10, PM10_2 5, PM2 5, and sulfate for the Six Cities (Dockery
13      et al., 1993); mean PM2 5 and sulfate for ACS  (Pope et al., 1995); mean PM10 and sulfate, and
14      PM10 exceedances of 100 //g/m3 for AHSMOG (Abbey et al., 1999).
15
16      8.2.3.3.2  Lipfert and Morris (2002): An Ecological Study
17           Although we have identified reasons for preferring to use prospective cohort studies to
18      assess the long-term exposure effects of particles and gases, additional useful information may
19      still be provided by ecological  studies, particularly by repeated cross-sectional studies that may
20      provide another tool for examining changes in air-pollution-attributable mortality over time.
21      Lipfert and Morris (2002) carried out cross-sectional regressions for five time periods using
22      published data on mortality, air pollution, climate, and socio-demographic factors using county-
23      level data. Data were available for TSP and gaseous co-pollutants as far back as 1960  and for
24      PM25, PM15, and SO4= from the IPN. Attributable mortality at ages 45+ for 1979-1981 was
25      associated with TSP 1960-64, less strongly with TSP 1970-1974, but not with concurrent (1979-
26      1981) TSP. Attributable mortality for ages 45+ in 1979-1981 was associated with PM25 and
27      SO4= but not PM15 for 1979-1984. However, SO4= for most intervals 1960-64 up to 1979-1981
28      was associated with mortality for most ages .  Concurrent SO2 (1979-1981) was associated with
29      mortality, but much less for earlier years.
30           Pollution-attributable mortality in 1989-91 was no longer significantly associated with TSP,
31      but remained significantly associated with PM2 5 and SO4= for ages 45+ for most time intervals:

        April 2002                                8-100       DRAFT-DO NOT QUOTE OR CITE

-------
 1      1979-84, 1999, forPM25, and 1970-74, 1979-81, 1979-84 (fines), 1982-88 for SO4=. Pollution-
 2      attributable mortality in 1995-1997 had little association with present or previous PM25 and
 3      PM10, but a reasonably consistent and positive relationship to SO4=. There appeared to be a
 4      systematic decrease in the TSP, IPN, PM2 5, and PM10 effects from the 1960s to the 1990s, and in
 5      the AIRS and IPN SO4= effect over time, but an increase in the AIRS PM2 5 effect and in the NO2
 6      and peak O3 effects.
 7           One of the journal editors (Ayres, 2002) notes that this study uses some other ecological
 8      variables that may improve the model.  Two of the ecological variables, (vehicle miles of travel
 9      per square mile per year by gasoline (VMTG)  and diesel (VMTD) vehicles respectively in a
10      county, also used in Janssen et al., 2002) are likely to have important associations with air
11      pollution.  As noted earlier, some ambient pollutants associated with fuel combustion have
12      higher concentrations near main roads, such as PM10_25 (EC if from diesel exhaust) , NO2, and
13      CO, whereas other pollutants such as O3 may have higher concentrations away from major
14      highways.
15
16      8.2.3.4  Population-Based Mortality Studies in Children
17           Older cross-sectional mortality studies suggest that the very young may represent an
18      especially susceptible sub-population for PM-related mortality. For example, Lave and Seskin
19      (1977) found mortality among those 0-14 years of age to be significantly associated with TSP.
20      More recently, Bobak and Leon (1992) studied neonatal (ages < 1 mo) and post-neonatal
21      mortality (ages 1-12 mo) in the Czech Republic and reported significant and robust associations
22      between post-neonatal mortality and PM10, even after considering other pollutants.  Post-neonatal
23      respiratory mortality showed highly significant associations for all pollutants considered, but only
24      PM10 remained significant in simultaneous regressions. The exposure duration was longer than a
25      few days, but shorter than in the adult prospective cohort studies. Thus, the limited available
26      studies reviewed in the 1996 PM AQCD were highly suggestive of an association between
27      ambient PM  concentrations and infant mortality, especially among post-neonatal infants.
28           More recent studies since the 1996 PM AQCD have focused specifically on ambient PM
29      relationships to (a) intrauterine mortality and morbidity and (b) early post neonatal mortality. In
30      a study by Pereira et al. (1998), of intrauterine (pre-natal) mortality during one year (1991-1992)
31      in Brazil, PM10 was not found to be a significant predictor, but involvement of CO was suggested

        April 2002                                8-101       DRAFT-DO NOT QUOTE OR CITE

-------
 1      by an association between increased carboxyhemoglobin (COHb) in fetal blood and ambient CO
 2      levels on the day of delivery measured in a separate study. Another study (Dejmek et al., 1999)
 3      evaluated possible impacts of ambient PM10 and PM2 5 exposure (monitored by EPA-developed
 4      VAPS methods) during pregnancy on intrauterine growth retardation (IUGR) risk in the highly
 5      polluted Teplice District of Northern Bohemia in the Czech Republic during three years
 6      (1993-1996).  Mean levels of pollutants (PM, NO2, SO2) were calculated for each month of
 7      gestation and three concentration intervals (low, medium, high) derived for each pollutant.
 8      Preliminary analyses found significant associations of IUGR with SO2 and PM10 early in
 9      pregnancy but not with NO2. Odds ratios for IUGR for PM10 and PM2 5 levels were determined
10      by logistic regressions for each month during gestation, after adjusting for potential confounding
11      factors (e.g., smoking, alcohol consumption during pregnancy, etc.). Definition of an IUGR birth
12      was any one for which the birth weight fell below the 10th percentile by gender and age for live
13      births in the Czech Republic (1992-93). The OR's for IUGR were significantly related to PM10
14      during the first month of gestation: that is, as compared to low PM10, the medium level PM10
15      OR = 1.47 (CI 0.99-2.16), and the high level PM10 OR = 1.85 (CI 1.29-2.66). PM25 levels were
16      highly correlated with PM10 (r = 0.98) and manifested similar patterns (OR  = 1.16, CI 0.08-0.69
17      for medium PM25 level; OR = 1.68, CI 1.18-2.40 for high PM25 level). These results suggest
18      effects of PM exposures (probably including fine particles such as sulfates,  acid aerosols, and
19      PAHs in the Teplice ambient mix) early in pregnancy (circa embryo implantation) on fetal
20      growth and development.
21           A recent study relating air pollution to birth weight in the metropolitan Reno, Nevada area
22      (Chen et al., 2002) examined the associations between air pollutant variables and birth weight
23      (BW) as a continuous variable and the prevalence of low birth weight (LEW, BW < 2500 gtn) as
24      a dichotomous variable. Mean daily concentrations of the pollution variables PM10, O3, and CO
25      were relatively low:  31.5 //g/m3 for PM10 (range 1 to 157), 27.2 ppb for O3  (range 2.8 to 62), and
26      1.0 ppm for CO (range 0.25 to 4.9).  Ordinary least squares regression of BW on one, two,  or
27      three air pollutants, and numerous covariates (e.g., age, race, education, prenatal care, maternal
28      behaviors) were included in the models. Third-trimester maternal exposure to PM10 was
29      significantly associated with an approximately 1 g reduction in BW per //g/m3 PM10, a finding
30      robust across different model specifications. Another finding was that the reduction in BW was
31      9 to 12 g for third-trimester exposures > 90th percentile PM10 (45 //g/m3). However, none of the

        April 2002                                8-102       DRAFT-DO NOT QUOTE OR CITE

-------
 1      PM odds ratios were significantly associated with increased risk of LEW.  Neither ambient CO
 2      nor O3 were associated were significantly associated with LEW, unlike findings for intrauterine
 3      growth reduction (IUGR) in Los Angeles by Ritz and Yu (1999) and Ritz et al. (2002).
 4           More consistent results indicating likely early post-natal PM exposure effects on neonatal
 5      infant mortality have emerged from other new studies. Woodruff et al. (1997), for example, used
 6      cross-sectional methods to evaluate possible association of post-neonatal mortality with ambient
 7      PM10 pollution.  This study involved an analysis of a cohort of circa 4 million infants born during
 8      1989 - 1991 in 86 U.S. metropolitan statistical areas (MSAs).  Data from the National Center for
 9      Health Statistics-linked birth/infant death records were combined at the MSA level with PM10
10      data from EPA's Aerometric database. Infants were categorized as having high, medium, or low
11      exposures based on tertiles of PM10 averaged over the first 2 postnatal months. Relationships
12      between this early neonatal PM10 exposure and total and cause-specific post-neonatal mortality
13      rates (from 1 mo to 1 y of age) were examined using logistic regression analyses, adjusting for
14      demographic and environmental factors.  Overall post-neonatal mortality rates per 1,000 live
15      births were 3.1 among infants in areas with low PM10 exposures, 3.5  among infants with medium
16      PM10 exposures, and 3.7 among highly PM exposed infants. After adjustment for covariates, the
17      odds ratio (OR) and 95% confidence intervals for total post-neonatal mortality for the high
18      versus the low exposure group was 1.10 (CI=1.04-1.16). In normal birth weight infants, high
19      PM10 exposure was associated with mortality for respiratory causes (OR = 1.40, CI=1.05-1.85)
20      and sudden infant death syndrome (OR = 1.26, CI=1.14-1.39).  Among low birth weight babies,
21      high PM10 exposure was positively (but not significantly) associated with mortality from
22      respiratory causes (OR = 1.18, CI=0.86-1.61).  However, other pollutants (e.g., CO) were not
23      considered as possible confounders.  This study provides results consistent with some earlier
24      reports indicating that outdoor PM air pollution may be associated with increased risk of post-
25      neonatal mortality (e.g., Bobak and Leon, 1992), but lack of consideration of other air pollutants
26      as potential confounders in this new study reduces the certainty that PM is the specific causal
27      outdoor air pollutant in this case.
28           Lipfert et al. (2000c) have reported replicating the basic findings of Woodruff et al. (1997)
29      using a similar modeling approach but annual average PM10 air quality data for one year (1990)
30      instead of PM10 averaged over the first two post natal months during 1989-1991.  The
31      quantitative relationship between the individual risk of infant mortality did not differ among

        April 2002                                 8-103        DRAFT-DO NOT QUOTE OR CITE

-------
 1      infant categories (by age, by birthweight, or by cause), but PM10 risks for SIDs deaths were
 2      higher for babies of smoking mothers.  SO4= was a strong negative predictor of SIDs mortality for
 3      all age and birth weight categories.  The authors (a) noted difficulties in ascribing the reported
 4      PM10 and SO4= associations to effects of the PM pollutants per se versus the results possibly
 5      reflecting interrelationships between the air pollution indices, a strong well-established
 6      East-West gradient in U.S. SIDS cases, and/or underlying sociodemographic factors (e.g., the
 7      socioeconomic or education level of parents) and (b) hypothesized that a parallel gradient in use
 8      of wood burning in fireplaces or woodstoves and consequent indoor wood smoke exposure might
 9      explain the observed cross-sectional study results.
10           The basic findings from Woodruff et al. (1997) also appear to be bolstered by  a more recent
11      follow-up study by Bobak and Leon (1999), who conducted a matched population-based
12      case-control  study covering all births registered in the Czech Republic from 1989 to 1991 that
13      were linked to death records. They used conditional logistic regression to estimate the effects of
14      suspended particles and nitrogen oxides on risk of death in the neonatal  and early post-neonatal
15      period, controlling for maternal socioeconomic status and birth weight, birth length, and
16      gestational age. The effects of all pollutants were strongest in the post-neonatal period and
17      specific for respiratory causes.  Only PM showed a consistent association when all pollutants
18      were entered in one model.  Thus, in this study, it appears that long-term exposure to PM is the
19      air pollutant  metric most strongly associated with excess post-neonatal deaths.
20           A study of changes in annual air pollution and infant mortality over time (rather than
21      spatially) in the U.S. was also recently conducted for the period  1981-1982 (Chay and
22      Greenstone,  2001a,b).  These studies used sharp, differential air quality changes across sites
23      attributable to geographic variation in the effects of the 1981-1982 recession to estimate the
24      relationship between PM air pollution and infant mortality.  During the narrow period of these
25      two years, there was substantial variation across counties in changes in particulate (TSP)
26      pollution and these differential pollution reductions appeared to be independent of changes in
27      numerous socioeconomic and health care factors that may be related to infant mortality.  The
28      authors found that a 1 ug/m3 reduction in TSP resulted in about 4-8 fewer infant deaths per
29      100,000 live births at the county level (a 0.35-0.45 elasticity), the estimates being remarkably
30      stable across a variety  of specifications. The estimated effects in this  study were driven almost
31      entirely by fewer deaths  occurring within one month and one day of birth (i.e., neonatal),

        April 2002                                 8-104        DRAFT-DO NOT QUOTE OR CITE

-------
 1      suggesting that fetal exposure to pollution (via the mother) may have adverse health
 2      consequences. Findings of the population reductions in infant birth weight in this study provide
 3      evidence consistent with the infant mortality effects found, suggestive of a causal relationship
 4      between PM exposure and infant mortality.
 5           The study by Loomis et al. (1999) of infant mortality in Mexico City during 1993-1995
 6      adds additional interesting information pointing towards likely fine particle impacts on infant
 7      mortality.  That is, in Mexico City (where mean 24-h PM2 5 = 27.4 //g/m3), infant mortality was
 8      found to be associated with PM2 5, NO2, and O3 in single pollutant GAM Poisson models, but
 9      much less consistently with NO2 and O3 than PM2 5 in multipollutant models. The estimated
10      excess risk for PM25-related infant mortality lagged 3-5 days was 18.2% (95% CI 6.4, 30.7) per
11      25 //g/m3 PM2 5.  It is not clear, however, the extent to which such a notable increased risk for
12      infant mortality might be extrapolated to U.S. situations, due to possible differences in prenatal
13      maternal or early postnatal infant nutritional status.
14
15      8.2.3.5  Salient Points Derived from Analyses of Chronic Particulate Matter Exposure
16              Mortality Effects
17           A review of the studies summarized in the previous PM AQCD (U.S. Environmental
18      Protection Agency, 1996a) indicates that past epidemiologic studies of chronic PM exposures
19      collectively indicate increases in mortality to be associated with long-term exposure to airborne
20      particles of ambient origins. The PM effect size estimates for total mortality from these studies
21      also indicate that a substantial portion of these deaths reflected cumulative PM impacts above
22      and beyond those exerted by acute exposure events.
23           The recent HEI-sponsored reanalyses of the ACS  and Harvard  Six-Cities studies (Krewski
24      et al., 2000) "replicated the original results, and tested those results against alternative risk
25      models and analytic approaches without substantively altering the original findings of an
26      association between indicators of paniculate matter air pollution and mortality." Several
27      questions, including the questions (1-4) posed at the outset of this Section (8.2.3) were
28      investigated by the Krewski et al. (2000) sensitivity analyses for the Six City and ACS studies
29      data sets. Key results emerging from the HEI reanalyses and other new chronic PM mortality
30      studies are as follow:
        April 2002                                8-105       DRAFT-DO NOT QUOTE OR CITE

-------
 1           (1)  A much larger number of confounding variables and effects modifiers were considered
 2      in the Reanalysis Study than in the original Six City and ACS studies.  The only significant air
 3      pollutant  other than PM2 5 and SO4 in the ACS study was SO2, which greatly decreased the PM2 5
 4      and sulfate effects when included as a co-pollutant (Krewski et al., 2000, Part n, Tables 34-38).
 5      A similar reduction in particle effects occurred in any multi-pollutant model with SO2. The most
 6      important new effects modifier was education.  The AHSMOG study suggested that other metrics
 7      for air pollution, and other personal covariates such as time spent outdoors and consumption of
 8      anti- oxidant vitamins, might be useful. Both individual- level covariates and ecological-level
 9      covariates shown in (Krewski et al., 2000, Part n, Table 33) were evaluated.
10           (2)  Specific attribution of excess long-term mortality to any specific particle component or
11      gaseous pollutant was refined in the reanalysis of the ACS study. Both PM25 and sulfate were
12      significantly associated with excess total mortality and cardiopulmonary mortality and to  about
13      the same  extent whether the air pollution data were mean or median long-term concentrations or
14      whether based on Original Investigator or Reanalysis Team data. The association of mortality
15      with PM15 was much smaller, though still significant,  and the associations with the coarse
16      fraction (PM15.2 5) or TSP were even smaller and not significant. The lung cancer effect was
17      significant only for sulfate with the original investigator data or for new investigators with
18      regional sulfate artifact adjustment for the 1980-1981  data (Krewski et al., 2000, Part II,
19      Table 31). Associations of mortality with long-term mean concentrations of criteria gaseous
20      co-pollutants were generally non-significant except for SO2 (Krewski et al., 2000, Part n,  Tables
21      32, 34-38) which was highly significant, and for cardiopulmonary disease with warm-season
22      ozone.  However, the regional association of SO2 with SO4 and SO2 with PM2 5 was very high,
23      and the effects of the separate pollutants could not be  distinguished. Krewski et al. (2000,
24      p. 234) concluded that, "Collectively, our reanalyses suggest that mortality may be associated
25      with more than one component of the complex mix of ambient air pollutants in urban areas of the
26      United States." In the most recent extension of the ACS study, Pope et al. (2002) confirmed the
27      strong association with SO2 but found little evidence of effects for long-term exposures to other
28      gaseous pollutants.
29           (3)  The extensive temporal data on air pollution concentrations over time in the Six City
30      Study allowed the Reanalysis Team to evaluate time scales for mortality for long-term exposure
31      to a much greater extent than reported in Dockery et al. (1993). The first approach was to

        April 2002                                8-106        DRAFT-DO NOT QUOTE OR CITE

-------
 1      estimate the log- hazard ratio as a function of follow up time using a flexible spline-function
 2      model (Krewski et al., 2000, Part H, Figures 2 and 3).  The results for both SO4 and PM25 suggest
 3      very similar relationships, with larger risk after initial exposure decreasing to 0 after about 4 or
 4      5 years, and a large increase in risk at about 10 years follow-up time.
 5           The analyses of the ACS Study proceeded somewhat differently, with less temporal data
 6      but many more cities.  Flexible spline regression models for PM2 5 and sulfate as function of
 7      estimated cumulative exposure (not defined) were very nonlinear and showed quite different
 8      relationships (Krewski et al., 2000, Part II, Figures 10 and 11).  The PM2 5 relationship shows the
 9      mortality log-hazard ratio increasing up to about 15 //g/m3 and relatively flat above about
10      22 //g/m3, then increasing again. The sulfate relationship is almost piecewise linear, with a low
11      near- zero slope below about 11 //g/m3 and a steep  increase above that concentration.
12           A third approach evaluated several time-dependent PM2 5 exposure indicators in the Six
13      City study.  They are:  (a) constant (at the mean) over the entire follow-up period; (b) annual
14      mean within each  of the 13 years of the study; (c) city-specific mean concentration for the earliest
15      years of the study, i.e., very long-term effect; (d) exposure estimate in 2  years preceding death;
16      (e) exposure estimate in 3 to 5 years preceding death; (f) exposure estimate > 5 years preceding
17      death.  The time-dependent estimates (a-e) for mortality risk are generally similar and statistically
18      significant (Krewski et al., 2000, Part II, Table 53), with RR of 1.14 to 1.19 per 24.5 //g/m3 being
19      much lower than the risk of 1.31 estimated for exposure at the constant mean for the period.
20      Thus, it is highly likely the duration and time patterns of long-term exposure affect the risk of
21      mortality, and further study of this question (along with that of mortality displacement from
22      short-term exposures) would improve estimates of life-years lost from PM exposure.
23           (4)  The Reanalysis Study also advanced our understanding of the  shape of the relationship
24      between mortality and PM. Again using flexible spline modeling, Krewski et al. (2000, Part n,
25      Figure  6) found a visually near-linear relationship between all-cause and cardiopulmonary
26      mortality residuals and mean sulfate concentrations, near-linear between cardiopulmonary
27      mortality and mean PM2 5, but a somewhat nonlinear relationship between all-cause mortality
28      residuals and mean PM2 5 concentrations that flattens above about 20 //g/m3. The confidence
29      bands around the fitted curves are very wide, however, neither requiring a linear relationship nor
30      precluding a nonlinear relationship if suggested by reanalyses.  An investigation of the mortality


        April 2002                                8-107       DRAFT-DO NOT QUOTE OR CITE

-------
 1      relationship for other indicators may be useful in identifying a threshold, if one exists, for chronic
 2      PM exposures.
 3           (5) With regard to the role of various PM constituents in the PM-mortality association, past
 4      cross-sectional studies have generally found the fine particle component, as indicated either by
 5      PM2 5 or sulfates, to be the PM constituent most consistently associated with mortality. While
 6      relative measurement errors of various PM indicators must be further evaluated as a possible
 7      source of bias in these estimate comparisons, the Six-Cities and AHSMOG prospective semi-
 8      individual studies both indicate that the fine mass components of PM are more strongly
 9      associated with mortality effects of chronic PM exposure than are coarse fraction indicators.
10
11
12      8.3 MORBIDITY EFFECTS OF PARTICULATE  MATTER EXPOSURE
13           This morbidity discussion is presented below in several subsections, dealing with:  (a) acute
14      cardiovascular morbidity effects of ambient PM exposure; (b) effects of short-term PM exposure
15      on the incidence of respiratory and other medical visits and hospital admissions; and (c) short-
16      and long-term PM exposure effects on lung function and respiratory symptoms in asthmatics and
17      non-asthmatics.
18
19      8.3.1  Cardiovascular Effects Associated with Acute Ambient Particulate
20             Matter Exposure
21      8.3.1.1  Introduction
22           Very little information specifically addressing acute cardiovascular morbidity effects  of PM
23      existed  at the time of the 1996 PM AQCD. Since that time, a significantly expanded body  of
24      literature has emerged, both on the ecologic relationship between ambient particles and
25      cardiovascular hospital admissions and on physiological and/or biochemical measures that  have
26      been associated with PM exposures.  The latter studies are particularly important in that they
27      suggest possible mechanisms.
28           This section begins with a brief summary of the conclusions that were reached in the  1996
29      PM AQCD regarding acute cardiovascular impacts of PM. Next, new studies are reviewed in the
30      two categories noted above, i.e., ecologic time series studies and individual-level studies of
31      physiological measures of cardiac function and/or biochemical measures in blood as they relate
        April 2002                               8-108        DRAFT-DO NOT QUOTE OR CITE

-------
 1      to ambient pollution. This review is followed by discussion of several issues that are important
 2      in interpreting the available data, including the identification of potentially susceptible sub-
 3      populations, the roles of environmental co-factors such as weather and other air pollutants,
 4      temporal lags in the relationship between exposure and outcome, and the relative importance of
 5      various size-classified PM components (e.g., PM25, PM10, PM10_25).
 6           Studies of cardiovascular PM effects presented in this section were identified by ongoing
 7      Medline searches in conjunction with other search strategies.  Specific studies were summarized
 8      in text and/or tables based on criteria that include the following:  (1) preference was given to
 9      results reported for PM10,  PM10_2 5, and PM2 5; (2) studies relating cardiovascular effects to levels
10      of ambient PM exposure in a quantitative manner are the focus of presentations; and (3) other
11      factors discussed earlier in Section 8.1 of this chapter.
12
13      8.3.1.2  Summary of Key Findings on Cardiovascular Morbidity from the 1996 Particulate
14              Matter Air quality Criteria Document
15           Just two studies were available for review in the  1996 PM  AQCD that provided data on
16      acute cardiovascular morbidity outcomes (Schwartz and  Morris, 1995; Burnett et al.,  1995).
17      Both studies were of ecologic time series design, using standard statistical methods. Analyzing
18      four years of data on the > 65 year old Medicare population in Detroit, MI, Schwartz and Morris
19      (1995) reported significant associations between ischemic heart  disease admissions and PM10,
20      controlling for environmental covariates. Based on an analysis of admissions data from
21      168 hospitals throughout Ontario, Canada, Burnett and colleagues (1995) reported significant
22      associations between fine  particle sulfate concentrations, as well as other air pollutants, and daily
23      cardiovascular admissions. The relative risk due to sulfate particles was slightly larger for
24      respiratory than for cardiovascular hospital admissions.  The 1996 PM AQCD concluded on the
25      basis of these studies that: "There is a suggestion of a relationship to heart disease, but the
26      results are based on only two studies, and the estimated effects are smaller than those for other
27      endpoints" (U.S. Environmental Protection Agency, 1996a p. 12-100). The PM AQCD went on
28      to state that acute impacts on CVD admissions had been  demonstrated for elderly populations
29      (i.e., > 65),  but that insufficient data existed to assess relative impacts on younger populations.
30           When viewed alongside the more extensive literature on acute CVD mortality that was
31      available at that time, the  evidence from ecologic time series studies reviewed in the 1996 PM

        April 2002                                8-109       DRAFT-DO NOT QUOTE OR CITE

-------
 1      AQCD was consistent with the notion that acute health risks of PM are larger for cardiovascular
 2      and respiratory causes than for other causes. Given the tendency for end-stage disease states to
 3      include both respiratory and cardiovascular impairment, and the associated diagnostic overlap
 4      that often exists, it was not possible on the basis of these studies alone to determine which of the
 5      two organ systems, if either, was more critically impacted.
 6
 7      8.3.1.3 New Particulate Matter-Cardiovascular Morbidity Studies
 8      8.3.1.3.1  Acute Hospital Admission Studies in United States Cities
 9           Numerous new studies have examined associations between daily measures of ambient PM
10      and daily hospital admissions for cardiovascular disease (see Table  8-17 and Table 8B-1 in
11      Appendix 8B). Of particular relevance are two new multi-city studies (Schwartz, 1999; Samet
12      et al., 2000a,b; Zanobetti et al., 2000a), which provide evidence substantiating significant PM
13      effects on cardiovascular-related hospital admissions and visits. Numerous other studies, carried
14      out by individual investigators in a variety of locales, present a more varied picture, especially
15      when gaseous co-pollutants have been analyzed on equal footing with PM.
16           For example, Schwartz (1999) extended the analytical approach he had used in Tucson
17      (described below) to eight more U.S. metropolitan areas, limiting analyses to a single county  in
18      each  location to enhance representativeness of the air pollution data. The locations analyzed
19      were: Chicago, IL; Colorado Springs, CO; New Haven, CT; Minneapolis, MN; St. Paul, MN;
20      Seattle, WA; Spokane, WA; and Tacoma, WA. Again, the analyses focused on total
21      cardiovascular (CVD) hospital admissions among persons >65 years old.  In univariate
22      regressions, remarkably consistent PM10 associations with CVD admissions were found across
23      the eight locations, with a 50 //g/m3 increase in PM10 associated with 3.6 to 8.6% increases in
24      admissions. The univariate eight-county pooled PM10 effect was 5.0% (CI 3.7-6.4), similar to the
25      6.1 % effect per 50 //g/m3 observed in the previous Tucson analysis. In a bivariate model that
26      included CO, the pooled PM10 effect size diminished somewhat to 3.8% (CI 2.0-5.5) and the CO
27      association with CVD admissions was generally robust to inclusion of PM10 in the model.
28           Additional new results were based on analyses of daily CVD hospital admissions in persons
29      65 and older in relation to PM10 in 14 cities from the NMMAPS multi-city study (Samet et al.,
30      2000a,b). Cities included Birmingham, AL; Boulder, CO; Canton, OH; Chicago, IL; Colorado
31      Springs,  CO; Detroit, MI; Minneapolis/ St. Paul, MN; Nashville, TN;  New Haven, CT;

        April 2002                                8-110        DRAFT-DO NOT QUOTE OR CITE

-------
TABLE 8-17. SUMMARY OF STUDIES OF PM10 OR PM2 5 AND TOTAL CVD HOSPITAL VISITS
^
to
o
o
to

Reference citation,
location, etc.


Outcome Measure

Mean Paniculate
Levels (IQR) ^g/m3

Co-pollutants
Analyzed with PM


Lag Structure
Effect measures standardized to
50 Mg/m3 PM10 or 25 ^g/m3
PM * PM **
rlV12.5 •> rlV110-2.5
U.S. Results Without Co-pollutants











oo
^




O
^
-LJ
H
1
O
n
s.**
O
H
O
O
H
W
0
^\
Samet et al. (2000a,b)
14 Cities
Schwartz (1999)
8 Counties
Linn et al. (2000)
Los Angeles
Schwartz (1997)
Tucson, AZ
Gwynn et al. (2000)
Buffalo, NY
Moolgavkar (2000b)
Cook County, IL
Moolgavkar (2000b)
Los Angeles County, CA
Moolgavkar (2000b)
Maricopa County, AZ
Zanobetti et al., 2000a
Cook County, IL
Tolbertetal., (2000a)
Atlanta, GA 1993-1998

Tolbertetal., (2000a)
Atlanta, GA 1998-1999


Total CVD admiss.
> 65 yrs
Total CVD admiss. > 65 yrs

Total CVD admiss. > 30 yrs

Total CVD admiss. > 65 yrs

CVD HA

Total CVD admiss. > 65 yrs

Total CVD admiss. > 65 yrs

Total CVD admiss. > 65 yrs

Total CVD admiss. > 65 yrs

Total CVD emerg. dept.
visits, > 16 yrs

Total CVD emerg. dept.
visits, > 16 yrs


Mean 24.4-45.3

Median 23-37

45, 18

42, IQR 23

mn/max 24. 1/90.8

35, IQR 22

44, IQR 26

41, IQR 19

Median 3 3, IQR 23

30.1, 12.4
Period 1

29.1, 12.0
Period 2


none

none

none

none

none

none

none

none

none

none


none



Oday

Oday

Oday

Oday

3 day

Oday

Oday

Oday

0-1 day avg.

0-2 day avg.


0-2 day avg.



5.5% (4.7, 6.2)

5.0% (3.7, 6.4)

3. 25% (2.04, 4.47)

6.07% (1.12, 1.27)

5.7% (-3.3, 15.5)

4.2% (3.0, 5.5)

3.2% (1.2, 5.3)
4.3% (2.5, 6.1)*
-2.4% (-6.9, 2.3)

6.6% (4.9, 8.3)

-8.2%(p=0.002)


5.1% (-7.9, 19.9)
6.1% (-3.1, 16.2)*
17.6% (-4.6, 45.0)**

U.S. Results With Co-pollutants
Schwartz (1999)
8 Counties

Schwartz (1997)
Tucson, AZ
Total CVD admiss. > 65 yrs


Total CVD admiss. > 65 yrs

Median 23-37


42, IQR 23

CO


CO

Oday


Oday

3. 8% (2.0, 5.5)


5.22% (0.17, 10.54)


-------
                TABLE 8-17 (cont'd). SUMMARY OF STUDIES OF PM10 OR PM2 5 AND TOTAL CVD HOSPITAL VISITS
to
o
o
to
oo
fe
H
6
o
o
H
O
O
H
W
O
O
Reference citation,
location, etc. Outcome Measure
Mean Paniculate
Levels (IQR)
Co-pollutants Lag Structure
Analyzed with PM
Effect measures standardized to
50 Mg/m3 PM10 or 25 ^g/m3
PM25*
U.S. Results With Co-pollutants (cont'd)
Moolgavkar (2000b) Total CVD admiss. > 65 yrs
Cook County, IL
Moolgavkar (2000b) Total CVD admiss. > 65 yrs
Los Angeles County, CA
35, IQR22
44, IQR 26
NO2 0 day
none 0 day
1.8% (0.4, 3.2)
-1.8% (-4.4, 0.9)
0.8% (-1.3, 2.9)*
Non-U.S. Results Without Co-pollutants
Burnett et al., (1997a) Total CVD admiss. all ages
Toronto, Canada
Stieb et al. (2000) Total CVD emerg. dept.
Saint John, Canada visits, all ages
Atkinson et al. (1999b) Total emerg. CVD admiss.
Greater London, England > 65 yrs
Prescott et al. (1998) Total CVD admiss. > 65 yrs
Edinburgh, Scotland
Wong et al. (1999) Total emerg. CVD admiss.
> 65 yrs
28, IQR 22
14.0, 9.0
28.5, 90-10 %tile
range: 30.7
20.7, 8.4
Median 45.0,
IQR 34.8
none 1-4 day avg.
none 1-3 day avg.
none 0 day
none 1-3 day avg.
none 0-2 day avg.
7.7% (0.9, 14.8)
5. 9% (1.8, 10.2) PM2 *
13.5% (5. 5, 22.0)**
29.3% (p=0.003)
14.4% (p = 0.055) PM2 5*
2.5% (-0.2, 5.3)
12.4% (4.6, 20.9)
4.1% (1.3, 6.9)
Non-U.S. Results With Co-pollutants
        Burnett etal., (1997a)
        Toronto, Canada

        Stieb et al. (2000)
        Saint John, Canada

        Atkinson et al. (1999b)
        Greater London, England
        Prescott etal. (1998)
        Edinburgh, Scotland
        Wong etal. (1999)
Total CVD admiss. all ages   28, IQR 22
Total CVD emerg. dept.
visits, all ages
14.0, 9.0
Total emerg. CVD admiss.    28.5, 90-10 %tile
>65 yrs                    range: 30.7
Total CVD admiss. > 65 yrs   20.7, 8.4
Total emerg. CVD admiss.    Median 45.0,
> 65 yrs	IQR 34.8
                    O3, NO2, SO2, CO    1-4 day avg.
CO, H2S, NO2, O3,   1-3 day avg.
SO2, total reduced
sulfur
NO2, O3, SO2, CO   0 day
                   SO2, NO2, O3, CO    1-3 day avg.
                   NO2, O3, SO2
                   0-2 day avg.
-0.9% (-8.3, 7.1)
-1.1% (-7.8, 6.0)PM25*
 8.1% (-1.3, 18.3)**
PM10 not significant; no
quantitative results presented

PM10 not significant; no
quantitative results presented
PM10 effect robust; no
quantitative results presented
PM10 effect robust; no
quantitative results presented
       *PM2 5 entries. **PM10.2 s. All others relate to PM10.

-------
 1      Pittsburgh, PA; Provo/Orem, UT; Seattle, WA; Spokane, WA; and Youngstown, OH. The range
 2      of years studied encompassed 1985-1994, although this varied by city. Covariates included SO2,
 3      NO2, O3, and CO; however these were not analyzed directly as regression covariates. Individual
 4      cities were analyzed first by Poisson regression methods on PM10 for lags from 0 to  5 days.
 5      An overall PM10 risk estimate was then computed by taking the inverse-variance weighted mean
 6      of the city-specific risk estimates. The city-specific risk estimates for PM10 were also examined
 7      for correlations with omitted covariates, including other pollutants. No relationship was
 8      observed between city-specific risk estimates and measures of socioeconomic status, including
 9      percent living in poverty, percent non-white, and percent with college educations. The overall
10      weighted mean risk estimate for PM10 was greatest for lag 0 and for the mean of lags 0-1.
11      For example, the mean risk estimate for the mean of lags 0-1 was a 6.0% increase in CVD
12      admissions per 50 //g/m3 PM10 (95% CI: 5.1 - 6.8). The mean risk was larger in a subgroup of
13      data where PM10 was less than 50 //g/m3, suggesting the lack of a threshold.  A weakness of this
14      study was its failure to report multipollutant results.  The authors argued that confounding by
15      co-pollutants was not present because the city-specific risk estimates did not correlate with city-
16      specific regressions of PM10 on co-pollutant levels. However, the validity of this method for
17      identifying meaningful confounding by co-pollutants at the daily time-series level has not been
18      demonstrated. Thus, it is not possible to conclude from these results alone that the observed
19      PM10 associations were independent of co-pollutants.
20           Janssen et al. (2002), in further analyses of the data set examined above by Samet et al.
21      (2000a,b), evaluated whether differences in prevalence in air conditioning (AC) and/or the
22      contribution of different sources to total PM10 emissions could partially explain the observed
23      variability in exposure-effect relations in the 14 cities. Cities were characterized and analyzed as
24      either winter or nonwinter peaking for the AC analyses.  Data on the prevalence of AC from the
25      1993 American Housing Survey of the United States Census Bureau (1995) were used to
26      calculate the percentage of homes with central AC for each metropolitan area. Data on PM10
27      emissions by source category were obtained by county from the U.S. EPA emissions and air
28      quality data web site (2000). In an analysis of all 14 cities, central AC was not strongly
29      associated with PM10 coefficients. However, separate analysis for nonwinter peaking and winter
30      peaking PM10 cities yielded coefficients for CVD-related hospital admissions that decreased
31      significantly with increased percentage of central AC for both groups of cities, as  shown in

        April 2002                                8-113       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Figure 8-1 la.  Another plot shown in Figure 8-1 Ib depicts the relationships between PM10
 2      percent emissions from highways and CVD, showing significant positive relationships. For both
 3      analyses, similar patterns were found for hospitalization for COPD and pneumonia.  The authors
 4      note that the stronger relationship for hospital admission rates for CVD over COPD and
 5      pneumonia may relate to the 10 times higher CVD hospital admissions rate (which would result
 6      in less error).  However, no co-pollutant analyses were reported. The ecologic nature and limited
 7      sample size also indicate the need for further study.
 8           Zanobetti et al. (2000a) re-analyzed a subset of 10 cities from among the 14 evaluated by
 9      Samet et al. (2000a,b). The same basic pattern of results obtained by Samet et al. (2000a,b) were
10      found, with strongest PM10 associations on lag 0 day, smaller effects on lag 1 and 2, and none at
11      longer lags. The cross-city weighted mean estimate at 0 day lag was excess risk = 5.6% (95%
12      CI 4.7, 6.4) per 50 //g/m3 PM10 increment. The 0-1 day lag average excess CVD risk = 6.2%
13      (95% CI 5.4, 7.0) per 50 //g/m3 PM10 increment. Effect size estimates increased when data were
14      restricted to days with PM10 < 50 //g/m3. As before, no evidence of gaseous (CO, O3, SO2)
15      co-pollutant modification of PM  effects was seen in the second stage analyses. Again, however,
16      co-pollutants were not tested as independent explanatory variables in the regression analysis.
17           Turning to some examples  of independent single-city analyses, PM10 associations with
18      CVD hospitalizations were also examined in a study by Schwartz (1997), which analyzed three
19      years of daily data for Tucson, AZ linking total CVD hospital admissions for persons > 65 years
20      old with PM10, CO, O3, and NO2. As was the above case in Chicago, only one site monitored
21      daily PM10, whereas multiple sites did so for gaseous pollutants (O3, NO2,  CO). Both PM10 and
22      CO were independently (i.e., robustly) associated with CVD-related admissions, whereas O3 and
23      NO2 were not. The percent effect of a 50 //g/m3 increase in PM10 changed only slightly from 6.07
24      (CI 1.12-11.27) to 5.22 (CI 0.17 - 10.54) when CO was included in the model along with PM10.
25           Morris and Naumova (1998) reported results for PM10, as well  as for O3, NO2, and SO2 in
26      an analysis of four years of congestive heart failure data among people > 65 years old in Chicago,
27      IL. As many as eight monitoring  sites were available for calculating daily gaseous pollutant
28      concentrations; however, only one site in Chicago monitored daily PM10.  Only same-day results
29      were presented, based on an initial exploratory analysis showing strongest effects for same-day
30      pollution exposure (i.e., lag 0). Associations between hospitalizations and PM10 were observed
31      in univariate regressions (3.9% [1.0, 6.9] per 50 //g/m3 PM10 increase), but these diminished

        April 2002                                8-114       DRAFT-DO NOT QUOTE OR CITE

-------
                0.0025
                0.0020 -
           z:    0.0015 -
                0.0010 -
                0.0005 -
 c
 CD
'O
it
 CD
 O
O
Q

O
                0.0000
                                New Haven, CT
                                   * Ratio of Summer to
                                     Winter PM10 levels.
                         Boulder, CO
                          o1.35
                             Pittsburgh,
                                .63
                 Colorado Springs, CO
                     01.75
                -  ,        »-74
             Seattle WA^v   Youngstown, OH
                1'82    ^vs      Spokane WA
                         "^    »1.29
                                         Cahtoji, OH
                                         o   v
                                       Provo UT
                                         2.11
                            10     20      30      40      50
                                     Central Air Conditioning (%)
                                                          Nashville.TN
                                                             2.09
                                                                60
                                                                       70
                                                                              80
                0.0025
                0.0020 -
i=    0.0015 H
           Q)

          i£
           CD
          Q
          O
                0.0010 -
                0.0005 -
                0.0000
                               Boulder, CO
                 Colorado Springs, CO
•                           Youngstown,
                 „,,-.,„..»,
                          Birmingham, AL      ^ Seattle, WA

                                      Nashville, TN
                                                       New Haven, Ct.
                                            Chicago, IL
                                                                           B
                                                                     Detroit, Ml
                              123456
                                 PM10 emission from highway vehicles (%)
 Figure 8-11.  Univariate relation between percentage of homes with central AC and
               regression coefficients for (A) CVD, for cities nonwinter peaking PM10
               concentrations (solid line) and winter peaking PM10 concentrations
               (dashed line) and (B) univariate relation between percentage of PM10
               from highway vehicles and  regression coefficients for CVD. Circle area
               is proportional to the inverse of the variance of the effect estimate.  Lines
               represent inverse variance regression equations (fixed-effects model).

 Source: Adapted by EPA from Janssen et al. (2002).
April 2002
                                  8-115
                                                DRAFT-DO NOT QUOTE OR CITE

-------
 1      somewhat in a multi-pollutant model (2.0%, [-1.4, 5.4]). Strong, robust associations were seen
 2      between CO and congestive heart failure admissions.  These results seem to suggest a more
 3      robust association with CO than with PM10. However, the observed differences might also be
 4      due in part to differential exposure misclassification for PM10 (monitored at one site) as
 5      compared with CO (eight sites).
 6           In one of two U.S. studies comparing multiple PM indices, Lippmann et al. (2000)
 7      analyzed  associations between PM10, PM2 5, or PM10_2 5 and various categories of CVD hospital
 8      admissions  among the elderly (65+ yr) in Detroit on 490 days in the period 1992-1994. The most
 9      striking findings were notable percent excess risk for: (a) ischemic heart disease (IHD) in
10      relation to PM indices, i.e. 8.9% (0.5, 18.0) per 50 //g PM10; 10.5% (2.8, 18.9) per 25 //g/m3
11      PM10.25; and 4.3% (-1.4, 10.4) per 25 //g/m3 PM25 (all at lag 2d); and (b) heart failure, i.e. 9.7%
12      (0.2, 20.1) per 50 //g/m3 PM10; 5.2% (-3.3, 14.5) per 25 //g/m3 PM10.25; and 9.1% (2.4, 6.2) per
13      25 //g/m3 PM2 5 (the first two  at lag 0 d and the latter at lag 1 d).  The PM effects generally were
14      robust when co-pollutants were added to the model.  As discussed earlier with regard to the
15      Lippmann et al. (2000) mortality findings,  it is difficult to discern whether the observed
16      associations with coarse fraction particles (PM10_25) are independently due to such particles or
17      may possibly be attributed to  the moderately correlated fine particle (PM25) fraction in Detroit.
18      Also, power was limited by the  small sample size.
19           Tolbert et al. (2000a) reported very preliminary results on multiple PM indices as they
20      relate to daily hospital emergency department visits for dysrhythmias (DYS) and all CVD
21      categories for persons aged 16 yrs or older, based on analyses of data from 18 of 33 participating
22      hospitals  in an ongoing study in Atlanta. During Period 1 of the study (1993-1998), PM10 from
23      the EPA AIRS database was reported to be negatively associated with CVD visits. In a
24      subsequent  one-year period (Aug. 1998 - Aug. 1999), when data became available from the
25      Atlanta PM supersite, positive but non-significant associations were seen between CVD and
26      PM10 (RR of 5.1% per 50 //g/m3 PM10) and PM2.5 (RR of 6.1% per 25 //g/m3 PM2 5); and
27      significant positive associations were seen with certain fine particle components, i.e., elemental
28      carbon (p < 0.005) and organic carbon (p < 0.02), along with CO (p < 0.005).  No multi-pollutant
29      results were reported. Study power was limited due to the short data record in Period 2.
30      In addition, caution applies to acceptance of the Tolbert et al. (2000a) findings until more
31      complete analyses from all participating hospitals are carried out and reported.

        April 2002                                8-116        DRAFT-DO NOT QUOTE OR CITE

-------
 1           In an analysis of 1992-1995 Los Angeles data, Linn et al. (2000) also found that PM10, CO,
 2      and NO2 were all significantly associated with increased cardiovascular admission in single-
 3      pollutant models among persons aged 30 yr and older.  Associations generally appeared to be
 4      stronger for CO than for PM10. No PM10 results were presented with co-pollutants in the model.
 5           Lastly, Moolgavkar (2000b) analyzed PM10, CO, NO2, O3, and SO2 in relation to daily total
 6      cardiovascular (CVD) and total cerebrovascular (CrD) admissions for persons aged >65 from
 7      three urban counties (Cook, IL; Los Angeles, CA; Maricopa, AZ) in the period 1987-1995.
 8      Consistent with most  studies, in univariate regressions, PM10 (and PM2 5 in LA) was associated at
 9      some lags with CVD admissions in Cook and LA counties, but not in Maricopa county.
10      However, in two-pollutant models in Cook and LA counties, the PM risk estimates diminished
11      substantially and/or were rendered non-significant, whereas co-pollutant (CO or NO2) risk
12      estimates were less affected.  Results of this study suggest that gaseous pollutants, with the
13      exception of 03, were  more strongly associated with CVD hospitalizations than was PM.
14           The above analyses of daily PM10 and CO in U.S. cities, overall, indicate that elevated
15      concentrations of both PM10 and CO may enhance risk of CVD-related morbidity leading to acute
16      hospitalizations. The Lippmann results appear to implicate PM25 and/or PM10_25 in increased
17      hospital admissions for some categories of CVD among the elderly.
18
19      8.3.1.3.2  Studies in Non-U.S. Cities
20           Four separate analyses of hospitalization data in Canada have been reported by Burnett and
21      coworkers since 1995 (Burnett et al., 1995, 1997a,b, 1999). A variety of locations, outcomes,
22      PM exposure metrics, and analytical approaches were used in these studies, which hinders
23      somewhat the ability to draw broad conclusions across the full group.  The first (Burnett et al.,
24      1995), reviewed briefly in the 1996 PM AQCD, analyzed six years of data from 168 hospitals in
25      Ontario, CN.  Cardiovascular (CVD) and respiratory hospital admissions were analyzed in
26      relation to sulfate and ozone concentrations. Sulfate lagged one day was associated with CVD
27      admissions, with a percent effect of 2.8 (CI 1.8-3.8) per 13  //g/m3 without O3 in the model and
28      3.3 (CI 1.7-4.8) with O3 included. When CVD admissions  were split out into  sub-categories,
29      larger associations were seen between sulfates and coronary artery disease and heart failure than
30      for cardiac dysrhythmias.  Sulfate associations with total admissions were larger for the elderly
        April 2002                                8-117        DRAFT-DO NOT QUOTE OR CITE

-------
 1      sub-population > 65 yr old (3.5% per 13 //g/m3) than for those <65 yr old (2.5% per 13 //g/m3).
 2      There was little evidence for seasonal differences in sulfate associations.
 3           Burnett et al. (1997c) analyzed daily congestive heart failure hospitalizations in relation to
 4      carbon monoxide and other air pollutants (O3, NO2, SO2, CoH) in ten large Canadian cities as a
 5      replication of an earlier U.S.  study by Morris et al. (1995).  The Burnett Canadian study
 6      expanded upon the previous work both by its size (11 years of data for each of 10 large cities)
 7      and also by including a measure of PM air pollution (coefficient of haze, CoH), whereas no PM
 8      data were included in the earlier Morris et al. study.  The Burnett study was restricted to the
 9      population > 65 years old.  The authors noted that all pollutants except O3 were correlated,
10      making it difficult to separate them statistically.  CoH, CO, and NO2 measured on the same day
11      as admission (i.e., lag 0) were all strongly associated with congestive heart failure admissions in
12      univariate models.  In multi-pollutant models, CO remained a strong predictor, whereas COH did
13      not (gravimetric PM measures were not evaluated).
14           The roles played by size-selected gravimetric and chemically speciated particle metrics as
15      predictors of CVD hospitalizations were explored in  analysis of data from metropolitan Toronto
16      for the summers of 1992-1994 (Burnett et al., 1997a). The analysis used  dichotomous sampler
17      (PM25, PM10, and PM10_25), hydrogen ion, and sulfate data  collected at  a central site as well as O3,
18      NO2, SO2, CO, and COH data collected at multiple sites in Toronto. Hospital  admissions
19      categories included total cardiovascular (i.e., the sum of ischemic heart disease, cardiac
20      dysrhythmias, and heart failure) and total respiratory. Model specification with respect to
21      pollution lags was completely data-driven, with all lags and averaging  times out to 4 days  prior to
22      admission evaluated in exploratory analyses and "best" metrics chosen on the  basis of maximal
23      t-statistics. The relative risks of CVD admissions were positive and generally statistically
24      significant for all pollutants analyzed in univariate regressions, but especially so for O3, NO2,
25      COH, and PM10_2 5 (i.e., regression t-statistics > 3). Associations for gaseous pollutants were
26      generally robust to inclusion of PM covariates, whereas the PM indices (aside from COH) were
27      not robust to inclusion of multiple gaseous pollutants. In particular, PM2 5 was not a robust
28      predictor of CVD admissions in multi-pollutant models: whereas an 25 //g/m3 increase in PM25
29      was associated with a 5.9% increase (t=l.8) in CVD admissions in a univariate model, the
30      percent effect was reduced to -1.1 (t=0.3) in a model that included O3, NO2, and SO2. COH, like
31      CO and NO2, is generally thought of as a measure of primary motor-vehicle emissions during the

        April 2002                                8-118        DRAFT-DO NOT QUOTE OR  CITE

-------
 1      non-heating season. The authors concluded that "particle mass and chemistry could not be
 2      identified as an independent risk factor for exacerbation of cardiorespiratory diseases in this
 3      study beyond that attributable to climate and gaseous air pollution."
 4           Burnett et al. (1999) later reported results of a more extensive attempt to explore cause-
 5      specific hospitalizations for persons of all ages in relation to a large suite of gaseous and PM air
 6      pollutant measures, using 15  years of Toronto data.  Cardiovascular admissions were split out
 7      into separate categories for analysis: dysrhythmias, heart failure, and ischemic heart disease.
 8      The analyses also examined several respiratory causes, as well as cerebrovascular and diseases of
 9      the peripheral circulation (the latter categories being included because they should show PM
10      associations if one mechanism of PM action is related to increased plasma viscosity, as suggested
11      by Peters et al. (1997a).  The PM metrics analyzed were PM2 5, PM10, and PM10_25 estimated from
12      daily TSP and  TSP sulfate data, based on a regression analysis on dichotomous sampling data
13      that were available every sixth day during an eight-year subset of the full study period.  This use
14      of estimated rather than measured PM components limits the interpretation of the PM results
15      reported here.  In general, use of estimated PM exposure metrics will tend to increase exposure
16      measurement error and thereby tend to decrease effects estimates.  Model specification for lags
17      was again data-driven, based on maximal t-statistics. Although  some statistically significant
18      associations with one or another PM metric were found in  univariate models, there were no
19      significant PM associations with any of the three CVD hospitalization outcomes in multi-
20      pollutant models.  For example, whereas an 25 //g/m3 increase in estimated PM2 5 was associated
21      with a 8.05% increase (t-statistic = 6.08) in ischemic heart disease admissions in a univariate
22      analysis, the PM25 association was reduced to 2.25% (n.s.) when NO2 and SO2 were included in
23      the model.  The gaseous pollutants dominated most regressions. There also were no associations
24      between PM and cerebral or peripheral vascular disease admissions.
25           The Burnett et al. studies provide some of the most extensive results for PM in conjunction
26      with multiple gaseous pollutants, but the inconsistent use of alternative PM metrics in the various
27      analyses confuses the picture somewhat.  A general finding appears to be lack of robustness of
28      associations between cardiovascular outcomes and PM in multi-pollutant analyses. This was
29      seen for COH in the analysis of 10 Canadian  cities (Burnett et al.,  1997c), for PM2 5 and PM10 in
30      the analysis of summer data in Toronto (Burnett et al., 1997a), and for linear combinations of
31      TSP and sulfates (i.e., estimated PM25, PM10, and PM10_25) in the analysis of 15 years of data in

        April 2002                                 8-119        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Toronto (Burnett et al., 1999). One exception was the association reported between CVD
 2      admissions to 168 Ontario hospitals and sulfate concentrations (Burnett et al., 1995), where the
 3      sulfate association was robust to the inclusion of O3. Also, although gravimetric PM variables
 4      were not robust predictors in the Toronto summer analysis, COH was (Burnett et al., 1997a),
 5      perhaps  reflecting the impact of primary motor vehicle emissions. This contrasts, however, with
 6      COH's lack of robustness in the 10-city analysis (Burnett et al., 1997c).
 7           Stieb et al. studied all-age acute cardiac emergency room visits in relation to a rich set of
 8      pollution covariates in Saint John, Canada for the period 1992-1996.  Daily data were available
 9      on PM2 5, PM10, fine fraction hydrogen and sulfate ions, COH, CO, H2S, NO2, O3, SO2, and total
10      reduced  sulfur. In a multi-pollutant model, neither PM10 nor PM2 5 were significantly related to
11      total cardiac ED visits, though O3 and SO2 were.
12           Several additional non-U.S. studies, mainly in the U.K., have also been published since the
13      1996 PM AQCD.  Most of these studies evaluated co-pollutant effects along with those of PM.
14      Interpretation is hindered somewhat, however, by the failure to report quantitative results for
15      PM10 in  the presence of co-pollutants.  In univariate models, Atkinson et al. (1999a) reported
16      significant associations of both ambient PM10 and black smoke (BS), as well  as all other
17      co-pollutants, with daily admissions for total cardiovascular disease and ischemic heart disease
18      for 1992-1994 in London, UK, using standard time series regression Methods. Co-pollutants
19      included NO2, O3, SO2, and CO. PM associations were observed for persons aged < 65 yr and for
20      persons  aged > 65 yr. In two-pollutant  models, the associations with PM10, NO2, SO2, and CO
21      were moderated by the presence of BS in the model, but the BS association was robust to
22      co-pollutants. Interpretation is hampered somewhat by the lack of quantitative results for
23      two-pollutant models. In another U.K.  study, associations with PM10, and to  a lesser extent BS,
24      SO2, and CO, were reported for analyses of daily emergency hospital admissions for
25      cardiovascular diseases from 1992-1995 for Edinburgh, UK (Prescott et al., 1998).
26      No associations were observed for NO2 and ozone. Significant PM10 associations were present
27      only in persons 65 and older. The authors reported that the PM10 associations were unaffected by
28      inclusion of other pollutants; however, results were not shown.  On the other hand, no
29      associations between PM10 and daily ischemic heart disease admissions were observed by
30      Wordley and colleagues (1997) in an analysis of two years of daily data from Birmingham, UK.
31      However, PM10 was associated with respiratory admissions and cardiovascular mortality during

        April 2002                               8-120        DRAFT-DO NOT QUOTE OR CITE

-------
 1      the same study period. This inconsistency of results across causes and outcomes is difficult to
 2      interpret, but may relate in part to the relatively short time series analyzed. The authors stated
 3      that gaseous pollutants did not have significant associations with health outcomes independent of
 4      PM, but no results were presented for models involving gaseous pollutants.
 5           In eight European cities, the APHEAII (Le Tertre et al., 2002) project examined the
 6      association between PM10 and hospital admissions for cardiac causes.  They found a significant
 7      effect of PM10 (0.5%; 0.2, 0.8) on admission for cardiac causes (all ages) and cardiac causes
 8      (0.7%;  0.4, 1.0) and ischemic heart disease (0.8%; 0.3, 1.2) for people over 65 years with the
 9      impact of PM10 per unit of pollution being half that found in the United States.  PM10 did not
10      seem to be confounded by O3 and SO2.  The PM10 effect was reduced when CO was incorporated
11      in the regression model and eliminated when controlling for NO2.
12           A study in Hong Kong by Wong et al (1999) found associations between CVD admissions
13      and PM10,  SO2, NO2, and O3 in univariate models, but did not examine multi-pollutant models.
14      Ye and colleagues analyzed a 16 year record of daily emergency hospital visits for July and
15      August in Tokyo among persons age 65 and  older (Ye et al., 2001).  In addition to PM10, the
16      study included NO2, ozone, SO2, and CO.  Models were built using an objective significance
17      criterion for variable inclusion. NO2 was the only pollutant significantly associated with angina,
18      cardiac insufficiency, and myocardial infarction hospital visits.
19
20      8.3.1.3.3 Summary and Conclusions
21           The ecologic time series studies reviewed here add substantially to the body of available
22      literature on acute CVD morbidity effects of PM and co-pollutants.  Two U.S. multi-city studies
23      offer the strongest current evidence for effects of PM10 on acute CVD hospital admissions.
24      However, uncertainties regarding the possible role of co-pollutants in the larger of the two
25      studies hinders interpretation with respect to independent PM10 effects. Among single-city
26      studies carried out in the U.S. and elsewhere by a variety of investigators (see Summary
27      Table 8-17), less consistent evidence for PM effects is seen. Of particular importance is the
28      possible roles of co-pollutants (e.g., CO) as confounders of the PM effect.  Among
29      13 independent studies that included gravimetrically-measured PM10 and co-pollutants, three
30      reported PM effects that appeared independent of co-pollutants (Schwartz, 1997; Lipmmann
31      et al., 2000; Prescott et al., 1998), eight reported no significant PM10 effects after inclusion of

        April 2002                               8-121        DRAFT-DO NOT QUOTE OR CITE

-------
 1      co-pollutants (Morris and Naumova, 1998; Moolgavkar, 2000b; Tolbert et al., 2000a; Burnett
 2      et al., 1997a; Steib et al., 2000; Atkinson et al., 1999b; Wordley et al. (1997); Morgan et al.,
 3      1998; Ye et al., 2001), and two studies were unclear regarding independent PM effects (Linn
 4      et al., 2000; Wong et al., 1999). In a recent quantitative review of published results from
 5      12 studies on airborne particles and hospital admissions for cardiovascular disease, Morris
 6      (2001) noted that adjustment for co-pollutants consistently reduced the PM10 effect, with
 7      reductions ranging from 10 to 320% across studies.  Thus, although several studies appear to
 8      provide evidence for PM effects on CVD hospital admissions independent of co-pollutant
 9      effects, still other studies examining co-pollutants yield results, showing PM effects in some
10      studies while not in others.
11           With respect to the question of particle size, only a handful of studies have examined the
12      relative impacts of different particle indicators (Lippmann et al., 2000, Burnett et al., 1997a,
13      Tolbert et al., 2000a, Steib et al., 2000, Moolgavkar, 2000b).  Perhaps due to statistical power
14      issues, no clear picture has emerged as to the particle size fraction most associated with acute
15      CVD effects.
16           Because hospitalization can be viewed as a less severe manifestation of the same
17      pathophysiologic mechanism that may be responsible for acute mortality following PM exposure,
18      it is of interest to assess the coherence between the morbidity results reviewed here and the
19      mortality results reviewed in Section 8.2.2 (Borja-Aburto et al., 1997, 1998; Braga et al., 2001;
20      Goldberg et al., 2000; Gouveia and Fletcher, 2001; Hoek et al., 2001; Kwon et al., 2001;
21      Michelozzi et al.,  1998; Morgan et al., 1998; Ponka et al., 1998; Schwartz et al., 1996a; Simpson
22      et al., 1997; Wordley et al., 1997; Zeghnoun et al., 2001; Zmirou et al., 1998). The mortality
23      studies reported significant associations between acute CVD mortality and measures of ambient
24      PM, though the PM metrics utilized and the  relative risk estimates varied  across studies. PM
25      measurement methods included gravimetrically analyzed filter samples (TSP, PM10, PM2 5,
26      PM10_25), beta gauge (particle attenuation of beta radiation), nephelometry (light scattering), and
27      black smoke (filter reflectance). Where tested, PM associations with acute CVD mortality
28      appeared to be generally more robust to inclusion of gaseous covariates than was the case for
29      acute hospitalization studies (Borja-Aburto et al., 1997,  1998; Morgan et al., 1998; Wordley
30      et al., 1997; Zmirou et al.,  1998). One study (Goldberg et al., 2000) which examined multiple
31      alternative PM metrics reported strongest associations with PM2 5 and no associations for PM10_2 5

        April 2002                                8-122        DRAFT-DO NOT QUOTE OR CITE

-------
 1      and hydrogen ion.  Three studies (Braga et al., 2001; Goldberg et al., 2000; Hoek et al., 2001),
 2      as noted in Section 8.2.2, provide data indicating that some specific cardiovascular causes of
 3      mortality (such as heart failure) were more strongly associated with air pollution than total
 4      cardiovascular mortality; but it was noted that ischemic heart disease (which contributes about
 5      half of all CVD deaths) was the strongest contribution to the association between air pollution
 6      and cardiovascular mortality.  These results for acute cardiovascular mortality are qualitatively
 7      consistent with those reviewed above for hospital admissions.
 8           Figure 8-12 illustrates PM10 excess risk estimates for single-pollutant models derived from
 9      selected U.S. studies of PM10 exposure and total cardiovascular disease (CVD) hospital
10      admissions, standardized to a 50 //g/m3 exposure to PM10.  Results are shown both for studies
11      yielding pooled outcomes for multiple U.S. cities and for studies of single U.S. cities.  The Samet
12      et al. (2000a) pooled cross-city results for 14 U.S. cities provides the most precise estimate for
13      relationships of U.S. ambient PM10 exposure to increased risk for CVD hospitalization. That
14      estimate, and those derived from most other studies depicted in Figure 8-6, generally appear to
15      confirm likely excess risk of CVD-related hospital admissions for U.S. cities in the range of
16      3-10% per 50 //g/m3 PM10, especially among the elderly (>65 yr).  Also, other individual-city
17      results from Detroit are indicative of excess risk for ischemic heart disease and heart failure in
18      the range of approximately 4.0 to 10.0% per 25 //g/m3 of PM25 or PM10_25, as are preliminary
19      individual-city findings from Atlanta suggestive of 4.3% and  10.5% excess risk per 25 //g/m3 of
20      PM2 5 and PM10_2 5, respectively. However, the extent to which PM affects CVD hospitalization
21      risk independently of or together with other co-pollutants (such as CO), remains to be further
22      resolved.
23
24      8.3.1.3.4  Individual-Level Studies of Cardiovascular Physiology
25           New studies carried out by various groups have evaluated longitudinal associations
26      between ambient PM and physiologic measures of cardiovascular function or biochemical
27      changes in the blood that may be associated with cardiac risks.  In  contrast to the ecologic time-
28      series studies discussed above, these studies measure outcomes and most covariates at the
29      individual level, making it possible to draw conclusions regarding individual risks, as well  as to
30      explore mechanistic hypotheses. Heterogeneity of responses across individuals, and across
31      subgroups defined on the basis of age, sex, pre-existing health status, etc., also can in principle

        April 2002                                8-123        DRAFT-DO NOT QUOTE OR CITE

-------
             Samet et al. (2000)
                14US Cities
               Schwartz (1999)
                 8 US Cities
             Moolgavkar (2000b)
               Maricopa, AZ

             Moolgavkar (2000b)
                  LA.CA
             Moolgavkar (2000c)
               Cook County

               Linn et al. (2000)
                   LA.CA

               Schwartz (1997)
                 Tucson,AZ

            Tolbert et al. (2000a)
                 Atlanta
       Morris and Naumova (1998)
              Chicago

           Lippmann etal.(2000) -
Total CVD
       Period 1 (AIRS Data)
 I             «	1
                                                            CHF
                           HF
                           IHD
                                  i	»	1
                                  Period 2 (Supersite Data)
                                 I      «	1
                                 -15       -10        -5         0          5        10
                                        Reconstructed  Excess Risk Percentage
                                                 50 i^g/m3 Increase in
      Figure 8-12.  Acute cardiovascular hospitalizations and particulate matter exposure excess
                    risk estimates derived from selected U.S. PM10 studies.  CVD =
                    cardiovascular disease.  CHF = congestive heart failure.
1      be assessed. While exposure assessment remains largely ecologic (i.e., the entire population is

2      usually assigned the same exposure value on a given day), exposure is generally well

3      characterized in the small, spatially-clustered study populations.  The recent studies fall into two

4      broad classes: those addressing cardiac rhythm or adverse events, and those addressing blood

5      characteristics.  While significant uncertainty still exists regarding the interpretation of results

6      from these new studies, the varied responses that have been reported to be associated with

7      ambient PM and co-pollutants are of much interest in regard to mechanistic hypotheses
       April 2002
                8-124
DRAFT-DO NOT QUOTE OR CITE

-------
 1      concerning pathophysiologic processes potentially underlying CVD-related mortality/morbidity
 2      effects discussed in preceding sections.
 3
 4      Cardiac Physiology and Adverse Cardiac Events
 5           Alterations in heart rate and/or rhythm have been hypothesized as possible mechanisms by
 6      which ambient PM exposures may exert acute effects on human health. Decreased heart rate
 7      variability, in particular, has been identified as a predictor of increased cardiovascular morbidity
 8      and mortality. Several independent studies have recently reported temporal associations between
 9      PM exposures and various measures of heart beat rhythm in panels of elderly subjects (Liao
10      et al.,  1999; Pope et al., 1999a,b,c; Dockery et al., 1999; Peters et al., 1999a, 2000a; Gold et al.
11      2000;  Creason et al., 2001).  Changes in blood pressure may also reflect increases in risk (Linn
12      et al.,  1999; Ibald-Mulli et al., 2001). Finally, one important new study has linked acute (2- and
13      24-h) ambient PM25 and PM10 concentrations with increased risk of myocardial infarction in
14      subsequent hours and days (Peters et al., 2001).
15           Liao and colleagues (1999) studied 26 elderly  subjects (age 65-89 years; 73% female) over
16      three consecutive weeks at a retirement center in metropolitan Baltimore, 18 of whom were
17      classified as "compromised" based on previous cardiovascular conditions (e.g., hypertension).
18      Daily six-minute resting electrocardiogram (ECG) data were collected, and time intervals
19      between sequential R-R intervals recorded. A Fourier transform was applied to the R-R interval
20      data to separate its variance into two major components: low frequency (LF, 0.04-0.15 Hz) and
21      high frequency (FTP, 0.15-0.40 Hz).  The standard deviation of all normal-to-normal (N-N; also
22      designated R-R) heartbeat intervals (SDNN) was computed for use as a time-domain outcome
23      variable.  PM2 5 was monitored indoors by TEOM and outdoors by dichotomous sampler.
24      Outdoor PM25 levels ranged from 8.0 to 32.2 //g/m3 (mean = 16.1 //g/m3). Regression analyses
25      controlled for inter-subject differences in  average variability, allowing each subject to serve as
26      his/her own control. Consistent associations were seen between decreases in all three outcome
27      variables  (LF, FTP, SDNN) and increases in PM2 5 concentrations (both indoors and outdoors),
28      with associations being stronger for the 18 "compromised"subjects. No analyses of heart rate
29      were reported.
30           Creason and colleagues (2001) recently reported results of a subsequent study using similar
31      methods among 56 elderly residents of a retirement  center in Baltimore County, MD. The

        April 2002                                8-125       DRAFT-DO NOT QUOTE OR CITE

-------
 1      11 men and 45 women ranged in age from 72 to 97 years and were all Caucasian. Associations
 2      between decreased HRV and ambient PM2 5 were again observed, though not significant at the
 3      0.05 level and smaller in magnitude than in the previous Baltimore study.  When two episodic
 4      PM2 5 days with rainfall were excluded from the 24-day data set, the PM2 5 associations increased
 5      in magnitude and became statistically significant. There was no evidence of larger effects among
 6      subsets of subjects with compromised health status. No results were presented for other
 7      pollutants besides PM2 5.
 8           Pope and colleagues (1999c) reported  similar findings in a panel of six elderly subjects
 9      (69-89 years, 5/6 male) with histories of cardiopulmonary disease, and one 23-year old male
10      subject suffering from Crohn's disease and arrhythmias.  Subjects carried Hotter monitors for up
11      to 48 hours during different weeks that varied in ambient PM10 concentrations. N-N heartbeat
12      intervals were recorded and used to calculate several measures of heart rate variability in the time
13      domain: the standard deviation of N-N intervals (SDNN), which is a broad measure of both high
14      and low frequency variations; the standard deviation of the  averages of N-N intervals in  all five
15      minute segments (SDANN), which is a measure of ultra-low frequency variations; and the root
16      mean squared differences between adjacent N-N intervals (r-MSSD), which is a measure of high
17      frequency variations.  Daily gravimetric PM10 data obtained from three sites in the study area
18      ranged from circa 10 //g/m3 to 130 //g/m3 during the study.  A simple step function in
19      concentration was observed with high levels occurring only during the first half of the 1.5 month
20      study period. Regression analysis with  subject-specific intercepts was performed, with and
21      without control for daily barometric pressure and mean heart rate.  Same-day, previous-day, and
22      the two-day mean of PM10 were considered. SDNN and SDANN were negatively associated with
23      both same-day and previous-day ambient PM10, and results  were unaffected by inclusion of
24      covariates. Heart rate, as well as r-MSSD, were both positively, but less strongly, associated
25      with PM10. No co-pollutants were studied.
26           The Pope et al. (1999c) study discussed above was nested within a larger cohort of
27      90 subjects who participated in a study of heart rate and oxygen saturation in the Utah Valley
28      (Dockery et al.,  1999; Pope et al., 1999b). The investigators hypothesized that decreases in
29      oxygen saturation might occur as a result of PM exposure, and that this could be a risk factor for
30      adverse cardiac outcomes. The study was carried out in winter months (mid November through
31      mid-March), when frequent inversions lead to fine particle episodes. PM10 levels at the three

        April 2002                                8-126        DRAFT-DO NOT QUOTE OR CITE

-------
 1      nearest sites averaged from 35 to 43 //g/m3 during the study, with daily 24-h levels ranging from
 2      5 to 147//g/m3. Two populations were studied: 52 retired Brigham Young University
 3      faculty/staff and their spouses, and 38 retirement home residents. Oxygen saturation (SpO2) and
 4      heart rate (HR) were measured once or twice daily by an optical sensor applied to a finger.
 5      In regression analyses that controlled for inter-individual differences in mean levels,  SpO2 was
 6      not associated with PM10, but was highly associated with barometric pressure. In contrast, HR
 7      significantly increased in association with PM10 and significantly decreased in association with
 8      barometric pressure in joint regressions. Including CO in the regressions did not change these
 9      basic findings.  This was the first study of this type to examine the interrelationships  among
10      physiologic measures (i.e.,  SpO2 and HR), barometric pressure, and PM10. The profound
11      physiological effects of barometric pressure noted here highlight the importance of carefully
12      controlling for barometric pressure effects in studies of cardiac physiology.
13           Gold and colleagues (2000) obtained somewhat different results in a study of heart rate
14      variability among 21 active elderly subjects, aged 53-87 yr, in a Boston residential community.
15      Resting, standing, exercising, and recovering ECG measurements were performed weekly using a
16      standardized protocol on each subject, which involved 25 min/week of continuous Hotter ECG
17      monitoring. Two time-domain measures were extracted: SDNN and r-MSSD (see above for
18      definitions). Heart rate also was analyzed as an outcome.  Continuous PM10 and PM25
19      monitoring was conducted by TEOM at a site 6 km from the study site, with PM data corrected
20      for loss of semivolatile mass. Data on CO, O3, NO2, SO2, temperature and relative humidity
21      were available from nearby sites.  Outcomes were regressed on PM2 5 levels in the 0-24 hour
22      period prior to ECG testing, with and without control  for HR and temperature.  As for the other
23      studies discussed above, declines in SDNN were associated with PM2 5 levels, in this case
24      averaged over 4 hours. These associations reached statistical significance at the 0.05 level only
25      when all testing periods (i.e., resting, standing,  exercise) were combined. In contrast to the above
26      studies, both HR and r-MSSD here were negatively associated with PM2 5 levels (i.e., lower HR
27      and r-MSSD) when PM2 5 was elevated.  These associations were statistically significant overall,
28      as well as for several of the individual testing periods, and were unaffected by covariate control.
29           Peters and colleagues (1999a) reported HR results from a retrospective analysis of data
30      collected as part of the MONICA study (monitoring of trends and determinants in cardiovascular
31      disease) carried out in Augsburg, Germany. Analyses focused on 2,681 men and women aged

        April 2002                                8-127        DRAFT-DO NOT QUOTE OR CITE

-------
 1      25-64 years who had valid ECG measurements taken in winter 1984-1985 and again in winter
 2      1987-1988. Ambient pollution variables included TSP, SO2, and CO. The earlier winter included
 3      a 10-day episode with unusually high levels of SO2 and TSP, but not of CO. Pollution effects
 4      were analyzed in two ways: dichotomously comparing the episode and non-episode periods, and
 5      continuously using regression analysis.  However, it is unclear from the report to what extent the
 6      analyses reflect between-subject vs. within-subject effects.  A statistically significant increase in
 7      mean heart rate was observed during the episode period versus other periods, controlling for
 8      cardiovascular risk factors and meteorology. Larger effects were observed in women.  In single-
 9      pollutant regression analyses, all three pollutants were associated with increased HR.
10           In another retrospective study, Peters and colleagues (2000a) examined incidence of cardiac
11      arrhythmias among 100 patients (mean age 62.2 yr; 79% male) with implanted cardioverter
12      defibrillators followed over a three year period. PM2 5 and PM10 were measured in  South Boston
13      by the TEOM method, along with black carbon, O3, CO, temperature and relative humidity; SO2
14      and NO2 data were obtained from  another site. The 5th percentile, mean, and 95th  percentiles of
15      PM10 concentrations were 7.8, 19.3, and 37.0 //g/m3, respectively.  The corresponding values for
16      PM25 were 4.6, 12.7, and 26.6 //g/m3. Logistic regression was used to analyze arrhythmia events
17      in relation to pollution variables, controlling for between-person differences, seasons, day-of-
18      week, and meteorology in two subgroups: 33 subjects with at least one arrhythmia event; and
19      6 subjects with 10 or more arrhythmia events. In the larger subgroup, only NO2 on the previous
20      day, and the mean NO2 over five days, were significantly associated with arrhythmia incidence.
21      In patients with 10 or more events, the NO2 associations were stronger. Also, some of the PM25
22      and CO lags became significant in this subgroup. These results should be interpreted cautiously
23      given the large number of statistical tests performed.
24           Linn and colleagues (1999) reported associations between both diastolic and systolic blood
25      pressure and PM10 in a panel study of 30 Los Angeles residents with severe COPD.  Recently,
26      Ibald-Mulli et al. (2001) reported similar findings from a study of blood pressure among 2607
27      men and women aged 25-64 years who participated in the MONICA study in Augsburg,
28      Germany.  Systolic blood pressure increased on average during an episode of elevated TSP and
29      SO2, but the effect disappeared after controlling for meteorological parameters including
30      temperature and barometric pressure.  However, when TSP and SO2 were analyzed as continuous
31      variables, both were associated with elevated systolic blood pressure, controlling for

        April 2002                               8-128        DRAFT-DO NOT QUOTE OR CITE

-------
 1      meteorological variables. In two-pollutant models, TSP was more robust than SO2. Further, the
 2      TSP association was greater in the subgroups of subjects with elevated blood viscosity and heart
 3      rates.
 4           An exploratory study of a panel of COPD patients (Brauer et al., 2001) examined several
 5      PM indicators in relation to CVD and respiratory health effects. The very low levels of ambient
 6      particle (PM10 mean - 18[7] //g/m3) and low variability in these levels plus the sample size of 16
 7      limit the conclusions that can be drawn. Nevertheless, for cardiovascular endpoints, single-
 8      pollutant models indicated that both systolic and diastolic BP decreased with increasing
 9      exposure, but this is not statistically significant. Also, the size of the ambient PM10 effect
10      estimate for AFEVj was larger than the effect estimate for ambient PM25 and personal PM2 5 but
11      not statistically significant.  While the quantitative health relationships outcome results are
12      inconclusive, the results related to PM indicators is informative while requiring future research.
13      This initial effort indicates that ambient PM10 consistently had the largest effect estimates while
14      models using personal exposure measurements did not show larger or more consistently positive
15      effect estimates relative to those using ambient exposure metrics.
16           An important study by Peters et al. (2001) reported associations between onset of
17      myocardial infarction and ambient PM (either PM10 or PM2 5) in a cohort of 772 MI patients
18      studied in Boston, MA  as part of the determinants of myocardial infarction onset study.  Precise
19      information on the timing of the MI, obtained from patient interviews, was linked with
20      concurrent air quality data measured at a single Boston site.  A case crossover design enabled
21      each Subject to serve as his/her own control.  One strength of this study was its analysis of
22      multiple PM indices and co-pollutants, including real-time PM2 5, PM10, the PM10-PM2 5
23      difference, black carbon, Ozone, CO, NO2, and SO2.  Only PM2 5 and PM10 were significantly
24      associated with MI risk in models adjusting for season, meteorological parameters, and day of
25      week. Both the mean PM2 5 concentration in the previous two hours and in the 24 hours lagged
26      one day were independently associated with MI, with odds ratios of 1.48 (1.09-2.02) for 25
27      ug/m3 and 1.62 (1.13-2.34) for 20 ug/m3,  respectively.  PM10 associations were similar.  The
28      non-significant findings for other pollution metrics should be interpreted in the context of
29      potentially differing exposure misclassification errors  associated with the single monitoring site.
30           The above studies present a range of intriguing findings suggesting possible effects of PM
31      on cardiac rhythm and adverse events. Four separate studies reported decreases in HR variability

        April 2002                                8-129       DRAFT-DO NOT QUOTE OR CITE

-------
 1      associated with PM in elderly cohorts, although r-MSSD (a measure of high-frequency HR
 2      variability) showed elevations with PM in one study (Pope et al., 1999a).  Also, all of the studies
 3      which examined HR found an association with PM; most reported positive associations, whereas
 4      one (Gold et al., 2000) reported a negative relationship. However, variations in methods and
 5      results across the studies argue for caution in drawing strong conclusions regarding PM effects
 6      from them, especially in light of the complex intercorrelations which exist among measures of
 7      cardiac physiology, meteorology, and air pollution (Dockery et al., 1999).
 8
 9      Viscosity and Other Blood Characteristics
10           Peters et al. (1997a) state that plasma viscosity is determined by fibrinogen  and other large
11      asymmetrical plasma proteins such as immunoglobulin M and K2-macr°gl°bulin.  They note that
12      in a cohort study of elderly men and women, fibrinogen concentrations were strongly related to
13      inflammatory markers such as neutrophil  count and acute-phase proteins, (C-reactive protein and
14      Kj-antichymotrypsin) and to self-reported infections. Fibrinogen contributes to plasma viscosity,
15      which is a risk factor for ischemic heart disease.
16           Support for a mechanistic hypothesis,  relating to  enhanced blood viscosity,  is suggested in
17      an analysis of plasma viscosity data collected in a population of 3256 German adults in the
18      MONICA study (Peters et al., 1997a).  Each subject provided one blood sample during October
19      1984 to June 1985. An episode of unusually high air pollution concentrations occurred during a
20      13 day period while these measurements were being collected. The authors reported that, among
21      the 324 persons who provided blood during the episode, there was a statistically significant
22      elevation in plasma viscosity as compared with the 2932 persons studied at other  times. The
23      odds ratio for plasma viscosity exceeding the 95th percentile was 3.6 (CI 1.6-8.1) among men
24      and 2.3 (CI 1.0-5.3) among women. Analysis of the distribution of blood viscosity data
25      suggested that these findings were driven by changes in the upper tail of the distribution rather
26      than by a general shift in mean viscosity, consistent with the likelihood of a susceptible
27      sub-population of individuals.
28           Peters et al. (2000b) reported on a prospective cohort study of a subset of male participants
29      from the above-described Augsburg, Germany MONICA study. Based on a survey conducted in
30      1984/85, a sample of 631 randomly selected men/aged 45-64 yr), free of cardiovascular disease at
31      entry, were evaluated in a 3-yr follow-up that examined relationships of air pollution to serum

        April 2002                               8-130        DRAFT-DO NOT QUOTE OR CITE

-------
 1      C-reactive protein concentrations.  C-reactive protein is a sensitive marker of inflammation,
 2      tissue damage, and infections, with acute and chronic infections being related to coronary events,
 3      as well as inflammation being related to systemic hypercoagulability and the onset of acute
 4      ischemic syndromes.  During the 1985 air pollution episode affecting Augsburg and other parts
 5      of Germany, the odds of abnormal  increases in serum C-reactive protein (i.e., >90th percentile of
 6      pre-episode levels = 5.7 mg/L) tripled and associated increases in TSP levels of 26 //g/m3 (5-day
 7      averages) were associated with an odds ratio of 1.37 (95% CI1.08-1.73) for C-reactive protein
 8      levels exceeding the 90th percentile levels in two pollutant models also including SO2 levels.
 9      The estimated odds ratio for a 30 //g/m3 increase in the 5-day mean for SO2 was 1.12 (95% CI
10      0.92 to 1.47; non-significant).
11           Two other recent studies also examined blood indices in relation to PM pollution (Seaton
12      et al., 1999; Prescott et al., 1999).  Seaton and colleagues collected sequential blood samples (up
13      to 12) over an 18 month period in 112 subjects (all over age 60) in Belfast and Edinburgh, UK.
14      Blood samples were analyzed for hemoglobin, packed cell volumes, blood counts, fibrinogen,
15      factor Vn, interleuken 6, C-reactive protein. In a subset of 60 subjects, plasma albumin also was
16      measured. PM10 data monitored by TEOM were collected from ambient sites in each city.
17      Personal exposure estimates for the three days preceding each blood draw were derived from
18      ambient data adjusted by time-activity patterns and I/O penetration factors. No co-pollutants
19      were analyzed. Data were analyzed by analysis of covariance, controlling for city, seasons,
20      temperature, and between-subject differences.  Significant changes in several of the blood indices
21      were observed in association with either ambient or estimated personal PM10 levels.  All changes
22      were negative, except for C reactive protein in relation to ambient PM10, which was positive.
23           Prescott et al. (1999) also investigated factors that might increase susceptibility to adverse
24      cardiovascular events resulting from PM exposure. Using data from a cohort of 1592 subjects
25      aged 55-74 in Edinburgh, UK, baseline measurements of blood fibrinogen and blood and plasma
26      viscosity were examined as modifiers of the effects of PM (indexed by BS) on the incidence of
27      fatal and non-fatal myocardial infarction or  stroke.  All three blood indices were strong predictors
28      of increased cardiac event risk.  However, there was no clear evidence of either a main effect of
29      BS, nor interactions between BS and blood  indices.
30           Two other new studies examined air pollution associations with plasma fibrinogen.
31      Pekkanen and colleagues (2000) analyzed plasma fibrinogen data from a cross-sectional survey

        April 2002                                8-131        DRAFT-DO NOT QUOTE OR CITE

-------
 1      of 4982 male and 2223 female office workers in relation to same-day and previous three-days
 2      concentrations of PM10, black smoke, NO2, CO, SO2, and ozone. In the full analysis, NO2 and
 3      CO were significantly associated with fibrinogen levels. When the analysis was restricted to the
 4      summer season, NO2 and CO, as well as PM10 and black smoke, showed significant univariate
 5      associations. Schwartz (2001) reported significant associations between plasma fibrinogen levels
 6      and PM10 exposures in a subset of the NHANES in cohort. PM10 associations also were
 7      observed for platelet and white cell counts. The PM10 associations were robust when ozone,
 8      NO2, or SO2 was included. CO was not analyzed.
 9           The above findings add support for some intriguing hypotheses regarding possible
10      mechanisms by which PM exposure may be linked with adverse cardiac outcomes. They are
11      especially interesting in terms of implicating both increased blood viscosity and C-reactive
12      protein, a biological marker of inflammatory responses thought to be predictive of increased risk
13      for serious cardiac events.
14
15      8.3.1.4 Issues in the Interpretation of Acute Cardiovascular Effects Studies
16      Susceptible subpopulations.  Because they lack data on individual subject characteristics,
17      ecologic time series studies provide only limited information on susceptibility factors based on
18      stratified analyses.  The relative impact of PM on cardiovascular (and respiratory) admissions
19      reported in ecologic time series studies are generally somewhat higher than those reported for
20      total admissions. This provides some limited support for hypothesizing that acute effects of PM
21      operate via cardiopulmonary pathways or that persons with pre-existing cardiopulmonary disease
22      have greater susceptibility to PM, or both.  Although there is  some data from the ecologic time
23      series studies showing larger relative impacts of PM on cardiovascular admissions in adults aged
24      >65 yr as compared with younger populations, the differences are neither striking  nor consistent.
25      One recent study reported larger CVD hospitalization effects among persons with  current
26      respiratory infections.  The individual-level studies of cardiophysiologic function assessed above
27      generally do suggest that elderly persons with pre-existing cardiopulmonary disease are
28      susceptible to subtle changes  in heart rate variability in association with PM exposures. Because
29      younger and healthier populations have not yet been assessed, it is not yet possible to say whether
30      the elderly clearly have especially increased susceptibility, but this does represent  a reasonable
31      working hypothesis.

        April 2002                                 8-132       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Role of other environmental factors.  The ecologic time series studies published since 1996 all
 2      have controlled adequately for weather influences.  Thus, it is deemed unlikely that residual
 3      confounding by weather accounts for the PM associations observed.  With one possible
 4      exception (Pope et al., 1999a), the roles of meteorological factors have not been analyzed
 5      extensively as yet in the individual-level studies of cardiac function.  Thus, the possibility of
 6      confounding in such studies cannot yet be readily discounted.  Co-pollutants have been analyzed
 7      rather extensively in many of the recent time-series studies of hospital admissions and PM. In
 8      some studies, PM clearly carries an independent association after controlling for gaseous co-
 9      pollutants.  In others, the "PM effects"  are markedly reduced once co-pollutants are added to the
10      model; but this may in part be due to colinearity between PM10 and co-pollutants and/or the
11      gaseous pollutants such as CO having independent effects on cardiovascular function.
12
13      Temporal patterns of responses following PM exposure. The evidence from recent time series
14      studies of CVD admissions suggests rather strongly that PM effects tend to be maximal at lag 0,
15      with some carryover to lag 1, with little evidence for important effects beyond lag 1.
16
17      Relation of CVD  effects to PM size and chemical composition attributes. Insufficient data
18      exist from the time series  CVD admissions literature or from the emerging individual-level
19      studies to provide clear guidance as to which ambient PM components, defined either on the
20      basis of size or composition, determine ambient PM CVD effect potency. The epidemiologic
21      studies published to date have been constrained by the limited availability of multiple PM
22      metrics. Where multiple metrics exist, they often are highly correlated or of differential quality
23      due to differences in numbers of monitoring sites and in monitoring frequency.
24
25      PM effects  on blood characteristics related to CVD events. Interesting, though limited, new
26      evidence has also been derived which is highly suggestive of associations between ambient PM
27      and increased blood viscosity, increased serum C-reactive protein, and fibrinogen (both related
28      to increased risks of serious cardiac events)
29
30
        April 2002                                8-133        DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.3.2  Effects of Short-Term Particulate Matter Exposure on the Incidence of
 2             Respiratory Hospital Admissions and Medical Visits
 3      8.3.2.1 Introduction
 4           Among the most severe morbidity measures evaluated with regard to PM exposure are
 5      hospital admissions. Hospital emergency department (ED) visits represent a related outcome that
 6      has also been studied in relation to air pollution.  Also doctors' visits represent a related health
 7      measure that, although less studied, is relevant to those who also suffer severe health effects.
 8      This latter category of pollution-affected persons can represent a large population, yet one largely
 9      unevaluated due to the usual lack of centralized data regarding doctors' visits.
10           This section evaluates present knowledge regarding the epidemiologic associations of
11      ambient PM exposure with respiratory hospital admissions and medical visits. It intercompares
12      various studies examining each of the size-related PM mass exposure measures (e.g., for PM10)
13      and study results for various PM chemical components vis-a-vis their relative associations with
14      health effects, and their respective extents of coherence with PM associations exhibited across
15      related health effects measures. In the following discussion, the main focus for quantitative
16      intercomparisons is on studies and results considering PM metrics that quantitatively measure
17      mass or a specific mass constituent, i.e.,:  PM10, PM10_25, PM25, sulfates (SO4=),  or acidic aerosols
18      (H+).  Study results for other related PM metrics (e.g., Black Smoke; BS) are also considered, but
19      only qualitatively, primarily with  respect to their coherence or lack of coherence with studies
20      using mass or composition metrics measured in North America.  In order to consider potentially
21      confounding effects of other co-existing pollutants, study results for various PM metrics are
22      presented both for: (1) when the PM metric is the only pollutant in the model; and, (2) the case
23      where a second pollutant (e.g., ozone) is also included. Results from models with more than two
24      pollutants included simultaneously are not used for quantitative estimates of coefficient size or
25      statistical strength, due to increased likelihood of bias and variance inflation due to multi-
26      collinearity of various pollutants (e.g., see Harris, 1975).
27           The approach taken in this section is: first, to summarize briefly results and implications of
28      the 1996 PM AQCD document regarding this topic; then the most important (pertinent for
29      present purposes) findings from newly available key studies published since the 1996 PM AQCD
30      are discussed in the text. More detailed descriptions of these and other new studies are provided
31      in tabular form in Appendix 8B.

        April 2002                                8-134       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Studies of respiratory hospital admissions and medical visits presented in this section were
 2      identified by ongoing Medline searches in conjunction with other search strategies. Specific
 3      studies were summarized in tables and/or text based on criteria that include the following:
 4      (1) preference was given to results reported for PM10, PM10_2 5, and PM2 5 and/or smaller PM,
 5      (2) studies relating respiratory hospital admissions and medical visits to levels of ambient PM
 6      exposure in a quantitative manner are the focus of presentation, and (3) other factors discussed
 7      earlier in Section 8.1.3 of this chapter.
 8
 9      8.3.2.2  Summary of Key Respiratory Hospital Admissions Findings from the 1996
10              Particulate Matter Air Quality Criteria Document
11           In the 1996 PM AQCD, it was found that both COPD and pneumonia hospitalization
12      studies showed moderate, but statistically significant, relative risks in the range of 1.06 to
13      1.25 (or 6 to 25% excess risk increment) per 50 //g/m3 PM10 increase or its equivalent.  While a
14      substantial number of hospitalizations for respiratory illnesses occur in those >65 years of age,
15      there are also numerous hospitalizations for those under 65 years of age.  Several of the
16      hospitalization studies restricted their analysis by age of the individuals, but did not explicitly
17      examine younger age groups.  One exception noted was Pope (1991), who reported an increase in
18      hospitalization for Utah Valley children (aged 0 to 5) for monthly numbers of admissions in
19      relation to PM10 monthly averages, as opposed to daily admissions in relation to daily PM levels
20      used in other studies. Studies examining acute associations between indicators of components of
21      fine particles (e.g., BS; sulfates, SO4=; and acidic aerosols, H+) and hospital admissions were also
22      reported as finding significant relationships.  While sulfates were especially predictive of
23      respiratory health effects, it was not clear whether the sulfate-related effects were attributable to
24      their acidity, to the broader effects of associated combustion-related fine particles,  or to other
25      factors.
26
27      8.3.2.3  New Respiratory-Related Hospital Admissions Studies
28           New studies since 1996 have confirmed PM associations with respiratory hospital
29      admissions.  These studies have examined various admissions categories, including: total
30      respiratory admissions for all ages and by age; asthma for all ages and by age; chronic obstructive
31      pulmonary disease (COPD) admissions (usually for patients > 64 yrs.), and pneumonia

        April 2002                                 8-135        DRAFT-DO NOT QUOTE OR CITE

-------
 1      admissions (for patients > 64 yrs.). Table 8B-2 in Appendix 8B summarizes important details
 2      regarding the study area, study period, study population, PM indices considered and their
 3      concentrations, the methods employed, study results and comments, and the "bottom-line" PM
 4      index percent excess risks per standard PM increment (e.g., 50 //g/m3 for PM10) from studies
 5      published since the 1996 PM AQCD.
 6           The percent excess risk (ER) estimates presented in Table 8B-2 are based upon the relative
 7      risks (RR's) provided by the authors, but converted into percent increments per standardized
 8      increments used by the U.S. EPA to facilitate direct intercomparisons of results across studies, as
 9      discussed in Section 8.1. The ER's shown in the table are for the most positively significant
10      pollutant coefficient.  The maximum lag model is used here to provide an estimate of the
11      pollutant-health effects impact.
12           Among the numerous new epidemiological studies published on PM10 morbidity, many
13      evaluated effects of relatively high PM10 concentrations. However, a large number of studies did
14      evaluate associations at low PM10 concentration levels and associations have been reported by
15      several investigators between acute PM10 exposures and total respiratory-related hospital
16      admissions for numerous U.S. cities with annual mean ambient concentrations extending to
17      below 50//g/m3.
18           The NMMAPS multi-city study (Samet et al., 2000a,b) of PM10 concentrations and hospital
19      admissions by persons 65 and older in 14 U.S. cities is of particular interest. As noted in
20      Table 8-18, this study indicates PM10 effects similar to other cities, but with narrower confidence
21      bands, due to its greater power derived by combining multiple cities in the same analysis. This
22      allows significant associations to be identified, despite the fact that many of the cities considered
23      have relatively small populations and that each of the  14 cities had mean PM10 below 50 //g/m3.
24      The cities considered and their respective annual mean/daily maximum PM10 concentrations
25      (in //g/m3) are: Birmingham (34.8/124.8); Boulder (24.4/125.0); Canton (28.4/94.8); Chicago
26      (36.4/144.7); Colorado Springs (26.9/147.2); Detroit (36.8/133.6);Minneapolis/St Paul
27      (36.8/133.6); Nashville  (31.6/128.0); New Haven (29.3/95.4); Pittsburgh (36.0/139.3);
28      Provo/Orem (38.9/241.0); Seattle (31.0/145.9); Spokane (45.3/605.8); and Youngstown
29      (33.1/104.0). As seen in Table  8-18, the PM10 association remained even when only those days
30      with PM10 less than 50 //g/m3 were considered. The city-specific value results ranged from -0.06
        April 2002                                8-136        DRAFT-DO NOT QUOTE OR CITE

-------
           TABLE 8-18. PERCENT INCREASE IN HOSPITAL ADMISSIONS PER 10-^g/m3
          	INCREASE IN PM,n IN 14 U.S. CITIES	
                                    CVD                   COPD                Pneumonia
                             % Increase    (95% CI)     %Increase    (95% CI)     % Increase    (95% CI)
Constrained lag models (Fixed Effect
One day mean (lag 0)
Previous day mean
Two day mean
(for lag 0 and 1)
PM10 <50 Mg.m3
(two day mean)
Quadratic distributed lag
Unconstrained distributed
Fixed effects estimate
Random effects estimate
1.07
0.68
1.17
1.47
1.18
Lag
1.19
1.07
Estimates)
(0.93,
(0.54,
(1.01,
(1.18,
(0.96,

(0.97,
(0.67,
1.22)
0.81)
1.33)
1.76)
1.39)

1.41)
1.46)
1
1
1
2
2

2
2
.44
.46
.98
.63
.49

.45
.88
(1.00,
(1.03,
(1.49,
(1.71,
(1.78,

(1.75,
(0.19,
1.89)
1.88)
2.47)
3.55)
3.20)

3.17)
5.64)
1.
1.
1.
2.
1.

1
2.
57
31
98
84
68

.9
07
(1.27,
(1.03,
(1.65,
(2.21,
(1.25,

(1.46,
(0.94,
1.87)
1.58)
2.31)
3.48)
2.11)

2.34)
3.22)
         Source: Samet et al. (2000a,b)
 1     for Boulder to 6.43 for Detroit, with a combined result of 9.9 for fixed effects and 8.7 for random
 2     effects models for 2 day mean PM10 for values less than 50 //g/m3 for CVD as an example.
 3          Janssen et al. (2002) did further analyses for the Samet et al. (2000a,b) 14-city data set
 4     examining the associations for the variable prevalence in AC and/or the contribution of different
 5     sources to total PM10.  For COPD and pneumonia, the associations were less significant, but the
 6     pattern of association were similar to that for CVD as discussed in Section 8.3.1.
 7          If day-to-day increases in air pollution cause rises in hospital admissions, as indicated by
 8     time-series studies, then short-term removal of pollution should lower admissions. However, it
 9     is rarely possible to test this hypothesis by examining a situation when pollution sources are
10     abruptly "turned off and then "turned on" again.  One past case in point was a steel mill strike
11     and concomitant reductions in both PM and respiratory admissions that were experienced in Utah
12     Valley,  but not in surrounding valleys with out the steel mill, as documented by Pope (1991).
13     A more broadly relevant case where this hypothesis was similarly tested was a study of air
14     quality improvements during the Atlanta Summer Olympics of 1996 (Friedman et al., 2001).
15     These improvements were compared to changes that occurred in children's hospital admissions,
16     while weather and other "natural" influences on admissions stayed unchanged from normal.

       April 2002                                8-137        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Compared to a baseline period, traffic related pollution declined, with PM10 levels declining by
 2      16%, and ozone by 28% as a result of the alternative mass transportation strategy implemented to
 3      reduce road traffic during the Games. At the same time, SO2, not related to traffic, actually
 4      increased during the Games. Concentrations of both PM and ozone also rose noticeably after the
 5      end of the Olympics. A significant reduction in asthma events was associated with ozone
 6      concentration, but the PM10 association was not statistically significant. While the high
 7      correlation between PM and ozone limit the ability to determine which pollutant may have
 8      accounted for the reduction in asthma events, this study supports the hypothesis that reductions
 9      of acute air pollution levels can provide immediate health improvements.
10           Other U.S. studies finding associations of respiratory-related hospital admissions or
11      medical visits with PM10 levels extending below 50 //g/m3 include: Schwartz (1995) in Tacoma;
12      Schwartz (1994) in Minneapolis; Schwartz et al. (1996b) in Cleveland;  Sheppard et al. (1999) in
13      Seattle; Gwynn et al. (2000) in Buffalo, NY; Linn et al.  (2000) in Los Angeles, Nauenberg and
14      Basu (1999) in Los Angeles; and Moolgavkar et al.  (1997) in Minneapolis-St. Paul, MN, but not
15      in Birmingham, AL. The excess risk estimates appear to most consistently fall in the range of
16      5-25% per 50 //g/m3 PM10 increment, with those for asthma visits and hospital admissions
17      usually being higher than those for COPD and pneumonia hospital admissions.
18           Similar associations between increased respiratory related hospital admissions/medical
19      visits and relatively low short-term PM10 levels were also reported by various investigators for
20      several non-U.S. cities. Wordley et al.  (1997), for example, reported positive and significant
21      associations between PM10 (mean = 25.6 //g/m3, max. = 131 //g/m3) and respiratory admissions in
22      Birmingham, UK; and Atkinson et al. (1999a) found significant increases in hospital admissions
23      for respiratory disease to be associated with PM10 (mean = 28.5 //g/m3) in London, UK. Hagen
24      et al. (2000) and Prescott et al. (1998) also found positive but non-significant PM10 associations
25      with hospital admissions in Drammen,  Sweden (mean = 16.8 //g/m3)  and Edinburgh, Scotland
26      (mean = 20.7 //g/m3), respectively. Admissions in Drammen considered relatively small
27      populations, limiting statistical power in this study.  Petroeschevsky et al. (2001)  examined
28      associations between outdoor air pollution and hospital  admissions in Brisbane, Australia during
29      1987-1994 using a light scattering index (BSP) for fine PM.  The levels of PM are quite low in
30      this city, relative to most U.S.  cities.  BSP was positively and significantly associated with total
31      respiratory admissions, but not for asthma.

        April 2002                                8-138       DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.3.2.3.1  Particulate Matter Mass Fractions and Composition Comparisons
 2           While PM10 mass is the metric most often employed as the particle pollution index in the
 3      U.S. and Canada, some new studies have begun to examine the relative roles of various PM10
 4      mass fractions and chemical constituents (such as SO4=) in the PM-respiratory hospital
 5      admissions association.  Several new studies report significant associations of increased
 6      respiratory-cause medical visits and/or hospital admissions with ambient PM2 5 and/or PM10_2 5
 7      ranging to quite low concentrations.  These include the Lippmann et al. (2000) study in Detroit,
 8      where all PM metrics (PM10, PM2 5, PM10_2 5, H+) were positively related to pneumonia and COPD
 9      admissions among the elderly (aged 65+ yr) in single pollutant models, with their RR values
10      generally remaining little changed (but with broader confidence intervals) in multipollutant
11      models including one or more gaseous pollutant  (e.g., CO, O3, NO2,  SO2).  Excess risks for
12      pneumonia admissions in the one pollutant model were 13% (3.7, 22) and 12% (0.8, 24) per
13      25 //g/m3 of PM25 and PM10_25, respectively; those for COPD admissions were 5.5% (-4.7, 17)
14      and 9.3% (-4.4, 25) per 25 //g/m3 PM25 and PM10_25, respectively. Also of note,  Moolgavkar
15      found ca. 5.0% excess risk for COPD hospital admissions among the elderly (64+ yr) in
16      Los Angeles to be significantly related to both PM2 5 and PM10_25 in one pollutant models; but the
17      magnitudes of the risk estimates dropped by more than half to non-statistically significant levels
18      in two-pollutant models including CO. In the same study, similar magnitudes of excess risk (i.e.,
19      in the range of ca. 4 to 7%) were found in one-pollutant models to be associated with PM2 5 or
20      PM10_25 for other age groups (0-19 yr; 20-64 yr) in Los Angeles, as  well.  Moolgavkar et al.
21      (2000) also found 5.6% (0.2, 11.3) excess risk for all-ages COPD hospital admissions per
22      25 //g/m3 PM2 5 increase in King County, WA.
23           Tolbert et al. (2000a) reported no significant associations of PM2 5 or PM10_2 5 with COPD
24      emergency department visits in Atlanta, based on data from less than half of all participating
25      hospitals and ca. 1 yr of supersite air quality data. However, more complete analyses from all
26      participating hospitals over a longer time period  are required before this can be adequately
27      evaluated.
28           Gwynn et al. (2000) considered a 2.5 yr period (May 1988-Oct. 1990) in the Buffalo, NY
29      region in a time-series analysis of daily mortality and hospital admissions for total, respiratory,
30      and circulatory hospital admissions categories. Pollutants considered included: PM10, H+, SO4=
31      COH, O3, CO, SO2, and NO2. The H+ and SO4= concentrations were determined from daily PM2 5

        April 2002                               8-139        DRAFT-DO NOT QUOTE OR CITE

-------
 1      samples not analyzed for mass (in order to avoid possible acid neutralization) using the
 2      sequential acid aerosol system.  Various modeling techniques were applied to control for
 3      confounding of effect estimates due to seasonality, weather and day-of-week effects.  They found
 4      multiple significant pollutant-health effect associations, the most significant being between SO4=
 5      and respiratory hospital admissions. When calculated in terms of increments employed across
 6      analyses in this report,  various PMRR's were:  PM10RR=1.11, 95% C.I.=1.05-1.18(for
 7      50 Aig/m3); H+ RR=1.06, 95% C.I.=1.03-1.09 (for 75 nmoles/m3= 3.6 Mg/m3, if as H2SO4); and
 8      SO4= RR=1.08, 95% C.I.=1.04-1.12 (for 155 nmoles/m3=15 //g/m3).  As in the Burnett et al.
 9      (1997a) study described below, H+ yielded the highest RR per //g/m3 of concentration. These
10      various PM metric associations were not significantly affected by inclusion of gaseous
11      co-pollutants in the regression model.  Thus, all PM components considered except COH were
12      found to be associated  with increased hospital admissions, but H+, SO4= and O3 had the most
13      coherent associations with respiratory admissions.
14           Lumley and Heagerty (1999) illustrate the effect of reliable variance estimation on data
15      from hospital admissions for respiratory disease on King County, WA for eight years (1987-94),
16      together with air pollution and weather information. However, their weather controls were
17      relatively crude (i.e., seasonal dummy variables and linear temperature terms). This study is
18      notable for having compared sub-micron PM (PMLO) versus coarse PM10.LO and for finding
19      significant hospital admission associations only with PML0. This may suggest that the PM2 5 vs.
20      PM10 separation may not  always be sufficient to differentiate submicron fine particle vs. coarse-
21      particle toxicities.
22           Asthma hospital admission studies conducted in various U.S. communities provide
23      additional important new data.  Of particular note is a study by Sheppard et al. (1999) which
24      evaluated relationships between measured ambient pollutants (PM10, PM2 5, PM10_2 5, SO2, O3 and
25      CO) and non-elderly adult (<65 years of age) hospital admissions for asthma in Seattle, WA.
26      PM and CO were found to be jointly associated with asthma admissions. An estimated 4 to 5%
27      increase in the rate of asthma hospital admissions (lagged 1 day) was reported to be associated
28      with interquartile range changes in PM indices  (19 //g/m3 for PM10, 11.8 //g/m3 for PM2 5, and
29      9.3 //g/m3 for PM10_25), equivalent to excess risk rates as follows:  13% (95% CI 05, 23) per
30      50 Aig/m3 for PM10; 9% (95% CI 3, 14) per 25 Aig/m3 PM2 5; 11%  (95% CI 3, 20) per 25 Aig/m3
31      PM10_2 5. Also of note for the same region is the Norris et al. (1999) study showing associations

        April 2002                                8-140       DRAFT-DO NOT QUOTE OR CITE

-------
 1      of low levels of PM2 5 (mean = 12 //g/m3) with markedly increased asthma ED, i.e., excess risk =
 2      44.5% (CI 21.7, 71.4) per 25 //g/m3 PM25.
 3           Burnett et al. (1997a) evaluated the role that the ambient air pollution mix, comprised of
 4      gaseous pollutants and PM indexed by various physical and chemical measures, plays in
 5      exacerbating daily admissions to hospitals for cardiac diseases and for respiratory diseases
 6      (tracheobronchitis, chronic obstructive long disease, asthma, and pneumonia).  They employed
 7      daily measures of PM25 and PM10_25, aerosol chemistry (sulfates and H+), and gaseous pollutants
 8      (ozone, nitrogen dioxide, sulfur dioxide, and carbon monoxide) collected in Toronto, Ontario,
 9      Canada, during the summers of 1992, 1993, and 1994. Positive associations were observed for
10      all ambient air pollutants for both respiratory and cardiac diseases.  Ozone was the most
11      consistently significant pollutant and least sensitive to adjustment for other gaseous and
12      particulate measures. The PM associations with the respiratory hospital admissions were
13      significant for: PM10 (RR=1.1 Ifor 50 //g/m3; CI=1.05-1.17); PM2 5 (fine) mass (RR=1.09 for
14      25 Aig/m3; CI=1.03-1.14); PM10.25 (coarse) mass (RR=1.13 for 25 //g/m3; CI=1.05-1.20); sulfate
15      levels (RR=1.11 for  155 nmoles/m3=15 //g/m3;  CI=1.06-1.17); and H+ (RR=1.40 for
16      75 nmoles/m3= 3.6 //g/m3,  as H2SO4; CI=1.15-1.70). After simultaneous inclusion of ozone in
17      the model, the associations with the respiratory hospital admissions remained significant for:
18      PM10(RR=1.10;CI=1.04-1.16);fmemass(RR=1.06;CI=1.01-1.12); coarse mass (RR= 1.11;
19      CI=1.04-1.19); sulfate levels (RR=1.06; CI=1.0-1.12); and H+(RR= 1.25; CI=1.03-1.53), using
20      the same increments. Of the PM metrics considered here, H+ yielded the highest RR estimate.
21      Regression models that included all recorded pollutants simultaneously (with high
22      intercorrelations among the pollutants) were also presented.
23           There have  also been numerous new time-series studies examining associations between air
24      pollution and respiratory-related hospital admissions in Europe, as summarized in Appendix 8B,
25      Table 8B-2, but most of these studies relied primarily on black smoke (BS) as their PM metric.
26      BS is a particle reflectance measure that provides an indicator of particulate blackness and is
27      highly correlated  with airborne carbonaceous particle concentrations (Bailey and Clayton, 1982).
28      In the U.S., Coefficient of Haze (COH) is a metric of particle transmittance that similarly most
29      directly represents a  metric of particle blackness and ambient elemental carbon concentration
30      (Wolff et al., 1983) and has been found to be highly correlated with BS (r = 0.9) (Lee et al.,
31      1972).  However, the relationship between  airborne carbon and total mass of overall aerosol

        April 2002                                8-141        DRAFT-DO NOT QUOTE OR CITE

-------
 1      (PM) composition varies over time and from locality to locality, so the BS-mass ratio is less
 2      reliable than the BS-carbon relationship (Bailey and Clayton, 1982).  This means that the BS-
 3      mass relationship is likely to be very different between Europe and the U.S., largely due to
 4      differences in local PM source characteristics (e.g., percentages of diesel powered motor
 5      vehicles).  Therefore, while these European BS-health effects studies are of qualitative use for
 6      evaluating the PM-health effects associations, they are not as useful for quantitative assessment
 7      of PM effects relevant to the U. S.
 8           Hagan et al. (2000) compared the association of PM10 and co-pollutants with hospital
 9      admissions for respiratory causes in Drammen, Norway during 1994-1997. Respiratory
10      admissions averaged only 2.2 per day; so, the power of this analysis is weaker than studies
11      looking at larger populations and longer time periods. The HEI IB Multi-city Report modeling
12      approach was employed. While a significant association was found for PM10 as a single
13      pollutant, it became non-significant in multiple pollutant models. In two pollutant models, the
14      associations and effect size of pollutants were generally diminished, and when all eight pollutants
15      were considered in the model, all pollutants became non-significant.  These results are typical of
16      the problems of analyzing and interpreting the coefficients of multiple pollutant models when the
17      pollutants are even moderately inter-correlated over time. A unique aspect of this work was that
18      benzene was considered in this community strongly affected by traffic pollution. In two  pollutant
19      models, benzene was most consistently still associated.  The authors conclude that PM is mainly
20      an indicator  of air pollution in this city and that emissions from vehicles seem most important for
21      health effects.  Thompson et al. (2001) report a similar result in Belfast, Northern Ireland, where,
22      after adjusting for multiple pollutants, only the benzene level was independently associated with
23      asthma emergency  department admissions.
24           The most recent European air pollution health effects analyses have mainly been conducted
25      as part of the APHEA study, which evaluated 15 European cities from  10 different countries with
26      a total population of over 25 million. All studies used a standardized data collection and analysis
27      approach, which included consideration of the same suite of air pollutants (BS, SO2, NO2, SO2,
28      and O3) and the use of time-series regression addressing:  seasonal and other long-term patterns;
29      influenza epidemics; day of the week; holidays; weather; and autocorrelation (Katsouyanni et al.,
30      1996). The general coherence of the APHEA results with other results gained under different
31      conditions strengthens the argument for causality in the air pollution-health effects association.

        April 2002                                8-142       DRAFT-DO NOT QUOTE OR CITE

-------
 1      In earlier studies, the general use of the less comparable suspended particle (SPM) measures and
 2      BS as PM indicators in some of the APHEA locations and analyses lessens the quantitative
 3      usefulness of such analyses in evaluating associations between PM and health effects most
 4      pertinent to the U.S. situation.  However, Atkinson et al. (2001) report results of PM10 analysis in
 5      a study of eight APHEA cities.
 6
 7      8.3.2.3.2 Methodological Studies
 8           One study by Lumley and Sheppard (2000) applied a simulation approach to examine the
 9      effects of seasonal confounding and model selection on hospital admissions effect estimates in
10      Seattle, Washington.  It was found that the bias introduced by model  selection was small, but
11      could be on the same order as the estimated health impacts.  This problem was the case when
12      seasonal adjustments were not accounted for in the model, and was larger when the maximum
13      lag of many was selected.  However, the distributed lag nature of air pollution effects was not
14      simulated, making the tests of the maximum lag vs. real effect unrealistic.  Also, in the now usual
15      case in which seasonal adjustments were included, any bias was consistently much smaller and
16      was non-significant in cases where there was no simulated PM effect. This suggests that model
17      selection bias is not a concern in the type of modeling routinely done today, and also points out
18      the need to consider statistical  significance when evaluating and inter-comparing effect estimates.
19           Several studies looked at the potential influence of exposure error on pollutant impact
20      estimates.  Lipfert (2000)  surveyed the sources and magnitudes of such errors and concluded that
21      they can have "profound effects on the results of epidemiological studies", noting especially
22      comparisons between fine particles and less accurately measured coarse particle associations
23      with health.  In a related paper, Lipfert and Wyzga (1999) consider this issue and argue against
24      the use of statistical significance for pollutant impact inter-comparisons because the distribution
25      characteristics of the variable can play a role in its strength of association.  They recommend the
26      use of effect size to inter-compare pollutants, even though differing choices of a particular
27      increment for the pollutant effect estimation (e.g., IQR vs. mean vs. median vs. max-mean, etc.)
28      will usually give differing rankings across pollutants, as their relative sizes are influenced by
29      pollutant distribution, as well.  The authors argue that, until uncertainties have been fully
30      explored, such as those introduced by  exposure error, such epidemiological studies should only
31      be considered as suggestive of causality. Huang and Batterman (2000) also look at exposure

        April 2002                               8-143        DRAFT-DO NOT QUOTE OR CITE

-------
 1      error, noting that most studies have not looked at the population exposure errors, and concluding
 2      that, "Unless exposure levels among groups are verified, it cannot be determined whether
 3      nonsignificant associations between exposure and health endpoints indicate a lack of measurable
 4      health effects, or are merely a result of exposure misclassification".  However, a recent study by
 5      Sheppard and Damian (2000) using quasi-likelihood simulation techniques investigated the effect
 6      of pure measurement error, but not spatial or within-day personal exposure variations,
 7      concluding that adjustment for measurement error does not alter the conclusions from the time
 8      series analyses typically reported in the literature.
 9           Schwartz (2001) examined another relevant methodological and mechanistic aspect of the
10      PM-health association: the harvesting question (i.e., as to whether the associations between air
11      pollution and health effects are due to the moving up of an event [e.g., death] that would have
12      happened in a few days, anyway, or not)? Using a smoothing technique, he estimated the "net"
13      change in mortality and hospital admissions in Chicago associated with PM, after accounting for
14      any decline in events in the follow-up period, ranging from 15 to 60 days. Analyses indicated
15      that the health effect estimates stayed the same, or increased, when any harvesting effects were
16      adjusted for in the analysis.  He concluded that the results are consistent with air pollution
17      increasing the size of the risk pool, and for most of the air pollution associated deaths being
18      advanced by months to years.
19          Dewanji and Moolgavkar (2000) implemented a flexible parametric model analysis in the
20      example of multiple hospital admissions for chronic respiratory disease in King County, WA,
21      that views the data on each subject as the realization of a point process, which allows
22      incorporation of subject specific covariate and the previous history of the process.  In single
23      pollutant analyses, measures of PM (PM10 and PM25)  and CO are associated with hospital
24      admissions.  The effect of PM was stronger than that of CO in the multipollutant models. This
25      result is inconsistent with other analysis of the same data (Moolgavkar et al., 2000) which find
26      that the effect of PM becomes insignificant when CO  is simultaneously considered in the
27      analysis.
28          Pollen is an atmospheric constituent that might potentially be a factor that may confound
29      PM-asthma admissions associations, if it is correlated with both PM and hospital admissions.
30      In a London study, airborne pollen did not confound the analysis of air pollution (including black
31      smoke) and daily admissions for asthma during the time period 1987-1992 (Anderson et al.,

        April 2002                                8-144        DRAFT-DO NOT QUOTE OR CITE

-------
 1      1998).  Moolgavkar et al. (2000) in a study in Seattle found that adding pollens to PM time-series
 2      regressions of respiratory admissions diminished the PM effect estimates more for PM10 than that
 3      for PM2 5.  Confounding by pollens would require correlation between daily PM2 5 levels and
 4      seasonal pollution events and weather-related specification events.  Further, for a different
 5      outcome measure, Delfino et al. (1996) found that pollen was not associated with asthma
 6      symptoms in an asthma panel (see Section 8.3.3)
 7
 8      8.3.2.4 Key New Respiratory Medical Visits Studies
 9           As discussed above, medical visits include both hospital emergency department (ED) visits
10      and doctors'  office visits. As in the past PM AQCD's, most of the available morbidity studies
11      presented in Appendix 8B,  Table 8B-3 are of ED visits and their associations with air pollution.
12      These studies collectively confirm the results provided in the previous AQCD, indicating a
13      positive and significant association between ambient PM levels and increased respiratory-related
14      hospital visits.
15           Of the medical visit and hospital admissions studies since the 1996 PM AQCD, the most
16      informative are those that evaluate health effects associations at levels below previously well-
17      implicated PM concentrations. In the case of medical visits, the Norris et al. (1999, 2000) studies
18      of asthma ED visits found significant PM-associated health effects among children in Seattle,
19      even at quite low average PM levels and even after incorporating the effects of other air
20      pollutants (study mean PM10 = 21.7 //g/m3; estimated mean PM2 5 =12 //g/m3).  Tolbert et al.
21      (2000b) reported a significant PM10 association with pediatric ED visits in Atlanta where the
22      maximum PM10 concentration was  105 //g/m3.  The Lipsett et al. (1997) study of winter air
23      pollution and asthma emergency visits in Santa Clara County, CA, may provide insight where
24      one of the principal sources of PM10 is residential wood combustion (RWC).  Their results
25      demonstrate an association between PM and asthma concentrations. Also, Delfino et al. (1997)
26      found significant PM10 and PM2 5 associations for respiratory ED visits among older adults in
27      Montreal when mean PM10 =21.7 //g/m3 and mean PM2 5 = 12.2 Mg/m3. Medina et al. (1997)
28      reported significant associations between doctor's asthma house visits and PM13 (which would
29      have a slightly higher concentration value than PM10) in Paris when mean PM13 = 25 //g/m3 and
30      maximum daily PM13 = 95 //g/m3. Hajat et al. (1999) reported significant PM10 associations with
31      asthma doctor's visits for children and young adults in London when mean PM10 = 28.2 //g/m3

        April 2002                               8-145        DRAFT-DO NOT QUOTE OR CITE

-------
 1      and the PM10 90th percentile was only 46.4 //g/m3. Overall, then, numerous new medical visits
 2      studies indicate PM-health effects associations at lower PM2 5 and PM10 levels than demonstrated
 3      previously for this health outcome.
 4
 5      8.3.2.4.1  Scope of Medical Visit Morbidity Effects
 6           Several of these recent medical visit studies consider a new endpoint for comparison with
 7      ED visits: visits in the primary care setting. In particular, key studies showing PM-health effects
 8      associations for this health outcome include: the study by Medina et al. (1997) for Paris, France
 9      which evaluated doctors' visits to patients in that city; the study by Hajat et al. (1999) that
10      evaluated the relationship between daily General Practice (GP) doctor consultations for asthma
11      and other lower respiratory disease (LRD) and air pollution in London, UK; the study by
12      Choudhury et al. (1997) of private asthma medical visits in Anchorage, Alaska; and the study by
13      Ostro et al.  (1999b) of daily visits by young children to primary care health clinics in Santiago,
14      Chile for upper or lower respiratory symptoms.
15           While limited in number, the above studies collectively provide new insight into the fact
16      that there is a broader scope of severe morbidity associated with PM air pollution exposure than
17      previously documented. As the authors of the London study note:  "There is less information
18      about the effects of air pollution in general practice consultations but, if they do exist, the public
19      health impact could be considerable because of their large numbers." Indeed, the studies of Paris
20      doctors' house calls and London doctors' GP office visits both indicate that the effects of air
21      pollution, including PM, can affect many more people than indicated by hospital  admissions
22      alone.
23           These new studies also provide indications as to the quantitative nature of medical visits
24      effects, relative to those for hospital admissions.  In the London case, comparing  the number of
25      admissions from the authors' earlier study (Anderson et al.,  1996) with those for  GP visits in the
26      1999 study (Hajat et al., 1999) indicates that there are approximately 24 asthma GP visits for
27      every asthma hospital admission in that city. Also, comparing the PM10 coefficients indicates
28      that the all-ages asthma effect size for the GP visits (although not statistically different) was
29      about 30%  larger than that for hospital  admissions. Similarly, the number of doctors' house calls
30      for asthma approximated 45/day in Paris (based on an average 9 asthma house calls in the SOS-
31      Medocina data base, representing 20%  of the total; Medina et al., [1997]), versus an average

        April 2002                                8-146        DRAFT-DO NOT QUOTE OR CITE

-------
 1      14 asthma admissions/day (Dab et al., 1996), or a factor of 3 more doctors' house calls than
 2      hospital admissions. Moreover, the RR for Paris asthma doctors' house calls was much higher
 3      than asthma admissions (RR=1.18 for 25 //g/m3 BS for house calls vs. RR=1.01 per 25 //g/m3 BS
 4      for hospital admissions). Thus, these new studies suggest that looking at only hospital
 5      admissions and emergency hospital visit effects may greatly underestimate the overall numbers
 6      of respiratory morbidity events in a population due to acute ambient PM exposure.
 7
 8      8.3.2.4.2 Evaluation of Factors Potentially Affecting Respiratory Medical Visit
 9               Study Outcomes
10           Some newly available studies have examined certain factors that might extraneously affect
11      the outcomes of PM-medical visit studies.  Stieb et al. (1998a) examined the occurrence of bias
12      and random variability in diagnostic classification of air pollution and daily cardiac respiratory
13      emergency department visits such as asthma, COPD, respiratory infection  and cardiac. They
14      concluded that there was no evidence of diagnostic bias in relation to daily air pollution levels.
15      Also, Stieb et al. (1998b) reported that for a population of adults visiting an emergency
16      department with cardiac respiratory disease, fixed site sulfate monitors appear to accurately
17      reflect daily variability in average personal exposure to particulate sulfate,  whereas paniculate
18      acid exposure was not as well represented  by fixed site monitors. Another study investigated
19      possible confounding of respiratory visit effects due to pollens. In London, Atkinson et al.
20      (1999a) studied the association between the number of daily visits to emergency departments for
21      respiratory complaints and measures of outdoor air pollution for PM10, NO2, SO2 and CO. They
22      examined different age groups and reported the strongest association for children for visits for
23      asthma, but were unable to separate the effects of PM10 and SO2.  Pollen levels did not influence
24      the results, similar to results from the asthma panel studies described below in Section 8.3.3.
25
26      8.3.2.5  Identification of Potential Susceptible Subpopulations
27           Associations between ambient PM measures and respiratory admissions have been found
28      for all age groups, but older adults and children have been indicated by a number of hospital
29      admissions studies to exhibit the most consistent PM-health effects associations in the literature.
30      As reported in this and previous PM AQCDs, numerous studies of older adults (e.g., those 65+
31      years of age) have related acute PM exposure with an increased incidence  of hospital admissions

        April 2002                                8-147        DRAFT-DO NOT QUOTE OR CITE

-------
 1      (e.g., see Anderson et al, 1998). However, only a limited number have specifically studied
 2      children as a subgroup. Burnett et al. (1994) examined the differences in air pollution-hospital
 3      admissions associations as a function of age in the province of Ontario, reporting that the largest
 4      percentage increase in admissions was found among infants (neonatal and post-neonatal, one year
 5      or less in age).
 6           Considerable efforts have aimed at identifying and quantifying air pollution effects among
 7      potentially especially susceptible sub-populations of the general public, especially among
 8      children. Burnett et al. (200Ib) studied the association between air pollution and hospitalization
 9      for acute respiratory diseases in children less than 2 years of age in Toronto, Canada during
10      1980-1994. In single pollutant analyses, PM25, PM10_25, ozone, NO2, and CO were all significant
11      predictors  of young children's respiratory admissions, but only ozone and CO stayed significant
12      in 2 pollutant models, with ozone also having a robust effect estimate in co-pollutant models.
13      These effects were found to be bigger than those for older children or adults studied in a previous
14      publication (Burnett et al.,  1994). Two other recent studies of children's morbidity support the
15      indication  of air pollution effects among children. Pless-Mulloli et al.  (2000) looked at
16      children's  respiratory health and air pollution near opencast coal mining sites in a cohort of
17      nearly 5,000 children aged 1-11 in England. Mean levels were not high (mean less than
18      20 //g/m3 PM10), but statistically significant PM10 associations were found with respiratory
19      symptoms. A roughly 5 percent increase General Practitioner medical visits was also noted, but
20      the effect was not significant in this cohort.  Dabaca et al. (1999) found an association between
21      levels of fine  PM  and emergency visits for pneumonia and other respiratory illnesses among
22      children less than  15 years of age living in the eastern part of Santiago, Chile, where the levels of
23      PM25 were very high (mean=71.3 //g/m3) during 1995-1996.  The authors found it difficult to
24      separate out the effects of various pollutants, but concluded that PM (especially the fine
25      component) is associated with the risk of these respiratory illnesses. Overall, these new studies
26      support past assertions that children, and especially  those less than 2 years of age, are especially
27      susceptible to the  adverse health effects of air pollution.
28           Several  new studies have further investigated the hypothesis that the elderly are especially
29      affected by air pollution. Zanobetti et al. (2000b) analyzed Medicare hospital admissions for
30      heart disease, COPD, and pneumonia in Chicago, IL between 1985 and 1994, finding that the
31      PM10 risk estimate was nearly doubled by the co-presence of respiratory infections, but that there

        April 2002                                8-148        DRAFT-DO NOT QUOTE OR CITE

-------
 1      was no effect modification by sex or race. Zanobetti et al. (2000a) similarly examined PM10
 2      associations with hospital admissions for heart and lung disease in ten U.S. cities, finding an
 3      overall association for COPD, pneumonia, and CVD. They found that these results were not
 4      significantly modified by poverty rate or minority status in this population of Medicare patients.
 5      Ye et al. (2001) examined emergency transports to the hospital. Both PM10 and N02 levels were
 6      significantly associated with daily hospital transports for angina, cardiac insufficiency,
 7      myocardial infarction, acute and chronic bronchitis, and pneumonia. The pollutant effect sizes
 8      were generally found to be greater in men than in women, except those for angina and acute
 9      bronchitis, which were the same across genders.  Thus, in these various studies, cardiopulmonary
10      hospital visits and admissions among the elderly were seen to be consistently associated with PM
11      levels across numerous locales in the U.S. and abroad, generally without regard to race or
12      income; but sex was sometimes an effect modifier.
13           Gwynn and Thurston (2001) examined race as a factor in the air pollution-hospital
14      admissions association.  This study considered persons of all ages in New York City during
15      1988-1990, which provided a large and diverse population ideal for investigating this question.
16      Although not statistically different from each other, the various air pollutants' relative risk
17      estimates for the Hispanic non-White category in NYC were generally larger in magnitude than
18      those of the non-Hispanic White group.  The greatest difference between the White and non-
19      White subgroups was observed for O3, but the same trend was found for PM10 and sulfates.
20      However, when insurance status was used as an indicator of socioeconomic/health coverage
21      status, higher RR's were indicated for the poor and working poor (i.e., those on Medicaid and the
22      uninsured) than for economically better off (i.e., the privately insured), even among the non-
23      Hispanic Whites.  This result is consistent with the past analyses in California by Nauenberg and
24      Basu (1999). Thus, the within-race analyses by insurance coverage suggested that most of the
25      generally higher effects of air pollution found for minorities (i.e.,  Hispanics and non-Whites)
26      were actually caused by overall socioeconomic and/or health care disparities in these populations
27      vs. the generally wealthier non-Hispanic White population.  This  suggests that those living in
28      poverty may represent an especially affected sub-population.
29           The respiratory-related hospital admissions studies summarized in Appendix 8B, Table
30      8B-2 reveals that the PM RR's for all children (e.g., 0-18) are not usually noticeably larger than
31      those for adults, but such comparisons of RR's must adjust for  differences in the baseline risks

        April 2002                                 8-149       DRAFT-DO NOT QUOTE OR CITE

-------
 1      for each group.  For example, if hospital admissions per 100,000 per day for young children are
 2      double the rate for adults, then they will have a pollution relative risk (RR) per //g/m3 that is half
 3      that of the adults given the exact same impact on admissions/100,000///g/m3/day.  Thus, it is
 4      important to adjust RR's or Excess Risks (ER's) for each different age groups' baseline, but this
 5      information is usually not available (especially regarding the population catchment for each age
 6      group in each study).
 7           One of the few indications that is notable when comparing children with other age group
 8      effect estimates in Table 8B-2 is the higher excess risk estimate for infants (i.e., the group <1  yr.
 9      of age) in the Gouveia and Fletcher (2000) study, an age group that has estimated risk estimate
10      roughly twice as large as for other children or adults.  This is confirmatory of the excess risk
11      pattern previously found in the above-noted Burnett et al. (1994) study for respiratory-related
12      hospital  admissions.
13
14      8.3.2.6  Summary of Key Findings on Acute Particulate Matter Exposure and
15              Respiratory-Related Hospital Admissions and Medical Visits
16           The results of new studies discussed above are generally consistent with and supportive of
17      findings presented in the previous PM AQCD (U.S. Environmental Protection Agency, 1996a),
18      with regard to ambient PM associations of short-term exposures with respiratory-related hospital
19      admissions/medical visits.  Excess risk estimates for specific subcategories of respiratory-related
20      hospital  admissions/medical visits for U.S. cities are summarized in Tables 8-19 to 8-22 and
21      graphically depicted in Figure 8-13. The  excess risk estimates fall most consistently in the range
22      of 5 to 25% per 50 //g/m3 PM10 increments, with those for asthma visits and hospital admissions
23      tending to be somewhat higher than for COPD and pneumonia hospital admissions.
24           More limited new evidence  substantiates increased risk of respiratory-related hospital
25      admissions due to ambient fine particles (PM2 5, PMLO, etc.) and also points towards such
26      admissions being associated with  ambient coarse particles (PM10_2 5).  Excess risk estimates tend
27      to fall in the range of ca. 5.0 to 15.0% per 25 //g/m3 PM25 or PM10_25 for overall respiratory
28      admissions or for COPD admissions, whereas larger estimates are found for asthma admissions
29      (ranging upwards to ca.  40 to 50% for children < 18 yr. old in one study).
30           Various new medical visits  studies (including non-hospital physician visits) indicate that
31      the use of hospital admissions alone can greatly understate the total clinical morbidity effects of

        April 2002                               8-150       DRAFT-DO NOT QUOTE OR CITE

-------
  TABLE 8-19. SUMMARY OF UNITED STATES PM10 RESPIRATORY HOSPITAL
                         ADMISSION STUDIES
Reference
Moolgavkar et al.
(1997)
Minneapolis, St. Paul
(MSP
Birmingham (BI)
BI
Gwynn et al. (2000)
Linn et al. (2000)
Schwartz et al.
(1996b)
Zanobetti and
Schwartz (2001)
Samet et al. (2000a,b)
Zanobetti et al.
(2000a)
Chen et al. (2000)
Zanobetti et al.
(2000b)
Lippman et al. (2000)
Moolgavkar et al.
(2000)
Moolgavkar (2000a)
Samet et al. (2000a,b)
Zanobetti et al.
(2000b)
Lippman et al. (2000)
Zanobetti et al.
(2000a)
Jacobs etal. (1997)
Sheppard et al. (1999)
Nauenberg and Basu
(1999)
Outcome
Measures
Respiratory
Respiratory
Respiratory
Respiratory
Respiratory
Respiratory
Respiratory
Respiratory
COPD
COPD
COPD
COPD
COPD
COPD
COPD
COPD (>64 yrs)
(median)
Pneumonia
Pneumonia
Pneumonia
Pneumonia
Asthma
Asthma
Asthma
Mean Levels
MSP PM10 34
MSP PM10 34
BIPM1043.4
BIPM1043.4
PM10
mn/max 24. 1/90.8
45.5
43
PM10 - 33 med
PM10 - 32.9
PM10 - 32.9
PM10-36.5
33.6
PM10-31
PM10 - 30.0
PM10 - 35, Chicago
PM10 -44, LA
PM10 - 41, Phoenix
PM10 - 44, LA
PM10 - 32.9
33.6
PM10-31
PM10 - 32.9
34.3
PM10-31.5
44.81
Co-Pollutants
Measured
03
03
gaseous
pollutants
CO, NO2, O3
SO3
—
SO2, O3, NO2,
CO
SO2, O3, CO
—
—
SO2, O3, NO2,
CO, H+
none
CO
CO
SO2, O3, NO2,
CO
—
SO2, O3, NO2,
CO, H+
SO2, O3, CO
O3, CO
CO, O3, SO2
03
Lag
1
1
0
0
0
0
—
—
0
1
0-1
—
0
3
3
2
2
0
2
0
2
0
1
0
1
1
0-1
—
1
0
Effect Estimate (95% CL)
8.7 (4.6, 13)
6.9(2.7, 11.3)
1.5 (-1.5, 4.6)
3.2 (-0.7, 7.2)
11% (4.0, 18)
3.3(1.7, 5)
5.8(0.5, 11.4)
w/ diabetes: 2.29 (-0.76, 5.44)
w/o diabetes: 1.50(0.42,2.6)
7.4(5.1,9.8)
7.5 (5.3, 9.8)
10.6(7.9, 13.4)
9.4(2.2, 17.1)
w/o prior RI: 8.8(3.3, 14.6)
w/ prior RI: 17.1 (-6.7, 46.9)
No Co Poll: 9.6 (-5. 1,27)
Co Poll: 1.0 (-15, 20)
No Co-Poll: 5.1(0, 10.4)
Co-Poll: 2.5 (-2.5, 7.8)
2% (-0.2, 4.3)
6.1(1.1, 11.3)
6.9 (-4.1, 19.3)
0.6 (-5. 16.7)
Two pollutant model
8.1(6.5,9.7)
6.7(5.3, 8.2)
w/o prior asthma: 11 (7.7, 14.3)
w/o prior asthma: 22.8 (5.1, 43.6)
No Co Poll: 22(8.3,36)
Co Poll: 24(8.2,43)
8.1(6.5,9.7)
6.11(NR)
13.7(5.5,22.6)
16.2(2.0,3.0)
April 2002
8-151
DRAFT-DO NOT QUOTE OR CITE

-------
        TABLE 8-20. SUMMARY OF UNITED STATES PM2 5 RESPIRATORY HOSPITAL
                                     ADMISSION STUDIES
Reference
Lumley and Heagerty
(1999)
Lippmann et al. (2000)
Moolgavkar et al.
(2000)
Outcome
Measures
Respiratory
COPD
COPD
Mean Levels
Mg/in3
PM1;NR
PM25-18
PM25-18.1
Two-Pollutants
Co-Pollutants
none
S02,03,N02,
CO,H+
none
CO
Lag
1
3
3
3
3
Effect Estimate
(95% CL)
5.9(1.1,11.0)
No Poll: 5.5 (-4.7, 17)
Co Poll: 2.8 (-9.2, 16)
6.4(0.9,12.1)
5.6(0.2,11.3)
       Moolgavkar (2000a)     COPD (>64 yrs)   PM25 - 22, LA
                         (median)
                                      PM2.5 - 22, LA
CO
2   5.1(0.9,9.4)

2   2.0 (-2.9, 7.1)
    Two pollutant model
Lippmann et al. (2000)

Sheppard et al. (1999)
Freidman et al. (2001)


Pneumonia

Asthma
Asthma


PM2.5

PM2.5
PM25
(36.7

-18

-16.7

- 30.8 decrease)

S02,03,N02,
CO,H+
CO, O3, SO2
03


1
1
1
3d.
cu
m
No Poll: 13(3.7,22)
Co Poll: 12(1.7,23)
8.7(3.3,14.3)
1.4(0.80-2.48)


        TABLE 8-21. SUMMARY OF UNITED STATES PM10 2 5 RESPIRATORY HOSPITAL
                                     ADMISSION STUDIES
Reference
Moolgavkar (2000a)
Lippmann et al. (2000)

Sheppard et al. (1999)
Outcome
Measures
COPD
COPD
Pneumonia
Asthma
Mean
A*g
PM10.2.5
PM10.2.5
PM10.2.5
PM10.2.5
Levels

-12
-12
-16.2
Two-Pollutant
Co-Pollutants
—
SO2, O3, NO2, CO, H+
S02,03,N02,CO,H+
CO, O3, SO2
Lag
3
3
3
3
3
1
Effect Estimates
(95% CL)
5.1% (-0.4, 10.9)
No Poll: 9. 3 (-4.4, 25)
Co Poll: 0.3 (-14, 18)
No Poll: 9.3 (-4.4, 25)
Co Poll: 0.3 (-14, 18)
11.1 (2.8,20.1)
1     air pollution.  Thus, these results support the hypothesis that considering only hospital
2     admissions and emergency hospital visit effects may greatly underestimate the numbers of
3     medical visits occurring in a population as a result of acute ambient PM exposure.  Those groups
4     identified in these morbidity studies as most strongly affected by PM air pollution are older
5     adults and the very young.
      April 2002                              8-152        DRAFT-DO NOT QUOTE OR CITE

-------
TABLE 8-22. SUMMARY OF UNITED STATES PM1(1, PM? „ and PM1(1,. ASTHMA MEDICAL VISIT STUDIES
                                          10,    2 5,
                                                       10 2 5
3.
to
o
to









H
O'
0
2;
O
H
0
0
H
W
O
O
H
W

Outcome Mean Levels Co-Pollutants
Reference Measures (/-tg/m3) Measured
PM10
Choudhury et al. (1997) Asthma PM10-41.5 Not considered
Lipsett et al. (1997) Asthma PM10 - 61.2 NO2, 03
Norris et al. (1999) Asthma PM10 - 21.7 CO, SO2, NO2
PM10
Norris et al. (2000) Asthma PM10 - Spokane 27.9 MP
PM10- Seattle 21. 5 MP
Tolbert et al. (2000b) Asthma PM10-38.9 O3

Tolbert et al. (2000a) Asthma PM10 -29.1 NO2, 03, CO, SO2
PM25
Norris et al. (1995) Asthma PM2 5 - 12.0 CO, SO2, NO2

Tolbert et al. (2000a) Asthma PM2 s - 19 A NO2, O3, CO, SO2
PM10.2.5
Tolbert et al. (2000a) Asthma PM10.2 5 - 9.39 NO2, O3, CO, SO2

*SP = single pollutant model; MP = multipollutant model.












Lag

0
2
1
1
3
3
1
1
0-2

1
1
0-2

0-2













Effect Estimate
(95% CL)

20.9(11.8,30.8)
34.7(16, 56.5) at 20 °C
SP 75. 9 (25.1, 147.4)
MP 75. 9 (16.3, 166)
2.4 (-10.9, 17.6)
56.2(10.4, 121.1)
SP 13.2 (1.2, 26.7)
MP 8.2 (-7.1, 26.1)
18.8 (-8.7, 54.4)

SP 44.5 (21.7, 71.4)
MP 51.2 (23.4, 85.2)
2.3 (-14.8, 22.7)

21.1 (-18.2, 79.3)













-------
               Tolbert et al. (2000) Atlanta -
               Morris et al. (2000) Seattle -
              Morris et al. (2000) Spokane -
               Morris etal. (1999) Seattle -
         Choudhury et al. (1997) Anchorage -
         Nauenberg and Basu (1999) LA.CA -
             Sheppard etal. (1999) Seattle -
           Zanobetti et al. (2000a) Chicago -
          Samet et al. (2000a) 14 US Cities -
              Moolgavkar (2000b) Phoenix -
               Moolgavkar (2000b) LA,CA -
              Moolgavkar (2000b) Chicago -
            Moolgavkar et al. (2000) King C -
          Moolgavkar et al. (1997) Minn-SP -
             Moolgavkar et al. (1997) Birm. -
              Chen  et al. (2000) Reno.NV -
           Zanobetti et al. (2000a) Chicago -
          Samet etal. (2000a) 14 US Cities -







_

h


Asthma Visits





1 * 1
Asthma Hospital Admissions
^
m
i « i
+i COPD Hospital Admissions
l-»-H
«H
1 A 1
w Pneumonia Hospital Admissions
                                  -25
25       50       75      100
      Excess Risk, %
                     125
150
        Figure 8-13.  Maximum excess risk of respiratory-related hospital admissions and visits
                      per 50-jUg/m3 PM10 increment in selected studies of U.S. cities.
 1      8.3.3  Effects of Particulate Matter Exposure on Lung Function and
 2             Respiratory Symptoms
 3           In the 1996 PM AQCD, the available respiratory disease studies used a wide variety of
 4      designs examining pulmonary function and respiratory symptoms in relation to PM10.  The
 5      models for analysis varied and the populations included several different subgroups. Pulmonary
 6      function studies were suggestive of short term effects resulting from ambient PM exposure. Peak
 7      expiratory flow rates showed decreases in the range of 2 to 5 1/min resulting from an increase of
 8      50 //g/m3 in 24-h PM10 or its equivalent, with somewhat larger effects in symptomatic groups
 9      such as asthmatics.  Studies using FEVj or FVC as endpoints showed less consistent effects.  The
10      chronic pulmonary function studies were less numerous than the acute studies, and the results
11      were inconclusive.
        April 2002
8-154
DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.3.3.1 Effects of Short-Term Particulate Matter Exposure on Lung Function and
 2             Respiratory Symptoms
 3           The available acute respiratory symptom studies discussed in the 1996 PM AQCD included
 4      several different endpoints, but typically presented results for: (1) upper respiratory symptoms,
 5      (2) lower respiratory symptoms, or (3) cough.  These respiratory symptom endpoints had similar
 6      general patterns of results.  The odds ratios were generally positive, the 95% confidence intervals
 7      for about half of the studies being statistically significant (i.e., the lower bound exceeded 1.0).
 8           The earlier studies of morbidity health outcomes of PM10 exposure on asthmatics were
 9      limited in terms of conclusions that could be drawn because of the few available studies on
10      asthmatic subjects. Lebowitz et al. (1987) reported a relationship with TSP exposure and
11      productive cough in a panel of 22 asthmatics but not for peak flow or wheeze.  Pope et al. (1991)
12      studied respiratory symptoms in two panels of asthmatics in the Utah Valley.  The 34 asthmatic
13      school children panel yielded estimated odd ratios of 1.28 (1.06, 1.56) for lower respiratory
14      illness (LRI) and the second panel of 21 subjects aged 8 to 72 for LRI of 1.01  (0.81, 1.27) for
15      exposure to PM10.  Ostro et al. (1991) reported no association for PM25 exposure in a panel of
16      207 adult asthmatics in Denver;  but, for a panel of 83 asthmatic children age 7 to 12 in central
17      Los Angeles, reported a relationship of shortness of breath to O3 and PM10, but could not separate
18      effects of the two pollutants (Ostro et al., 1995). These few studies did not indicate a consistent
19      relationship for PM10 exposure and health outcome in asthmatics.
20           Numerous new studies of short-term PM exposure effects on lung  function and respiratory
21      symptoms published since 1996 were identified by an ongoing medline search.. Most of these
22      followed a panel of subjects over one or more periods and evaluated daily lung function and/or
23      respiratory symptom associations with changes in ambient PM10, PM10_2  5, and/or PM2 5. Lung
24      function was usually measured daily with many studies including forced expiratory volume
25      (FEV), forced vital capacity (FVC) and peak expiratory flow rate (PEF).  Most analyses included
26      both morning and afternoon measurements. A variety of respiratory symptoms were measured,
27      including cough, phlegm, difficulty breathing, wheeze, and bronchodilator use. Finally, several
28      measures of airborne particles were used, including: PM10, PM25, PM10_25, ultrafme PM, TSP,
29      BS, and sulfate fraction of ambient PM.
30           These various studies are summarized in several tables presented in Appendix 8B.  Data on
31      physical and chemical aspects of ambient PM levels (especially for PM10, PM10_2 5. PM2 5, and

        April 2002                                8-155        DRAFT-DO NOT QUOTE OR CITE

-------
 1      smaller size fractions) are of particular interest, as are new studies examining health outcome
 2      effects and/or exposure measures not studied as much in the past. Each table is organized by
 3      study location, PM measure, etc. Where possible, results are presented in terms of the units
 4      described earlier. Specific studies were selected for summarization based on the following
 5      criteria:
 6           • Peak flow was used as the primary lung function measurement of interest.
 7           • Cough, phlegm, difficulty breathing, wheeze, and bronchodilator use were summarized as
 8            measures of respiratory symptoms when available.
 9           • Quantitative relationships were estimated using PM10, PM2 5, PM10_2 5, and/or smaller PM
10            as independent variables.
11           • The analysis of the study was done such that each individual served as their own control.
12      Other factors are discussed earlier in Section 8.1.3 of this chapter selection of Studies for
13      Review, as well  as in Chapter 1.
14
15      8.3.3.1.1  Lung  Function and Respiratory Symptom Effects in Asthmatic Subjects
16           Tables 8B-4 and 8B-5 in Appendix B summarize short-term PM exposure effects on lung
17      function and respiratory symptoms, respectively, in asthmatic subjects. The peak flow analyses
18      results for asthmatics tend to show small decrements for PM10 and PM2 5 as shown in studies to
19      include Gielen et al. (1997), Peters et al. (1997b), Romieu et al. (1997), and Pekkanen et al.
20      (1997) listed in summary Table 8-23 for PM10, and Table 8-24 for PM25, and in more detail in
21      Appendix 8B, Table 8B-4. Pekkanen et al. (1997) reported similar changes in peak flow to be
22      related to several sizes of PM with PN 0.032-0.10 -0.970 (0.502) l(cm3) and PM10.32 -0.901
23      (0.536) and PM10 -1.13 (0.478) for morning PEF lag 2.  Peters et al. (1997c) report the strongest
24      effects on peak flow were found with ultrafme particles. PMMC001.0 x: -1.21 (-2.13, -0.30);
25      PMMC001.25: -1.01 (-1.92, -0.11); and PM10, -1.30 (-2.36, -0.24).
26           Penttinen et al. (2001) using biweekly spirometry over 6 months on a group of 54 adult
27      asthmatics found that FVC, FEVj, and spirometric PEFR were inversely, but mostly
28      nonsignificantly-associated with ultra fine particle concentrations. Compared to the effect
29      estimates for self-monitored PEFR, the effect estimates for spirometric PEFR tended to be larger.
30      The strongest associations were observed in the size range of 0.1 to 1 //m.
31

        April 2002                                8-156       DRAFT-DO NOT QUOTE OR CITE

-------
TABLE 8-23. SUMMARY OF ASTHMA PM  PFT STUDIES
                                 10
3.
to
o
o
to









-------
TABLE 8-23 (cont'd). SUMMARY OF ASTHMA PM10 PFT STUDIES
3.
to
o
o
to










-------
 1           In the Uniontown reanalysis, peak flow for PM21 for a 14 //g/m3 increment was -0.91 1/m
 2      (-1.14, -1.68) and PM10.2A for 15 //g/m3 +1.04 1/m (-1.32, +3.4); for State College PM2A -0.56
 3      (-1.13, +0.01) andPM10.21 -0.17 (-2.07, +1.72). The Schwartz andNeas (2000) reanalyses
 4      allows comparison of fine and coarse effects using two pollutant models for fraction of PM.
 5           Coull et al. (2001) reanalyzed data from the Pope et al. (1991) study of PM effects on
 6      pulmonary function of children in the Utah Valley, using additive mixed models which allow for
 7      assessment of heterogeneity of response or the source of heterogeneity. These additive models
 8      describe complex covariate effects on each child's peak expiratory flow while allowing for
 9      unexplained population heterogeneity and serial correlation among repeated measurements.  The
10      analyses indicates that there is heterogeneity in that population with regard to PM10 (i.e.,
11      specifically that there are three subjects in the Utah Valley study who exhibited  a particularly
12      acute response to PM10).  However the limited demographic data available in the Utah  Valley
13      Study does not explain the heterogeneity in PM sensitivity among the school children population.
14           The peak flow analyses results for asthmatics tend to show small decrements for  both PM10
15      and PM2 5. For PM10, the available point estimates for morning PEF lagged one day showed
16      decreases, but the majority of the studies were not statistically significant, see Table 8-22 and as
17      shown in Figure 8-14 as an example of PEF outcomes. Lag 1  may be more relevant for morning
18      measurement of asthma outcome from the previous day.  The figure presents studies which
19      provided such data. The results were consistent for both AM and PM peak flow analyses. The
20      effects using 2 to five-day lags averaged about the same as did the zero to one-day lags, but the
21      effects had wider confidence limits.  Similar results were found for the PM2 5 studies, although
22      there were fewer studies. Several studies included PM2 5 and PM10 independently in their
23      analyses of peak flow. Of these, Naeher et al. (1999), Tiittanen et al. (1999), Pekkanen et al.
24      (1997), and Romieu et al. (1996) all  found similar results for PM25 and PM10. The study of
25      Peters et al. (1997c) found slightly larger effects for PM25. The study of Schwartz and Neas
26      (2000) found larger effects for fine particle measures (PM2 5, sulfate,  etc.) than for the coarse
27      fraction. Naeher et al. (1999) found that FT was significantly related to a decrease in morning
28      PEF. Overall,  then,  PM10 and PM2 5 both appear to affect lung  function in asthmatics, but there is
29      only limited evidence for a stronger effect of fine versus  coarse fraction particles. Also, of the
30      studies provided, few if any analyses were able to separate out the effects of PM10 and PM2 5 from
31      other pollutants.

        April 2002                                8-159       DRAFT-DO NOT QUOTE OR CITE

-------
                                                                Romleuetal. (1996)
                                                                *     (Mexico)
                                                                Pekkanenetal. (1997)
                                                                      (Finland)
                                                                Romleu et al. (1996)
                                                                     (Mexico)
                                                                 Gielenetal, (1937)
                                                                   (Netherlands)
                -10,0
-5.0              o.O              5-°
  Change In Pulmonary Function, L/min
                                                                                  10.0
       Figure 8-14.  Selected acute pulmonary function change studies of asthmatic children.
                     Effect of 50 Aig/m3 PM10 on morning Peak flow lagged one-day.
 1           The effects of PM on respiratory symptoms in asthmatics tended to be positive, although
 2     they were much less consistent than the effects on lung function. Vedal et al. (1998) reported
 3     that increases in PM10 were associated with increased reporting of cough,  phlegm production, and
 4     sore throat and that children with diagnosed asthma are more susceptible to the effects than are
 5     other children.  Similarly, in the Gielen et al. (1997) study of a panel of children, most of whom
 6     had asthma, low levels  of PM increased symptoms and medication use. Peters et al. (1997c)
 7     study of asthmatics examined particle effects by size which indicated that fine particles were
 8     associated with increases in cough, of which MC 0.01-2.5 was the best predictor.
 9           Delfino et al. (1998) used an asthma symptom score to evaluate the  effect of acute pollutant
10     exposures. PM10 1-and 8-hr maximum had larger effects than the 24-hr mean. Subgroup
11     analyses showed effects of current day PM maximums were strongest in 10 more frequently
12     symptomatic children;  the odds ratios for adverse symptoms from  90th percentile increases were
13     2.24 (1.46, 3.46), for l-hrPM10; 1.82 (1.18, 2.8), for 8-hr PM10, and 1.50 (0.80-2.80) for 24-hr
14     PM10. Analyses suggested that effects of O3 and PM10 were largely independent.
15           Romieu et al. (1996) found children with mild asthma to be more strongly affected by high
16     ambient levels of PM observed in northern Mexico City than in a study (Romieu et al., 1997)
       April 2002
               8-160
DRAFT-DO NOT QUOTE OR CITE

-------
 1      conducted in a nearby area with lower PM10 levels (mean PM10 = 166.8 //g/m3 versus 54.2
 2      Mg/m3). Yu et al. (2000) reported estimates of odd ratios for asthma symptoms and 10 //g/m3
 3      increments in PM10 and PM10 values of 1.18(1.05, 1.33) and 1.09(1.01, 1.18), respectively.
 4      Multipollutant models with CO and SO2 yielded for PM10, 1.06 (0.95, 1.19) for PM10, and 1.11
 5      (0.98, 1.26) for PMX 0, thus showing a lower value for PMX 0 and a lower value for PM10 with a
 6      loss of significance. The correlation between CO and PMl 0 and PM10 was 0.82 and 0.86. Ostro
 7      et al. (2001) studied a panel of inner-city African American children using a GEE model with
 8      several measures of PM, including PM10 (both 24-hour average and 1-hour max.) and PM25,
 9      demonstrating positive associations with daily probability of shortness of breath, wheeze, and
10      cough.
11           Most studies showed increases in cough, phlegm, difficulty breathing, and bronchodilator
12      use, although these increases were generally not statistically significant for PM10 (see
13      Tables 8-25, 8-26, 8-27, and 8-28; and, for cough as an example, see  Figure 8-15). For PM25
14      results, see Table 8-29. Several studies included two indicators for PM; PM10_2 5 or PM10 and
15      PM25 in their analyses. The studies of Peters  et al. (1997c) and Tiittanen et al. (1999) found
16      similar effects for the two PM measures, whereas the Romieu et al. (1996) study found slightly
17      1 arger effects for PM2 5.
18
19      8.3.3.1.2  Lung Function and Respiratory Symptom Effects in Nonasthmatic Subjects
20           Results of the PM10 peak flow analyses in non-asthmatic studies (see Appendix 8B,  Table
21      8B-6) were inconsistent, with fewer studies reporting results in the same manner as for the
22      asthmatic studies (see Table 8-30). Many of the point estimates showed increases rather than
23      decreases. Similar results were found in the PM2 5 studies (see Summary Table 8-31). The
24      effects on respiratory symptoms in non-asthmatics (see Appendix  8B, Table  8B-7) were similar
25      to those in asthmatics  (see Table 8-32). Most studies showed that PM10 increases cough, phlegm,
26      difficulty breathing, and bronchodilator use, although these increases were generally not
27      statistically significant. For PM2 5 see Tables  8-32 and for PM coarse studies see Table 8-33.
28      The Schwartz and Neas (2000) reanalyses allows comparison of fine  and coarse particle effects
29      on healthy school children using two pollutant models of fine and coarse PM. CM was estimated
30      by subtracting PM21 from PM10 data.  They report for cough for reanalysis of the Harvard Six
31      City Diary Study in the two PM pollutant model PM2 5 (increment 15  //g/m3)  OR = 1.07 (0.90,

        April 2002                                8-161        DRAFT-DO NOT QUOTE OR CITE

-------
TABLE 8-25. SUMMARY OF ASTHMA PM1(1 COUGH STUDIES
                                   10
3.
to
o
to











-------
                          TABLE 8-26. SUMMARY OF ASTHMA PM1(1 PHLEGM STUDIES
                                                            10
3.
to
o
o
to







Reference citation,
location, etc.
Vedal et al. (1998)
Peters etal. (1997b)
Romieuetal. (1997)
Romieuetal. (1996)
Vedal et al. (1998)
Romieuetal. (1997)
Romieuetal. (1996)
Peters etal. (1997b)
Outcome Measure
OR Phlegm
OR Phlegm
OR Phlegm
OR Phlegm
OR Phlegm
OR Phlegm
OR Phlegm
OR Phlegm
Mean Paniculate
Levels (Range) ^g/m3
19.1 (1, 159)
47 (29, 73)
(12, 126)
166.8 (29, 363)
19.1 (1, 159)
(12, 126)
166.8 (29, 363)
47 (29, 73)
Co-Pollutants
Measured
None
SO2, sulfate, H+
Ozone
Ozone
None
Ozone
Ozone
SO2, sulfate, H+
Lag
Structure
Oday
Oday
Oday
Oday
2 day
2 day
2 day
1-5 day
Effect measures
50 ,wg/n
1.28 (0.86, 1.89)
1.13(1.04, 1.23)
1.05 (0.83, 1.36)
1.21 (1.00, 1.48)
1.40(1.03, 1.90)
1.00(0.86, 1.16)
1.16(0.91, 1.49)
1.17(1.09, 1.27)
standardized to
13PM10








oo
              TABLE 8-27. SUMMARY OF ASTHMA PM10 LOWER RESPIRATORY ILLNESS (LRI) STUDIES
00




o
§
H
6
0
o
H
0
0
H
W
O
HH
7°
o
H
W

Reference citation,
location, etc.
Vedal et al. (1998)
Gielen etal. (1997)
Romieuetal. (1997)
Romieuetal. (1996)

Vedal et al. (1998)
Gielen etal. (1997)
Segala etal. (1998)
Romieuetal. (1997)
Romieuetal. (1996)

Delfinoetal. (1998)






Outcome Measure
LRI
LRI
LRI
LRI

LRI
LRI
LRI
LRI
LRI

LRI





Mean Paniculate
Levels (Range)
19.1(1, 159)
30.5 (16, 60)
(12, 126)
166.8 (29, 363)

19.1(1, 159)
30.5 (16, 60)
34.2 (9, 95)
(12, 126)
166.8 (29, 363)

24 h 26 (6, 51)
8-h43 (23-73)
1-h 57 (30-108)



Co-pollutants
Measured
None
Ozone
Ozone
Ozone

None
Ozone
SO2, NO2
Ozone
Ozone

Ozone
Ozone
Ozone



Lag
Structure
Oday
Oday
Oday
Oday

2 day
2 day
2 day
2 day
2 day

Oday
Oday
Oday



Effect measures standardized to
50 ,wg/m3 PM10
1.10(0.82, 1.48)
1.26 (0.94, 1.68)
1.00 (0.95, 1.05)
1.21(1.10, 1.42)

1.16(1.00, 1.34)
1.05 (0.74, 1.48)
1.66 (0.84, 3.30)
1.00 (0.93, 1.08)
1.10(0.98, 1.24)

1.47(0.90-2.39)
2.17(1.33 -3.58)
1.78(1.25-2.53)



-------
                             TABLE 8-28.  SUMMARY OF ASTHMA PM1(1 BRONCHODILATOR USE STUDIES
3.
to
o
o
to
                                                                                10
Reference citation,
location, etc.
Outcome Measure
  Mean Paniculate
Levels (Range)
Co-pollutants
 Measured
               Effect measures standardized to
Lag Structure           50 Atg/m3 PM10
         Gielenetal. (1997)
         Hiltermann et al. (1998)
         Peters etal. (1997b)
         Gielenetal. (1997)
         Hiltermann et al. (1998)
         Peters etal. (1997b)
                         OR bronchodilator use
                         OR bronchodilator use
                         OR bronchodilator use
                         OR bronchodilator use
                         OR bronchodilator use
                         OR bronchodilator use
                     30.5 (16, 60)
                     39.7(16,98)
                     47 (29, 73)
                     30.5 (16, 60)
                     39.7(16,98)
                     47 (29, 73)
                      Ozone                0 day
                      Ozone, NO2, SO2      0 day
                      SO2, sulfate, H+        0 day
                      Ozone                2 day
                      Ozone, NO2, SO2      1-7 day
                      SO2, sulfate, H+	1-5 day
                                 0.94 (0.59, 1.50)
                                 1.03 (0.93, 1.15)
                                 1.06 (0.88, 1.27)
                                 2.90(1.81,4.66)
                                 1.12(1.00, 1.25)
                                 1.23 (0.96, 1.58)
oo
H
6
o
o
H
O
O
H
W
O
O
H
W

-------
                                                         Glelenetal. (1997)
                                                            (Netherlands)
                                              Romieu et al, (1997)
                                                   (Mexico)
                                              Peters etal. (1997a)
                                               (Czech Republic)
                                                         Vedal etal. (1998)
                                                             (Canada)
             0.5
                        1.0               2.0
                                  Odds Ratio for Cough
      4.0
8.0
       Figure 8-15.  Odds ratios with 95% confidence interval for cough per 50-jUg/m3 increase in
                     PM10 for selected asthmatic children studies at lag 0.
 1
 2
 3
 4
 5
 6
 1
 8
 9
10
11
12
13
14
15
1.26) and PM10_25 (increment 8 //g/m3) OR 1.18 (1.04, 1.34) in contrast to lower respiratory
symptom results of PM25 OR 1.29 (1.06, 1.57) andPM10.25 1.05 (0.9, 1.23).
     Jalaludin et al.  (2000) analyses using a multipollutant model evaluated O3, PM10, and NO2.
They found in metropolitan Sydney that ambient ozone and PM10 concentrations are poorly
correlated (0.13). For PEFR PM10 only was 0.0045 (0.0125) p-0.72, and with O3, 0.0051
(0.0124), p-0.68. Ozone was also unchanged in the one- and two-pollutant models. Gold et al.
(1999) attempted to study the interaction of PM2 5 and ozone on PEF. The authors found
independent effects of the two pollutants, but found that the joint effect was slightly less than the
sum of the independent effects.
     Three authors,  Schwartz and Neas (2000), Tiittanen et al.  (1999) and Neas et al. (1999),
used PM10_2 5 as a coarse fraction particulate measure. Schwartz and Neas (2000) found that PM10
was significantly related to cough. Tiittanen found that one day lag of PM10_25 was related to
morning PEF, but there was no effect on evening PEF.  Neas et al. found no effects of PM10_2 5 on
PEF.
       April 2002
                                         8-165
DRAFT-DO NOT QUOTE OR CITE

-------
         TABLE 8-29.  SUMMARY OF ASTHMA PM2< RESPIRATORY SYMPTOM STUDIES
         Reference                       Mean Participate                              Effect measures
         citation,           Outcome      Levels (Range)    Co-pollutants      Lag      standardized to
         location, etc.        Measure          Mg/ni3          Measured     Structure    25 /-tg/m3 PM25
         Peters et al.
         (1997b)

         Romieu et al.
         (1996)

         Tittanen et al.
         (1999)

         Romieu et al.
         (1996)

         Tittanen et al.
         (1999)

         Ostro et al.
         (2001)

         Peters et al.
         (1997b)

         Romieu et al.
         (1996)

         Romieu et al.
         (1996)

         Romieu et al.
         (1996)

         Romieu et al.
         (1996)
 OR cough


 OR cough


 OR cough


 OR cough


 OR cough


 OR cough


 OR cough


OR Phlegm


OR Phlegm


 ORLRI


 ORLRI
 50.8 (9, 347)     SO2, sulfate, H+     0 day
85.7 (23, 177)


  15 (3, 55)


85.7 (23, 177)
85.7 (23, 177)
85.7 (23, 177)
85.7 (23, 177)
85.7 (23, 177)
   Ozone
Oday
NO2, SO2, CO,     0 day
    ozone
   Ozone
2 day
  15 (3, 55)       NO2, SO2, CO,     2 day
                    ozone
 40.8 (4, 208)      Ozone, NO2
                 3 day
   Ozone
   Ozone
   Ozone
   Ozone
1.22(1.08, 1.38)


1.27(1.08, 1.42)


1.04 (0.86, 1.20)


1.16(0.98, 1.33)


1.24(1.02, 1.51)


1.02 (0.98, 1.06)
 50.8(9,347)     SO2, sulfate, H+    1-5 day     1.02(0.90,1.17)
Oday      1.21(0.98,1.48)
2 day      1.16(0.99,1.39)
Oday      1.21(1.05,1.42)
2 day      1.16(1.05,1.42)
 1      8.3.3.2  Long-Term Particulate Matter Exposure Effects on Lung Function and
 2              Respiratory Symptoms

 3      8.3.3.2.1 Summary of the 1996 Particulate Matter Air Quality Criteria Document Key
 4               Findings

 5           In the 1996 PM AQCD, the available long-term PM exposure-respiratory disease studies

 6      were limited in terms of conclusions that could be drawn.  At that time, three studies based on a

 7      similar type of respiratory symptom questionnaire administered at three different times as part of

 8      the Harvard Six-City and 24-City Studies provided data on the relationship of chronic respiratory

 9      disease to PM. All three studies suggest a long-term PM exposure effect on chronic respiratory

10      disease. The analysis of chronic cough, chest illness and bronchitis tended to be significantly

11      positive for the earlier surveys described by Ware et al. (1986) and Dockery et al. (1989). Using
        April 2002
                        8-166
                      DRAFT-DO NOT QUOTE OR CITE

-------
p>
us.
to
O
o
to










10% AM PEFR Deer.
OR >10% AM PEFR Deer.
OR >10% AM PEFR Deer.
Morning PEFR
% change in morning PEFR
Evening PEFR
Evening PEFR
Evening PEFR

Evening PEFR
Evening PEFR
Evening PEFR

Evening PEFR
OR >10% PM PEFR Deer.

OR >10%PM PEFR Deer.

OR >10%PM PEFR Deer.

OR >10%PM PEFR Deer.

OR >10%PM PEFR Deer.

OR >10%PM PEFR Deer.

% change in evening PEFR



Mean Particulate
Levels (Range) /-ig/m3

51 (23,878)
28 (5, 122)
32
28 (5, 122)
42 (5, 146)
42 (5, 146)
42 (5, 146)
32
(not given)
32
(not given)
(not given)

28 (5, 122)
28 (5, 122)
51(23,878)

32
42 (5, 146)

42 (5, 146)

42 (5, 146)

34 (?, 106)

34 (?, 106)

34 (?, 106)

(not given)



Co-pollutants
Measured

Ozone
NO2, SO2, CO, ozone
Ozone
NO2, SO2, CO, ozone
NO2, SO2
NO2, SO2
NO2, SO2
Ozone
NO2, SO2, CO
Ozone
Sulfate fraction
Sulfate fraction

NO2, SO2, CO, ozone
NO2, SO2, CO, ozone
Ozone

Ozone
NO2, SO2

N02, S02

N02, S02

NO2, SO2, sulfate

NO2, SO2, sulfate

NO2, SO2, sulfate

N02, S02, CO



Lag Structure


Iday
Oday
1-5 day
1-4 day
Iday
2 day
1-5 day
Oday
1 day
Oday
Oday
Oday

Oday
Oday
Oday

1-5 day
Oday

2 day

1-5 day

Oday

2 day

1-5 day

1 day



Effect measures standardized to
50 Mg/m3 PM10

-0.20 (-0.47, 0.07)
1.21 (-0.43,2.85)
2. 64 (-6.56, 11.83)
-1.26 (-5. 86, 3.33)
1.04(0.95, 1.13)
1.02(0.93, 1.11)
1.05(0.91, 1.21)
-8.16(-14.81,-1.55)
0.07 (-0.50, 0.63)
-1.44 (-7.33, 4.44)
-1.52 (-2. 80, -0.24)
-0.93 (-1.88, 0.01)

0.72 (-0.63, 1.26)
2.33 (-2.62, 7.28)
-0.14 (-0.45, 0.17)

1.47 (-7.31, 10.22)
1.17(1.08,1.28)

1.08(0.99,1.17)

1.16(1.02,1.33)

1.44(1.02,2.03)

1.14(0.83,1.58)

1.16(0.64,2.10)

-0.22 (-0.57, 0.16)




-------
TABLE 8-31. SUMMARY OF NON-ASTHMA PM1(1 RESPIRATORY SYMPTOM STUDIES
                                       10
3.
to
o
o
to











oo
1
oo




O
H
1
O
0
o
H
0
0
H
W
O
O
H
w
Reference citation,
location, etc.
Schwartz & Neas (2000)

Boezenetal. (1998)
Van der Zee etal. (1999)
Urban areas
Tittanen etal. (1999)
Van der Zee etal. (1999)
Urban areas
Van der Zee etal. (1999)
Urban areas
Tittanen etal. (1999)
Boezenetal. (1998)
Tittanen etal. (1999)

Schwartz & Neas (2000)
Van der Zee etal. (1999)
Urban areas
Van der Zee etal. (1999)
Urban areas
















Outcome Measure
OR cough-no
other symptoms
OR cough
OR cough

OR cough
OR cough

OR cough

OR cough
OR Phlegm
OR Phlegm

LRI
LRI

LRI
















Mean Paniculate Levels
(Range) ^g/m3
(not given)

42 (5, 146)
34 (?, 106)

28 (5, 122)
34 (?, 106)

34 (?, 106)

28 (5, 122)
42 (5, 146)
28 (5, 122)

(not given)
34 (?, 106)

34 (?, 106)
















Co-pollutants
Measured
Sulfate fraction

NO2, SO2
NO2, SO2, sulfate

NO2, SO2, CO, ozone
NO2, SO2, sulfate

NO2, SO2, sulfate

NO2, SO2, CO, ozone
NO2, SO2
NO2, SO2, CO, ozone

Sulfate fraction
NO2, SO2, sulfate

NO2, SO2, sulfate
















Lag
Structure
Oday

Oday
Oday

Oday
2 day

1-5 day

1-4 day
Oday
2 day

Oday
Oday

2 day
















Effect measures standardized to
50 mg/m3 PM10
1.20 (1.07, 1.35)

1.06(0.93, 1.21)
1.04(0.95, 1.14)

1.00(0.87, 1.16)
0.94 (0.89, 1.06)

0.95(0.80, 1.13)

1.58(0.87,2.83)
1.11(0.91, 1.36)
Positive but not significant


0.98 (0.89, 1.08)

1.01(0.93, 1.10)

















-------
                 TABLE 8-32. SUMMARY OF NON-ASTHMA PM  RESPIRATORY OUTCOME STUDIES
                                                      2 5
3.
to
o
o
to









-------
TABLE 8-33.  SUMMARY OF NON-ASTHMA COARSE FRACTION STUDIES OF RESPIRATORY ENDPOINTS
3.
to
o
o
to










-------
 1      a design similar to the earlier one, Dockery et al. (1996) expanded the analyses to include
 2      24 communities in the United States and Canada.  Bronchitis was found to be higher (odds ratio
 3      =1.66) in the community with the highest particle strong acidity when compared with the least
 4      polluted community.  Fine particulate sulfate was also associated with higher reporting of
 5      bronchitis (OR = 1.65, 95% CI1.12, 2.42).
 6           Interpretation of such studies requires caution in light of the usual difficulties ascribed to
 7      cross-sectional studies.  That is, evaluation of PM effects is based on variations in exposure
 8      determined by a different number of locations. In the first two studies, there were six locations
 9      and, in the third, twenty-four.  The results  seen in all studies were consistent with a PM gradient,
10      but it was impossible to separate out effects of PM and any other factors or pollutants having the
11      same gradient.
12           Chronic pulmonary function studies by Ware et al. (1986), Dockery et al. (1989), and Neas
13      et al. (1994) had good monitoring data and well-conducted standardized pulmonary function
14      testing over many years,  but showed no effect for children from airborne particle pollution
15      indexed by TSP, PM15, PM25 or sulfates. In contrast, the Raizenne et al. (1996) study of U.S. and
16      Canadian children found significant associations between FEVj and FVC and acidic particles
17      (FT).  Overall, the available studies provided only limited evidence suggestive of pulmonary lung
18      function decrements being associated with chronic exposure to PM indexed by various measures
19      (TSP, PM10, sulfates, etc.).  However, it was noted that cross-sectional studies require very large
20      sample sizes to detect differences because they cannot eliminate person to person variation,
21      which is much larger than the within person variation.
22
23      8.3.3.2.2  New Studies of Long- Term Particulate Matter Exposure Respiratory Effects
24           Several  studies have been published  since 1996 which evaluate effects of long-term PM
25      exposure on lung function and respiratory illness, as summarized in Appendix 8B, Table 8B-8.
26      The new  studies examining PM10 and PM25 in the United States include McConnell et al. (1999),
27      Abbey et al. (1998), Berglund et al. (1999), Peters et al. (1999a,b), Gauderman et al. (2000), and
28      Avol et al. (2001), which all examined effects in California cohorts but produced inconsistent
29      results. McConnell et al. (1999) noted that as PM10 increased across communities, an increase in
30      bronchitis risk per interquartile range also  occurred, results consistent with those reported by
31      Dockery et al. (1996), although the high correlation of PM10,  acid, and NO2 precludes clear

        April 2002                                 8-171        DRAFT-DO NOT QUOTE OR CITE

-------
 1      attribution of the McConnell et al. bronchitis effects specifically to PM alone. Avol et al. (2001)
 2      reported that, for 110 children that moved to other locations as a group, subjects who moved to
 3      areas of lower PM10 showed increased growth in lung function and subjects who moved to
 4      communities with higher PM10 showed slowed lung function growth.
 5           For non-U.S. studies, particularly interesting results were obtained by Leonard! et al. (2000)
 6      as part of the Central European Air Quality and Respiratory Health (CESAR) study. Blood and
 7      serum samples were collected from school children aged 9-11 yrs. in each of 17 communities in
 8      Central Europe (N = 10 to 61 per city). Numbers of lymphocytes increased as PM concentrations
 9      increased across the cities. Regression slopes, adjusted for confounder effects, were largest and
10      statistically significant for PM2 5, but small and non-significant for PM10_25.  A similar positive
11      relationship was found between IgG concentration in serum and  PM2 5 gradient, but not for PM10
12      or PM10_2 5. These results tend to suggest a PM effect on immune function more strongly due to
13      ambient fine  particle than coarse particle exposure.
14           Other non-U.S. studies examined PM measures such as TSP and BS in European countries.
15      In Germany,  Heinrich et al. (2000) reported a cross-sectional survey of children, conducted twice
16      (with the same 971 children included  in both surveys).  TSP levels decreased between surveys as
17      did the prevalence of all respiratory symptoms (including bronchitis).  Also, Kramer et al. (1999)
18      reported a study in six East and West  Germany communities, which found yearly decreasing TSP
19      levels to be related to ever-diagnosed  bronchitis from 1991-1995. Lastly, Jedrychowski et al.
20      (1999) reported an association between both BS and SO2 levels in various areas of Krakow,
21      Poland, and slowed lung function growth (FVC and FEVj).
22
23      8.3.3.2.3  Summary of Long- Term Particulate Matter Exposure Respiratory Effects
24           The methodology used in the long-term studies varies much more than the methodology in
25      the short-term studies.  Some studies reported highly significant  results (related to PM) while
26      others reported no  significant results.  The cross-sectional studies are often confounded, in part,
27      by unexplained differences between geographic regions. The studies that looked for a time trend
28      are also confounded by other conditions that were changing over time.  Probably the most
29      credible cross-sectional study remains that described by Dockery et al. (1996) and Raizenne et al.
30      (1996). This study, reported in the previous 1996 PM AQCD, found differences in peak flow
31      and bronchitis rates associated with fine particle strong acidity.

        April 2002                                8-172       DRAFT-DO NOT QUOTE OR CITE

-------
 1          Newly available studies since the 1996 PM AQCD, overall, provide evidence consistent
 2     with the findings from the above 24-City Study. Most notably, several U.S. and European
 3     studies report associations between PM measures and bronchitis rates and/or lung function
 4     decrements or slowed lung function growth. One also provided evidence of PM effects on
 5     immune function in school children, with stronger associations for fine particle indicators than
 6     for ambient coarse particles.
 7
 8
 9     8.4 DISCUSSION OF EPIDEMIOLOGIC STUDIES ON HEALTH
10          EFFECTS OF AMBIENT PARTICULATE MATTER
11     8.4.1  Introduction
12          Numerous PM epidemiology studies assessed in the 1996 PM AQCD implicated ambient
13     PM as a likely contributor to mortality and morbidity effects associated with ambient air
14     pollution exposures in epidemiology studies.  Since preparation of the last previous PM AQCD
15     in 1996, the epidemiologic evidence concerning ambient PM-related health effects has expanded
16     greatly.  Past regulatory decisions have played an important role in the selection of PM indices
17     and in the evolution of the PM epidemiologic literature base.  The adoption of PM10 standards in
18     1987, and of PM25 standards in 1997, have generated ambient air concentration databases that
19     made it possible for research to address and resolve many of previously unresolved linkages
20     between airborne PM and human health;  and the newly authorized network of speciation
21     samplers holds promise for further advances in the near future on the identification  of the more
22     influential components of the ambient air pollution mixture. The most important types of
23     additions to the database beyond that assessed in the 1996 PM AQCD, as evaluated above in this
24     chapter, are:
25
26     (1) New multi-city studies on a variety of endpoints which provide more precise estimates of the
27     average PM effect sizes than most smaller-scale individual city studies, but also showing much
28     greater heterogeneity among studies than previously observed;
29
       April 2002                               8-173        DRAFT-DO NOT QUOTE OR CITE

-------
 1      (2) More studies of various health endpoints using ambient PM10 and/or closely related mass
 2      concentration indices (e.g., PM13 and PM7), which substantially lessen the need to rely on
 3      non-gravimetric indices (e.g., BS or COH);
 4
 5      (3) New studies evaluating relationships of a variety of endpoints to the ambient PM coarse
 6      fraction (PM10_2 5), the ambient fine-particle fraction (PM2 5), and  even ambient ultrafme particles
 7      measures (PM0 x and smaller) using direct mass measurements and/or estimated from site-specific
 8      calibrations;
 9
10      (4) A few new studies in which the relationship of some health endpoints to ambient particle
11      number concentrations were evaluated;
12
13      (5) Many new studies which evaluated the sensitivity of estimated PM effects to the inclusion of
14      gaseous co-pollutants in the model;
15
16      (6) Preliminary attempts to evaluate the effects of air pollutant combinations or mixtures
17      including PM components, based on empirical combinations (e.g., factor analysis) or source
18      profiles;
19
20      (7) Numerous new studies of cardiovascular endpoints, with particular emphasis on assessment
21      of cardiovascular risk factors as well as symptoms;
22
23      (8) Additional new studies on asthma and other respiratory conditions potentially exacerbated by
24      PM exposure;
25
26      (9) New analyses of lung cancer associations with long-term exposures to ambient PM.
27
28      (10) New  studies of infants and children as a potentially  susceptible population.
29
30           As discussed in Sections 8.2 and 8.3, numerous new PM epidemiology studies, both of
31      short-term and long-term PM exposure, show statistically significant excess risk for various

        April 2002                               8-174        DRAFT-DO NOT QUOTE OR CITE

-------
 1      mortality and/or morbidity endpoints in many U.S. cities and elsewhere to be associated with
 2      ambient PM indexed by a variety of ambient community monitoring methods.
 3           Still, several methodological issues discussed in the 1996 PM AQCD continue to be
 4      important in assessing and interpreting the overall PM epidemiology database and its
 5      implications for estimating risks associated with exposure to ambient PM concentrations in the
 6      United States.  The fundamental issue essentially subsuming all of the other modeling issues is
 7      the selection of an appropriate statistical model. These critical methodological issues are:
 8      (1) potential confounding of PM effects by co-pollutants (especially major gaseous pollutants
 9      such as O3, CO, NO2, SO2);  (2) the attribution of PM effects to specific PM components (e.g.,
10      PM10, PM10_25, PM25, ultrafines, sulfates,  metals, etc.) or source-oriented indicators (motor
11      vehicle emissions, vegetative burning, etc.); (3) the temporal relationship between exposure and
12      effect (lags, mortality displacement, etc.); (4) the general shape of exposure-response
13      relationship(s) between PM and/or other pollutants and observed health effects (e.g., potential
14      indications of thresholds for PM effects); and (5) the consequences of measurement error.
15           Assessing the above issue(s) in relation to the PM epidemiology data base remains quite a
16      challenge. The basic issue is that there are an extremely large number of possible models, any of
17      which may turn out to give the best statistical "fit" of a given set of data, and only some of which
18      can be dismissed a priori as biologically or physically illogical or impossible, except that
19      putative cause clearly cannot follow effect in time. Most of the models for daily time series
20      studies are fitted by adjusting for changes over long time intervals and across season, by day of
21      week, weather, and climate. Many of the temporal and weather variable models have been fitted
22      to data using semi-parametric methods such as spline functions or local regression smoothers
23      (loess).  The goodness of fit of these base models has been evaluated by criteria suitable for
24      generalized linear models with Poisson or hyper-Poisson responses (number of events) with a log
25      link function, particularly the Akaike Information  Criterion (AIC) and the more conservative
26      Bayes or Schwarz information criterion (BIC), which adjust for the number of parameters
27      estimated from the data.  The Poisson over-dispersion index and the auto-correlation of residuals
28      are also often used.  It is  often assumed, but rarely proven,  that the best-fitting models with PM
29      would be models with the largest and most significant PM  indices.  Also, if high correlations
30      between PM and one or more gaseous pollutants emitted from a common source (e.g., motor
31      vehicles) exist in a given area, then disentangling their relative individual partial contributions to

        April 2002                                8-175       DRAFT-DO NOT QUOTE OR CITE

-------
 1      observed health effects associations becomes very difficult.  However, there have been very few
 2      attempts at broad, systematic investigations of the model selection issue and little reporting of
 3      goodness-of-fit criteria among competing models that provide a better basis by which to better
 4      assess or compare models.
 5           One  systemic analysis of model choice was carried out by Clyde et al. (2000), using
 6      Bayesian Model Averaging for the same Birmingham, AL, data as analyzed by Smith et al.
 7      (2000).  Several different calibrated information criterion priors were tried in which models with
 8      large numbers of parameters are penalized to various degrees. After taking out a baseline trend
 9      (estimated using a GLM estimate with a 30-knot thin-plate smoothing spline), 7,860 models were
10      selected for use in model averaging. These included lags 0-3 of a daily monitor PM10, an
11      area-wide  average PM10 value with the same lags, temperature (daily extremes and average)
12      lagged 0-2 days, humidity (dewpoint, relative  humidity min  and max, average specific humidity)
13      lagged 0-2, and atmospheric pressure, lagged 0-2.  The model choice is sensitive to the
14      specification of calibrated information criterion priors, in particular disagreeing as to whether
15      different PM10 variables should be  included or not. For example, one or another PM10 variable is
16      included in all the top 25 Akaike Information  Criterion (AIC) models, but only in about 1/3 of
17      the top Bayes Information Criterion (BIC) models. Both approaches give a relative risk estimate
18      of about 1.05 (to be compared to the Schwartzvalueofl.il fora 100 unit increase), with
19      credibility intervals of (0.94, 1.17)  for the AIC prior and  (0.99, 1.11) for the BIC prior.
20      A validation study in which randomly selected data were predicted using the different priors
21      favored Bayesian model averaging with BIC prior over model selection (picking the best model)
22      with BIC or any approach with AIC. This method could be useful in assessing multi-pollutant
23      models.
24           The possibility that an observed effect is "real" (i.e., likely to be found in an independent
25      replication of the study) or merely  a statistical artifact is usually characterized by its confidence
26      interval or by its estimated significance level.  In most of this document, confidence intervals, or
27      credible intervals for Bayesian analyses, are reported in order to emphasize that the effect size is
28      not known with certainty, but some values are more nearly consistent with the data than effect
29      size values outside the interval. P-values or t-values are implicitly associated with a null
30      hypothesis of no effect.  A nominal significance level of  5% (i.e., a 95% confidence interval) is
31      usually used as a guide for the reader, but P-values should not be used as a rigid decision-making

        April 2002                                8-176       DRAFT-DO NOT QUOTE OR CITE

-------
 1      tool. If the observed confidence intervals were arrived at by a number of prior model
 2      specification searches, eliminating some worse fitting models, the true interval may well be
 3      wider.
 4           Given the now extremely large number of published epidemiologic studies of ambient PM
 5      associations with health effects in human populations and the considerably wide diversity in
 6      applications of even similar statistical approaches (e.g., "time-series analyses" for short-term PM
 7      exposure effects), it is neither feasible nor useful here to try to evaluate the methodological
 8      soundness of every individual study. Rather, two feasible approaches are likely to yield useful
 9      evaluative information: (1) an overall characterization of evident general commonalities (and/or
10      notable marked differences) among findings from across the body of studies dealing with
11      particular PM exposure indices and types of health outcomes; and (2) more thorough, critical
12      assessment of key newly published multi-city analyses of PM effects, given that greater scientific
13      weight is likely ascribable to their results than those of smaller sized studies  (often of individual
14      cities) yielding presumably less precise effects estimates. However, while pooling estimates
15      across cities may give more precise estimates of mean effect size, the uncertainty in the estimated
16      mean effect may also be inflated by differences in effect size among cities.
17           In the sections that follow, each of the five issues listed above (e.g., potential confounding
18      of PM effects by co-pollutants and so on) are critically discussed. In addition, given that the
19      newer multi-city study results, e.g, the NMMAPS analysis of the 90 largest U.S.  cities (Samet
20      et al., 2000a,b) show evidence of more geographical heterogeneity in the estimated PM risks
21      across cities and regions than had been seen in the studies assessed in the 1996 PM AQCD, the
22      issue of geographical heterogeneity in PM effects estimates  also warrants further evaluation here
23      (as is done in Section 8.4.9).
24
25      8.4.2 Assessment of Confounding by Co-Pollutants
26      8.4.2.1 Introduction
27           Airborne particles are found among a complex mixture of atmospheric pollutants, some of
28      which are well measured (such as gaseous criteria co-pollutants O3, CO, NO2, SO2) and others
29      which are not routinely measured.  The basic question here is one of determining the extent to
30      which observed health effects can be attributed to airborne particles acting alone or in
31      combination with other air pollutants. Many of the pollutants are closely correlated due to

        April 2002                                 8-177        DRAFT-DO NOT QUOTE OR CITE

-------
 1      emissions by common sources and dispersion by common meteorological factors (so that it may
 2      be difficult to disentangle their effects (as noted in Section 8.1.1), because some are in the
 3      pathway of formation of other pollutants, e.g.:  NO —* NO2  —¥ NO34  —> Particle Mass.
 4           It is widely accepted that some PM metrics are associated with health effects, and that PM
 5      has effects independent of the gaseous co-pollutants. The extent to which ambient gaseous
 6      co-pollutants may have health effects independent of PM is less certain, but this is important in
 7      considering the extent to which health effects attributed to PM may actually be due in part to
 8      co-pollutants or to some other environmental factors, and conversely.  EPA produces Air Quality
 9      Criteria Documents for four gaseous pollutants: CO, NO2, SO2, and O3. The possible health
10      effects of the gaseous pollutants exerted independently from PM, and in some cases jointly with
11      PM, are discussed in those documents.  They are also considered to some extent in this section
12      and elsewhere in this document because they affect quantitative assessments of the effects of
13      various PM metrics when these other pollutants are also present in the atmosphere. The gaseous
14      pollutants may also be of interest as PM effect modifiers, or through interactions with PM.
15           Co-pollutant models have received a great deal of attention in the last few years because
16      there now exist improved statistical methods for estimating PM effects by analyses of daily time
17      series of mortality (Schwartz and Marcus,  1990; Schwartz, 1991) or hospital admissions
18      (Schwartz, 1994) and/or in prospective cohort studies (Dockery et al.,  1993). For example, in the
19      most recent AQCD for NO2 (U.S. EPA, 1993), there are only three epidemiology studies on
20      mortality, a daily time series study (Lebowitz, 1971) and two ecological analyses (Hickey et al.,
21      1970; Mendelsohn and Orcutt, 1979).  The results of these earlier studies  are described by U.S.
22      EPA (1993) as non-significant or inconclusive.  By comparison, many of the studies using the
23      new methods have found significant positive relationships between mortality and one or more of
24      the four gaseous criteria pollutants in daily time series studies, and between SO2 and mortality in
25      the reanalyses of two large prospective cohort studies (Krewski et al., 2000). In the daily time
26      series studies, the estimated PM effect is relatively stable when the co-pollutant is included in the
27      model in some cities,  whereas the estimated PM effect in other cities changes substantially when
28      certain co-pollutants are included.  In the Krewski et al.  (2000) analyses, the estimated effect of
29      SO4= is greatly decreased when SO2 is also included as a predictor in a proportional hazards
30      model. How should these findings be interpreted?
        April 2002                                8-178        DRAFT-DO NOT QUOTE OR CITE

-------
 1           A number of the analyses presented below also discuss models in which multiple particle
 2      metrics are present, either with or without the gaseous criteria pollutants.  These mixtures are
 3      encountered in urban air.  Included among the studies evaluating both fine and coarse particles
 4      simultaneously are:  Burnett et al. (2000),  Chock et al. (2000), Clyde et al. (2000), Fairley et al.
 5      (1999), Lippmann et al. (2000), Mar et al.(2000); Cifuentes et al. (2001), and Castillejos et al.
 6      (2000).
 7           Gaseous co-pollutant levels may be correlated with total PM mass, but may be even more
 8      strongly correlated with specific PM constituents due to their emission from a common source
 9      (e.g., CO and NO2 from motor vehicle exhaust). The levels of a specific gaseous co-pollutant
10      may serve as an indicator of the day-to-day variation in the contribution of a distinct emission
11      source and to the varying composition of airborne PM.  In a model with total PM mass, a gaseous
12      co-pollutant may serve as a surrogate for the source-apportioned contribution to ambient air PM.
13      It would be interesting to evaluate models with  both source-relevant particle components (e.g.,
14      attributable to motor vehicles, coal combustion, oil combustion) and gaseous pollutants. The
15      closest approach is Model n in Burnett et al. (2000).
16           Carbon monoxide and NO2 may be acting as indicators of distinct emission sources
17      (primarily motor vehicles) and as indicators of PM from these sources (primary particles and
18      secondary nitrate particles). However, there are other sources of NO2, such as emissions from
19      coal- or oil-burning electric power plants.
20           The role of gaseous pollutants as surrogates for source-apportioned PM may be distinct
21      from confounding.  The true health effect may be independently associated with a particular
22      ambient PM constituent that may be more or less toxic than the particle mix as a whole. Thus,
23      a gaseous co-pollutant may give rise to the appearance of confounding in a regression model.
24      By serving as an indicator of the more toxic particles, the gaseous co-pollutant could greatly
25      diminish the coefficient for total particle mass.  In such a model, the coefficient for total particle
26      mass would most properly be interpreted an indicator of the other, less-toxic particles.  The
27      conceptual issues in evaluating potential confounding are at least as complex as the  technical
28      aspects discussed below. We restrict our discussion to daily time series studies.
29
30
31

        April 2002                                8-179       DRAFT-DO NOT QUOTE OR CITE

-------
 1           The conceptual problems in answering the question about confounding are:
 2
 3      (a) Biological plausibility:  Can some of the gaseous criteria pollutants cause increases in
 4      mortality or hospital admissions rates in the (presumably most susceptible) sub-populations at
 5      current levels of exposure to ambient concentrations?  If so, are these increases in mortality or
 6      hospital admissions likely to be associated with cardiovascular or respiratory causes?
 7
 8      (b) Exposure plausibility: Do some members of the population have personal exposure to both
 9      the particle metrics of ambient origin and the gaseous pollutants of ambient origin?  Also, do
10      susceptible subpopulations have greater or smaller personal exposure to ambient particles or
11      gases than the population as a whole?
12
13           The technical problems in answering these question(s) are:
14
15      (c) Is the model mis-specified (omission of predictive regressors, inclusion of correlated but non-
16      predictive  regressors, non-linearity, lags, measurement error from use of proxy variables)?
17
18      (d) Is there a bias in effect size estimates as a result of model mis-specification?
19
20      (e) Are the estimates of effect size standard errors sensitive to model mis-specification?
21
22      (f) Do some  of the mis-specification errors compound each other, e.g., non-linearity combined
23      with measurement error?
24
25      (g) Are effect size estimates and their standard errors really significantly different among
26      models?
27
28
29
        April 2002                                 8-180       DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.4.2.2  Issues
 2      8.4.2.2.1 Conceptual Issues in Assessing Confounding
 3           These concerns overlap two of Hill's (1965) suggested criteria for causal inference.
 4           (a) Biological plausibility: It is generally accepted that O3, NO2, and SO2 are associated
 5      with diminished pulmonary function and increased respiratory symptoms as well as more serious
 6      consequences, and CO exposure has been associated with cardiovascular effects.  While one may
 7      question whether adverse health effects occur in most healthy people at current exposure to
 8      ambient concentrations, there may be a susceptible sub-population for whom ambient gaseous
 9      pollutants cause health effects.  One should remember that less than 20 years ago, current levels
10      of exposure to ambient concentrations  of PM were thought to be safe. It would be premature to
11      conclude that the gaseous co-pollutants at current ambient levels are not associated with
12      respiratory and cardiovascular health effects in susceptible subpopulations.
13           Ambient gaseous co-pollutants can be potential confounders of ambient PM if:  (a) both the
14      gas and PM are able to cause the same  health effects; (b) if personal exposure is correlated with
15      ambient concentrations for both particles and gases respectively; (c) if the personal exposure to
16      gases and to particles are correlated, and; (d) if the ambient concentrations of particles and gases
17      are correlated. If any of these conditions fail, then we may have any of the conditions called
18      "under-fitting", "over-fitting", or "mis-fitting" described in Section 8.4.2.2.2.
19           (b) Exposure plausibility: While most Americans  spend most of their time in indoor
20      microenvironments, there is still sufficient personal exposure to O3 to cause frank respiratory
21      symptoms among sensitive children or adults exercising outdoors when ambient O3
22      concentrations are high (hence the declaration of "ozone alert"  days).  It is also likely that some
23      fraction  of ambient CO also contributes to indoor air pollution and total personal CO exposure.
24      Nitrogen dioxide,  while reactive, also penetrates indoors; and an ambient pollution component of
25      total personal exposure to NO2 can be identified among individuals without indoor NO2 sources
26      and close to strong outdoor sources such as highways. While there may be some, perhaps many,
27      individuals exposed to elevated concentrations of the gaseous criteria pollutants, in order to
28      contribute to the health effects associated with ambient concentrations of a co-pollutant (e.g,.
29      PM), the ambient  gaseous pollutants must be significantly and positively correlated with the
30      exposure to the co-pollutant.
31

        April 2002                                8-181       DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.4.2.2.2 Statistical Issues in the Use of Multi-Pollutant Models
 2           Confounding describes a condition in which one observable potentially explanatory
 3      variable in an epidemiological study can stand in for another one, leading to a confusion as to
 4      which variable may be causing the outcome. In most PM epidemiology studies, the gaseous
 5      pollutants can often stand in for the PM metric because there is frequently a high degree of
 6      positive linear correlation among PM metrics and all criteria gaseous pollutants but ozone.  This
 7      condition, known as multi-collinearity, is necessary to establish confounding, but not sufficient.
 8           We will demonstrate these important concepts graphically using causal pathway models.
 9      Figure 8-16 shows  a model with two pollutants (A and B) whose ambient concentrations and
10      personal exposure are correlated, and both are capable of producing the health outcome. If both
11      A and B exposure concentrations are used in a regression model for the health outcome, and both
12      are in fact causal, then the model is correctly specified. If the personal exposures are available,
13      then estimates of the health effects of both A and B as covariates or regressors will be unbiased,
14      but are likely to have large variances because exposures to A and B are correlated.  If only
15      ambient concentrations of A and B are available, then the effect estimates will be biased as well
16      as having large variances, but may still be predictive of health effects for personal exposures to
17      A and B of ambient origin. Disentangling the effects of A and B may be difficult.  This
18      corresponds to common notions  of confounding.
19
20
                     Amhipnt A 	fr-  Ambient A is used as a regressor
                     AmDemA        +      Personal exposure to A
                                            Personal exposure to B
                                        Ambient B is used as a regressor
                                                                      Health Outcomes
        Figure 8-16.  Graphical depiction of actual confounding of the effects of ambient A and
                     ambient B.
 1           In the figures below, a solid line with an arrow suggests a causal relationship, dot-dash
 2      lines suggest a non-direct association, and dotted lines suggest the absence of a pathway, either
 3      for exposure or outcome.  In principle, it is possible to carry out additional studies to determine
        April 2002                                8-182       DRAFT-DO NOT QUOTE OR CITE

-------
 1      whether A and B are both capable of causing independent health effects, although clinical trials
 2      with a small number of participants may not have sufficient power to detect and cannot be used
 3      to study highly adverse effects (death, hospital admissions) associated with particles or gases at
 4      current ambient concentrations and personal exposure levels. Case-crossover study designs may
 5      allow larger populations with adverse events to be studied, but are limited by the amount of
 6      personal exposure data that can be attributed to the case prior to the occurrence of the adverse
 7      outcome. There are a growing number of studies relating ambient concentrations and personal
 8      exposures for particles and gases.
 9           In Figure 8-16, we assume  that both pathways occur in nature, hence a large increase in the
10      standard errors or variances of the effect size estimates in a multi-pollutant study a natural
11      description of confounding.  However, variance inflation and effect size instability may also be
12      found in the absence of confounding, as shown in Figures 8-17 through 8-20. In summary,
13      multi-pollutant models may be useful tools for assessing whether the gaseous co-pollutants may
14      be potential confounders of PM effects, but cannot determine if in fact they are. Variance
15      inflation and effect size instability can occur in non-confounded models as well as in confounded
16      models. Our usual regression diagnostic tools can only determine whether there is a potential for
17      confounding.  Therefore, although multi-colinearity leading to effect size estimate instability and
18      variance inflation are necessary conditions for confounding, they are not sufficient by themselves
19      to determine whether confounding exists.
20
21

                     Amhipnt A 	fc.  Ambient A is used as a reqressor 	„
                     AmoentA       +•      Personal exposure to A              1
                         '                           !
                         I                           I                  Health Outcomes
                                                                          J
                       	fc.       Personal exposure to B      __
                            r  Ambient B is not used as a regressor

Figure 8-17.  Graphical depiction of under-fitting of A and B.  Only ambient A is used as
              a regressor (covariate) and B is omitted.  The estimate of A's effect is biased
              because it includes the causal effect of the omitted variable B which is
              correlated with A.
        April 2002                                 8-183        DRAFT-DO NOT QUOTE OR CITE

-------
             Amhipnt A        i»-  Ambient A is used as a reqressor
             AmmentA        »      Personal exposure to A
                                                             Health Outcomes

                                                                  A
                                   Personal exposure to B
                                Ambient B is used as a regressor
Figure 8-18.  Only A is causal, B is not related to the outcome, but both regressors are

             included in the model, a likely cause of variance inflation.
             Amhipnt A 	h-  Ambient A is used as a regressor
             AmmentA        >      Personal exposure to A
                                                             Health Outcomes
                                                                  t
                             **     Personal exposure to B     	
                             >-  Ambient B is useg as a regressor





Figure 8-19.  Graphical depiction of over-fitting of A and B. Both A and B are causal, and

             both ambient A and ambient B are used as regressors (covariates).  However,

             there is no relationship of ambient B to personal B. The estimate of A's

             effect is biased because it includes the hypothetical causal effect of ambient

             B, which is correlated with A, but for which there is no personal exposure.
             Amhipnt A  	<»> Ambient A is not used as a reqressor_
             Arm em A      »•       Personal exposure to A
                                   Personal exposure to B
                                Ambient B is useg as a regressor
                                                             Health Outcomes

                                                                  A
Figure 8-20.  Graphical depiction of mis-fitting of the effects of A and B.  Only A is causal,

             B is not related to the outcome, but B is used as a regressor in the model and

             the effects of A are transferred to B.
April 2002                               8-184        DRAFT-DO NOT QUOTE OR CITE

-------
 1           The most commonly used methods include multi-pollutant models in which both the
 2      putative causal agent (PM) and one or more putative co-pollutants are used to estimate the health
 3      effect of interest.  If the effect size estimate for PM is "stable", then it is often assumed that the
 4      effects of confounding are minimal.  "Stable" is usually interpreted as meaning that the
 5      magnitude of the estimated effect is similar in models with PM alone and in models with PM and
 6      one or more co-pollutants, and the statistical significance or width of the confidence interval for
 7      the PM effect is similar for all models, with or without co-pollutants. These (usually
 8      unquantified) criteria diagnose confounding in a narrow sense, interpreted as synonymous with
 9      multi-collinearity, not as a failure of the study design or other forms of model mis-specification.
10           (c) Model mis-specification assumes many forms. The omission of predictive regressors
11      ("underfitting", defined by Chen et al., 2000) may produce biased estimates of the effects of truly
12      predictive regressors that are included in the model.  Inclusion of unnecessary or non-predictive
13      regressors along with all truly predictive regressors ("over-fitting") will produce unbiased
14      estimates of effect, but may increase the estimated standard error of the estimated effect if it is
15      correlated with other predictors. Omitting a truly predictive regressor while including a
16      correlated but non-causal variable ("mis-fitting") will attribute the effect of the causal regressor
17      to the non-causal regressor. Interaction terms are candidates for omitted regressor variables. It is
18      important to avoid the "mis-fitting" scenario. Assuming there is a linear relationship when the
19      true concentration-response function is non-linear will produce a biased estimate of the effect
20      size, high or low at different concentrations. One of the most common forms of model mis-
21      specification is to use the wrong set of multi-day lags, which could produce any of the
22      consequences described as "under-fitting" (e.g., using single-day lags when a multi-day or
23      distributed lag model is needed), "over-fitting" (e.g., including a longer span of days than is
24      needed), or "mis-fitting" (e.g., using a limited set of lags while the effects are in fact associated
25      with different set of lags). Different PM metrics and gaseous pollutants may have different lag
26      structures, so that in a multi-pollutant model, forcing both PM and gases to have the same lag
27      structure is  likely to yield "mis-fitting". Finally, classical exposure measurement errors (from
28      use of proxy variables) attenuates (biases) effect size estimates under most assumptions about the
29      correlations among the regressors and among their measurement errors (Zeger et al., 2000).
30
31

        April 2002                                 8-185       DRAFT-DO NOT QUOTE OR CITE

-------
 1           (d) Bias: All of the mis-specifications listed in (c) can bias the effect size estimate except
 2      for "over-fitting" and measurement error of Berkson type.  The estimates of the standard error of
 3      the effect size estimate under "over-fitting" or Berkson error cases are inflated, however.
 4           (e) Estimates of effect size standard errors are usually sensitive to model mis-specification.
 5      When all truly predictive regressors are added to an "underfit" model, the uncertainty will almost
 6      always be reduced sufficiently that the standard errors of estimated effect size are reduced
 7      ("variance deflation"). Adding correlated non-causal variables to "over-fitted" or "mis-fitted"
 8      models will further increase the estimated standard errors ("variance inflation"). Variance
 9      inflation can occur whenever a covariate is highly correlated with the regressor variable that is
10      presumably the surrogate for the exposure of interest.  Confounding with the regressor variable
11      can occur only when the covariate is correlated (a) with the regressor variable proxy for the
12      exposure of interest and (b) with the outcome of interest in the absence of the exposure of
13      interest.
14           (f) Mis-specification errors may compound each other. If the concentration-response
15      function is nonlinear but there is measurement error in the exposures, then different sub-
16      populations will have greater or smaller risk than assigned by a linear model. Consider the
17      hypothetical case of a "hockey-stick" model with a threshold.  If there were no  exposure
18      measurement error, then the part of the population with measured concentrations above the
19      threshold would have excess risk, whereas those below would not. If exposures were measured
20      with error, even if the measured concentration were above the threshold, some people would
21      actually have exposures below the threshold and no excess risk. Conversely, if the measured
22      concentration was below the threshold, some people would actually have concentrations above
23      the threshold and would have excess risk.  The flattening of a non-linear concentration-response
24      curve by measurement error is a well known phenomenon that may be detected by standard
25      methods (Cakmak et al., 1999).
26           (g) The question of whether effect size estimates and their standard errors are really
27      significantly different among models is usually not addressed quantitatively.  Some authors
28      report various goodness-of-fit criteria such as AIC, BIC, deviance, or over-dispersion index, e.g.,
29      (Chock et al., 2000; Clyde et al., 2000), but the practice is not yet so wide-spread as to assist in
30      analyses of secondary data for use in this document.  These models are not strictly nested.
31

        April 2002                                8-186        DRAFT-DO NOT QUOTE OR  CITE

-------
 1           Variance inflation may also happen with a correctly specified model when both pollutants
 2      are causal and highly correlated, compared to a model in which only one pollutant is causal and
 3      the non-causal pollutant is omitted. The situation where variance or the standard error decreases
 4      when an additional variable is added (variance deflation) suggests that the model with the
 5      covariate is more nearly correct and that the standard errors of all covariates may decrease.
 6      Statistical significance is a concept of dubious usefulness in assessing or comparing results of
 7      many models from the same data set.  Still, it is a familiar criterion, and one we address by using
 8      a nominal two-sided 5% significance level for all tests and 95% confidence intervals for all
 9      estimates, acknowledging their limitations.  There is at present no consensus on what clearly
10      constitutes "stability" of a model estimate effect size, e.g., effect sizes that differ by no more than
11      20% (or some other arbitrary number) from the single-pollutant models.  Simple comparison of
12      the overlap of the confidence intervals of the models is not used because the model estimates use
13      the same data, and the confidence intervals for effect size in different models are more-or-less
14      correlated. In analyses with missing days  of data for different pollutants, comparisons must also
15      incorporate differences in sample size or degrees of freedom.  Some examples of (a) changes in
16      the statistical significance of PM effects in different models are evident from inspection of
17      Figures 8-21 to 8-25 and of (b) relative  stability in significance of PM effects in Figure 8-26.
18           In any case, statistical comparisons cannot answer questions about either conceptual or
19      statistical issues in  confounding with claims about statistical significance.  If the model is
20      mis-specified in any of the numerous ways described above, then effect size estimates or their
21      estimated standard  errors are biased. Statistical assessments alone can determine if the PM
22      metric is too closely correlated with the other pollutants to provide an accurate quantitative effect
23      size estimate, which is of course useful  information even if we conclude that it is not feasible to
24      estimate the separate effects of PM and  its gaseous co-pollutants. Confounding cannot occur if
25      the gaseous co-pollutants cannot produce the health outcome, or  if there is no personal exposure
26      to the gaseous co-pollutants, or the personal exposure to them is  not correlated with their ambient
27      concentrations.
28           We will start  by considering what can be learned from the most commonly used approach
29      to diagnose potential confounding, fitting  multi-pollutant models and evaluating the stability of
30      the estimated particle effect sizes against inclusion of co-pollutants.
31

        April 2002                                8-187        DRAFT-DO NOT QUOTE OR CITE

-------
          San Diego
a
S.
                                   Angeles
                                 Cleveland
                               Dallas - Ft. Worth
                                                         Pittsburgh
                                                       San Bernadino
                                                    ^^5^t°o'^0'1
                                                      V^V1 -g??
                                                        Philadeiphia
                                                         ^V*

                                                          Seattle
                                                                               \
                                                                                 San Antonio
                                                                              Santa Ana -Anaheim
                                                                                 Minneapolis
Figure 8-21.  Effects of PM10 on total mortality in 20 large U.S. cities, as a function of
              co-pollutant models. EPA presentation of results from Samet et al. (2000b).
April 2002
8-188        DRAFT-DO NOT QUOTE OR CITE

-------
                                                                                  >VV"
                                                                                 Model Type for Sulfur Dioxide
        Figure 8-22.  Effects of particles and gases on total mortality in eight Canadian cities.
                     EPA presentation of results in Burnett et al. (2000).
 1           The studies identified in Table 8-34 are too numerous to allow detailed narrative
 2      description here.  Rather, Table 8-34 provides a summary emphasizing the points of greatest
 3      relevance in evaluating multi-pollutant models.  The issue of the stability of the effect size
 4      estimate in multi-pollutant models may perhaps be assessed best by reporting the range of effect
 5      size estimates across different co-pollutant models, in the absence of quantitative goodness-of-fit
 6      comparison criteria in almost all of the papers cited.  Thus, in addition to identifying the study,
 7      the endpoint (usually total mortality), the PM metric, and the lags used in the analyses, the
 8      minimum and maximum effect size estimates and the co-pollutants (if any) for which the
 9      estimates were calculated are included. It is not uncommon for the single-pollutant PM model to
10      have the maximum PM effect size, in which case the co-pollutant is listed as "nothing."
11           There is some agreement on what constitutes "variance deflation".  If an additional
12      covariate is added to a baseline model (e.g., with PM alone) and the model predicts the outcome
        April 2002
8-189
DRAFT-DO NOT QUOTE OR CITE

-------
                            Maricopa County, Az.
                              Cook County, II.
                       S  2
                              i   i   i
                           012345
                                  PM,0Lag
                 Los Angeles
                                                            10
                                                          E
                                                          S
                                                          ir -10
                                                          5  -5
            0123456
                   PM,0Lag
                 Los Angeles
            0123456
                   PM2 5 Lag
        Figure 8-23.  Effects of PM10 or PM25 on circulatory mortality in three U.S. cities as a
                     function of lag days.  Dark shading is a co-pollutant model, light shading is
                     a single-pollutant PM model, medium shading shows overlap between single-
                     pollutant and co-pollutant models. EPA presentation of results in
                     Moolgavkar (2000b).
 1     better with the covariate, then the reduction in variance (or deviance for generalized linear or
 2     additive models [GLM or GAM]) outweighs the loss of degrees of freedom for variability.
 3     Although not always true, it is reasonable to expect a decrease in the estimated asymptotic
 4     standard error of the effect size estimate, but improved goodness-of-fit may not reduce the
 5     standard errors  of all parameters in equal proportion because introducing the new covariate
 6     modifies the covariate variance-covariance matrix. The weighted inverse covariance matrix
 7     provides an exact estimate for standard errors in ordinary linear regression models, and
 8     approximately so in GLM or GAM.  The effects on other parameter estimates are rarely reported.
 9           "Variance inflation" may occur under several circumstances, including "under-fitting" and
10     "mis-fitting" in which a truly predictive covariate is omitted or replaced by a correlated proxy,
       April 2002
8-190
DRAFT-DO NOT QUOTE OR CITE

-------
                        Model Type for Fine Particles      Model Type for Coarse Particles     Model Type for Nitrate in PMie
                          Model Type for Ozone
                                          Model Type for Carbon Monoxide    Model Type for Nitrogen Dioxide
Figure 8-24.  Total mortality from particles and gases in Santa Clara County, CA.
               EPA presentation of results in Fairley (1999).
                   Total Mortality
Circulatory Mortality
Respiratory Mortality
                                             PM:, Model Type
                E
                S,  5
                8
                                                              B  K
                                                                          \ /
                                                                   ,   ,   ,  Y  ,
                                                                    PMlc.-j Model Type
Figure 8-25.  Cause-specific fine or coarse particle mortality in Detroit, MI.
               EPA presentation of results in Lippmann et al. (2000).
April 2002
       8-191         DRAFT-DO NOT QUOTE OR CITE

-------
        i
        I ,
             Total Mortality
Elderly Mortality
                                PMF Model Type
Cardiovascular       Respiratory Mortality      Other Mortality
  Mortality
                                                   PMF Model Type
                                                                                     ^f ^°- ^-°- o.*P-
                                                                                       PMF Mode] Type
        Figure 8-26.   Effects of fine particles on total mortality in Mexico City. EPA presentation
                      of results in Borja-Aburo et al. (1998).
 1      and "over-fitting" in which a non-predictive covariate correlated with the PM metric is also
 2      included in the model.  The potential for over-fitting can be diagnosed by evaluating the
 3      eigenvalues of the correlation matrix of the predictors, with very small values identifying near-
 4      collinearity. However, the complete covariate correlation matrix is almost never reported,
 5      including all weather variables and nonlinear functions entered separately as covariates.
 6      Nonetheless, even a correlation matrix among all pollutants would be informative. Furthermore,
 7      composite correlation matrices in multi-city studies may conceal important differences among
 8      the correlation matrices.
 9           In the absence of any better criterion, we arbitrarily define "variance deflation" and
10      "variance inflation" as occurring when the estimated standard error of the effect size estimate
11      differs by more than 25% from the single-pollutant models. These are included in Table 8-34 for
12      models for total mortality. Figures 8-21 through 8-26 show results for two multi-city studies, one
13      in the U.S, (Samet et al., 2000b) and one in Canada (Burnett et al., 2000), as well as for some
14      single-city studies in the U.S.  and Mexico (Lippmann et al., 2000; Fairley et al., 1999; Borja-
15      Abuto et al., 1999).  We do not show other studies because of limitations in space. Readers may
16      form their own judgements about "stability" and "variance inflation/deflation". There are
17      examples of both parameter estimate stability and instability, and of variance inflation and
18      deflation, as noted in Table 8-34.
19
        April 2002
                    8-192
              DRAFT-DO NOT QUOTE OR CITE

-------
3.
TABLE 8-34. CHARACTERIZATION
      INFLATION OR DEFLATION
OF CO-POLLUTANT EFFECTS ON THE STABILITY AND VARIANCE
OF PM EFFECT SIZE ESTIMATE (IN TERMS OF EXCESS RR)
l^
o
o
to










oo
VO
oo



o
^
H
6
0
o
H
0
0
H
W
O
O
H
w
Authors
Samet et al
Samet et al.
Samet et al.
Samet et al.
Samet et al.
Samet et al
Samet et al.
Samet et al.
Samet et al.
Samet et al.
Samet et al
Samet et al.
Samet et al.
Samet et al.
Samet et al.
Samet et al
Samet et al.
Samet et al.
Samet et al.
Samet et al.
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar

Year
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b

Oty
Los Angeles
New York
Chicago
Dallas-FtW
Houston
San Diego
SantaAna-
Phoenix
Detroit
Miami
Philadelphia
Minneapolis
Seattle
San Jose
San Bernard
Cleveland
Pittsburgh
Oakland
Atlanta
San Antonio
Chicago
Chicago
Chicago
Chicago
Chicago
Chicago
Los Angeles
Los Angeles

Endpoint
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.

PM
Metric
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10

Lag
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
0
1
2
3
4
5
0
1

PM
Alone Age or
Exc. RR Season
1.9
5.7
1.6
-2
1
5.6
3.5
3.3
2.4
3.5
3.9
2.4
1.4
1.6
1.3
-0.2
2
10.8
0.2
3.5
2.4
2.1
1.6
1.1
0.8
-0.5
0.4
0.9

Min.
ExcRR
-0.3
1.1
-0.2
-2
0.1
2.1
1.9
-0.4
2.1
3
3.5
NA
1.4
-1.4
1
-0.2
1.3
5.7
-5.4
3.5
1.9
1.9
1.6
1
0.8
-0.5
-2
-4

PM with
O3, NO2,
CO
O3, SO2
03
nothing
O3, NO2
O3, SO2
O3! NO2
O3, CO
O3, CO
03
03
NA
nothing
O3! NO2
O3, NO2
nothing
O3, NO2
O3, CO
O3, NO2
nothing
CO
CO
CO
CO
CO
nothing
CO
CO

Max.
ExcRR
2.1
6.1
1.6
3.4
1.1
5.8
5.3
10.2
3.3
3.5
6.5
NA
7.4
1.9
1.6
0.4
2.2
10.8
0.25
4
2.4
2.1
1.6
1.1
0.8
-0.4
0.4
0.9

PM
with
03
03
nothing
03
03
03
03
O3, NO2
O3, SO2
nothing
O3, NO2
NA
03
03
O3, SO2
O3, NO2,
CO
O3, SO2
nothing
nothing
O3, CO
nothing
nothing
either
nothing
either
CO
nothing
nothing

Variance
Inflation Inflation Inflation Inflation Deflation
O3, CO O3, NO2
O3! CO O3, NO2 O3, SO2
O3, SO2


O3, CO O3! NO2
O3! CO O3, NO2
O3, CO O3, NO2 O3, SO2
O3, CO O3, NO2 O3, SO2

O3! NO2 O3, SO2

O3 O3, CO
O3, CO O3! NO2

O3 O3, CO O3, NO2 O3, SO2
O3! CO O3, NO2 O3, SO2
O3, CO O3! NO2 O3, SO2
O3, CO O3, NO2
CO
CO
CO
CO
CO
CO
CO


-------
3.
TABLE 8-34 (cont'd). CHARACTERIZATION OF CO-POLLUTANT EFFECTS ON THE STABILITY AND VARIANCE
         INFLATION OR DEFLATION OF PM EFFECT SIZE ESTIMATE (IN TERMS OF EXCESS RR)
l^
o
o
to













oo
^
VO
•^




O
^
H
1
O
o
2!
-*— I
O
H
O

0
H
W
O


H
W


Authors
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar
Moolgavkar

Moolgavkar

Moolgavkar
Moolgavkar
Moolgavkar
Fairley

Fairley
Fairley


Burnett et al.

Burnett et al.

Burnett et al.

Burnett et al.

Lippmann
etal.






Year
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b
2000b

2000b

2000b
2000b
2000b
1999

1999
1999


2000

2000

2000

2000

2000







Oty
Los Angeles
Los Angeles
Los Angeles
Los Angeles
Phoenix
Phoenix
Phoenix
Phoenix
Phoenix
Phoenix
Los Angeles
Los Angeles

Los Angeles

Los Angeles
Los Angeles
Los Angeles
San Jose

San Jose
San Jose


Canada 8

Canada 8

Canada 8

Canada 8

Detroit







Endpoint
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.
Total Mort.

Total Mort.

Total Mort.
Total Mort.
Total Mort.
Total Mort.

Total Mort.
Total Mort.


Total Mort.

Total Mort.

Total Mort.

Total Mort.

Total Mort.






PM
Metric
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM10
PM2.5
PM2.5

PM2.5

PM2.5
PM2.5
PM2.5
PM25

PM10.2.5
PM10,
nitrate

PM2.5

PM25

PM10.2.5

PM10.2.5

PM25







Lag
2
3
4
5
0
1
9
3
4
5
0
1

2

3
4
5
0

0
0


0

1

0

1

3





PM
Alone Age or
Exc. RR Season
2.4
0.8
0.6
-2
2
3
4.5
-0.8
0.5
6.1
1.5
1.4

0.6

-0.6
0.2
-1.2
8.5

4.5
4.6


2.3

3

1.2

1.8

3.1






Min.
ExcRR
-1.9
-3.9
-1.8
-5.5
0.2
0.4
0.8
-0.8
-0.8
3
0.4
0.5

-1.2

-1.2
-1.1
-1.5
-0.1

-5.6
4.4


1.1

1.9

0.2

1

2.8

1.6





PM with
CO
CO
CO
CO
CO
CO
CO
nothing
CO
CO
CO
CO

CO

CO
CO
CO
PM10,
nitrate
PMF
NO2


03

four
gases
03

CO,NO2,
SO2
03

PM10.2.5




Max.
ExcRR
2.4
0.8
0.6
-2
2
3
4.5
-0.8
0.5
6.1
1.5
1.4

0.6

-0.6
0.2
-1.2
10.8

4.5
5.6


2.6

3

2.6

1.8

3.9






PM Variance
with Inflation Inflation Inflation Inflation Deflation
nothing
nothing
nothing
nothing CO
nothing CO
nothing CO
nothing CO
CO CO
nothing CO
nothing CO
nothing CO
nothing CO

nothing CO

nothing CO
nothing CO
nothing CO
NO2 O3, CO, nitrate
NO2
nothing
PM25 NO2 PM25


CO NO2

nothing CO NO2 O3, CO, NO2 Model I

SO2

nothing

SO2

PM10.2.5




-------
3.
TABLE 8-34 (cont'd). CHARACTERIZATION OF CO-POLLUTANT EFFECTS ON THE STABILITY AND VARIANCE
         INFLATION OR DEFLATION OF PM EFFECT SIZE ESTIMATE (IN TERMS OF EXCESS RR)
l^
o
o
to















oo
VO
01




o

£j
H
6
0
o
H
0
0
w
o
o
H
w

Authors
Lippmann
etal.

Chock et al.

Chock et al.

Chock et al.

Chock et al.

Chock et al.

Chock et al.

Chock et al.

Chock et al.

Goldberg
etal.
Goldberg
etal.










Borja-Abuto



Year City
2000 Detroit


2000 Pittsburgh

2000 Pittsburgh

2000 Pittsburgh

2000 Pittsburgh

2000 Pittsburgh

2000 Pittsburgh

2000 Pittsburgh

2000 Pittsburgh

2000a Montreal

2000a Montreal

previous
previous

previous
previous
any previous
any previous
age previous
65+

1999 Mexico City



Endpoint
Total Mort.


Total Mort.

Total Mort.

Total Mort.

Total Mort.

Total Mort.

Total Mort.

Total Mort.

Total Mort.

Total Mort.

Total Mort.

cancer
acute LRI

airways
congestive
coronary
cardio-
airways


Total Mort.


PM
Metric
PM10.2.5


PM10

PM10

PM10

PM10

PM2.5

PM2.5

PM10.2.5

PM10.2.5

PM2.5 (est)

PM2.5 (est)




disease
heart
artery
vascular
disease


PM25


PM
Alone
Lag Exc. RR
3


Tbl.30

Tbl. 90

Tbl.60

Tbl. 90

Tbl. 90

Tbl. 90

Tbl. 90

Tbl. 90

MeanO,
1,2
Mean 0,





failure
disease
disease
no
meds.

4


4


3.1

1.3

2

2.2

2.6

1.5

0.7

1.3

5.8

1,2

4.9
12.9

3.5
10.9
4.9
7.4
11.2


3.4


Age or Min.
Season Exc RR
3.1

2.8
age < 75 2.6

age < 75 1.3

age 75+ 2

age 75+ 2.2

age < 75 2.6

age 75+ 1

age < 75 0.7

age 75+ 1.3

5.1

with previous

4.3
12.8

1.6
9.4
4.4
7.3
10.5


3.4



PM with
03

PM2,
CO

nothing

nothing

nothing

nothing

all 4
gases
nothing

nothing

CO

illness

03
03

03
03
03
03
03


NO2


Max.
ExcRR
4.1


4.3

2

3.7

2.7

3.3

1.5

0.8

1.4

5.9



4.9
13

3.6
10.9
4.9
7.4
11.7


4.2


PM Variance
with Inflation Inflation Inflation Inflation Deflation
SO2

PM25
NO2 NO2 CO, SO2 all 4
gases
all 4
gases
CO NO2 CO, NO2 all 4
gases
all 4 all 4
gases gases
all 4
gases
nothing

all
4 gases
all 4
gases
SO2 CO
(slight)


nothing
SO2

SO2
nothing
nothing
nothing O3 SO2
SO2


O3, NO2 NO2 O3, NO2



-------
3.
TABLE 8-34 (cont'd). CHARACTERIZATION OF CO-POLLUTANT EFFECTS ON THE STABILITY AND VARIANCE
         INFLATION OR DEFLATION OF PM EFFECT SIZE ESTIMATE (IN TERMS OF EXCESS RR)
l^J
o
o
to















oo
1 .
VO
Oi






H
1
O
0
o
H
0
0
H
W
O
O
H
w


Authors
Castillej os
etal.
Castillej os
etal.
Castillej os
etal.
Ci&entes
etal.
Ci&entes
etal.
Ci&entes
etal.
Ci&entes
etal.
Ci&entes
etal.

Ci&entes
etal.
Atkinson
etal.
Katsouyanni
etal.


Year City
2000 Mexico City

2000 Mexico City

2000 Mexico City

2000 Santiago

2000 Santiago

2000 Santiago

2000 Santiago

2000 Santiago


2000 Santiago

2000 8 European

2001 29
European

PM
Endpoint Metric
Total Mort. PM10

Total Mort. PM2 5

Total Mort. PM10.2 5

Total Mort. PM2 5

Total Mort. PM2 5

Total Mort. PM2 5

Total Mort. PM10.2 5

Total Mort. PM10.2 5


Total Mort. PM10.2 5

Resp. Mort. PM10

Total Mort. PM10

PM
Alone Age or Min. Max. PM Variance
Lag Exc. RR Season Exc RR PM with Exc RR with Inflation Inflation Inflation Inflation Deflation
Avg. 9.5 9.5 nothing 13 NO2 O3, NO2
1,2
Avg. 3.7 0.4 PM10.25 3.7 nothing O3, NO2 PM10.25
1,2
Avg. 10.5 10.3 PM25 11 O3 O3, NO2 PM25
1,2
Avg. 4.5 summer 2.8 O3, 4.5 nothing
1, 2 PM10.2.5
Avg. 1.6 winter 0.8 NO2 1.8 PM10.25 PM10.25 CO SO2
1,2
Avg. 1.8 all year 0.8 NO2 1.8 nothing PM10.25 CO
1,2
Avg. 5.5 summer 3.7 PM10_25 5.5 nothing
1,2
Avg. 1.8 winter -0.5 PM10.25 1.8 nothing PM10.25 CO NO2
1,2

Avg. 2.3 all year 0.3 PM10.25 2.3 nothing PM10.25 CO NO2
1,2
4.6 3.6 SO2 5.6 NO2 O3 SO2

Avg. 3.4 1.8 NO2 3 SO2 O3 NO2
0, 1
Note: Stadies with ozone as the only co-pollutant are omitted.

























































-------
 1      8.4.2.2 Assessments of Confounding Using Multi-Pollutant Models with Observed Gases
 2           The most common approach to evaluation of confounding of PM effects by gaseous
 3      co-pollutants is to compare the estimates of the PM effect size in models with and without the
 4      gases. The single pollutant model is, in general, of the form
 5
 6                                RR = exp(pPM x PM + other covariates)                      (8-1)
 7
 8      and the corresponding multi-pollutant model is of the form
 9
10                  RR = exp(pPM x PM + pgasl x [gasl] + Pgas2 x [gas2] + other covariates)        (8-2)
11
12      If the estimates of PPM in model specifications in Equations 8-1 and 8-2 are very different, or if its
13      estimated standard error is much larger in Equation 8-2 than in Equation 8-1, then one may
14      conclude that PM is confounded with the gaseous co-pollutants, particularly if PM and the co-
15      pollutants are highly inter-correlated, as they often are. Variance inflation (large standard error)
16      of the estimated PPM in the multi-pollutant model suggests that the pollutants are collinear.
17      A large change in estimated PPM without much variance inflation suggests that some of the total
18      effect might be shared among PM and the gaseous pollutants, whereas a large change in the PM
19      coefficient along with a large increase in the estimated variance of the PM regression coefficient
20      suggests only that the PM coefficient is unstable.
21           The results in Table 8-34 show numerous cases of variance inflation, an expected
22      consequence of the multi-colinearity of PM and gaseous pollutants in most cities where
23      combustion products from motor vehicles, power plants, home heating, and industrial processes
24      dominate the urban air mix.  We have not tabulated findings for which the only co-pollutant is
25      ozone (Dominici et al., 2000a; Lipfert et al., 2000; Samet et al., 2000a,c), as these results appear
26      to add little to the findings in (Samet et al., 2000b).  Samet et al. (2000b) found that including
27      ozone along with PM10 tends to slightly increase the PM10 effect, but increasing the variance
28      substantially only in Cleveland and Seattle. Moolgavkar (2000b) finds that adding CO to a PM10
29      model substantially increases its variability at most single-lags in Chicago and Phoenix, but less
30      so in Los Angeles. Adding CO to the PM2 5 model in Los Angeles results in a substantial
31      reduction in the uncertainty of the PM2 5 effect, one of the few cases of variance  deflation,

        April 2002                                8-197        DRAFT-DO NOT QUOTE OR CITE

-------
 1      implying a better-fitting model.  This suggests that it may be easier to separate the effects of CO
 2      from the effects of PM2 5 than from the effects of PM10 in Los Angeles.
 3           Some of the studies in which either PM10, PM25, or PM10_2 5 coefficients are evaluated in
 4      multi-pollutant models are discussed below. A partial list of recent studies for which such
 5      assessments can be done is given in Table 8-35.  Only a subset of these studies are discussed as
 6      examples of what can be learned from multi-pollutant analyses. Table 8-34 presents a summary
 7      of the results, ordered starting with studies having clearer indications of PM effect size instability
 8      against co-pollutants through to studies having clear indications of PM effect size stability.
 9
10      Samet et al. (2000b) mortality in 20 U.S. cities
11           Most cities in Samet et al. (2000b) show a considerable reduction in effect size with
12      inclusion of ozone and another pollutant in the model, compared  to the model with PM10 alone.
13      The maximum PM10 effect size across co-pollutant models is rarely much larger than the single-
14      pollutant PM10 effect, except in Dallas-Fort Worth, Phoenix, and  Seattle.  The overall impression
15      in Figure 8-21 is that the co-pollutant models are neither consistently stable nor unstable, so that
16      the sensitivity of the model to co-pollutants may vary substantially from city to city. The results
17      shown are all for a single lag day 1. Results in Moolgavkar (2000b) suggest that different single-
18      day lags in different cities may  affect the  apparent stability and variance inflation or deflation of
19      the estimated PM effects.
20
21      Moolgavkar (2000b) total and cardiovascular mortality in 3 U.S. cities.
22           Results for total mortality are shown in Figures 1, 2, and 3 in the Moolgavkar (2000b)
23      paper, using CO as the only co-pollutant.  The assessment of stability and variance
24      inflation/deflation against co-pollutants and lags is shown in Table 8-34 for total mortality where
25      CO is the only co-pollutant, for consistency with Samet et al. (2000b).  The results for
26      cardiovascular mortality are shown in Figure 8-23. It is clear that the results depend on both city
27      and lag. The  PM10 models for total mortality in Los Angeles and Phoenix show systematic
28      attenuation  of effect when CO is added, whatever the lag. The PM10 effect size in Chicago  is
29      only  somewhat attenuated by CO. Both Chicago and Phoenix show variance inflation by CO at
30      all lags. In Los Angeles, however, the PM10 effect shows little variance inflation by CO  except at
31      lags 0 ands  5, even though the PM10 effect is strongly attenuated.  The PM2 5 effect size on total

        April 2002                                 8-198        DRAFT-DO NOT QUOTE OR CITE

-------
         TABLE 8-35. SOME NEW DAILY TIME SERIES STUDIES FOR
      MORTALITY OR MORBIDITY WITH CO-POLLUTANT MODELS AND
                      GRAVIMETRIC PM INDICES
Study
City or Cities
PM Indices
Co-Pollutants
Studies in the U.S. and Canada
Burnett et al. (2000)
Chock et al. (2000)
Dominici et al. (2000a)
Fairleyetal. (1999)
Goldberg et al. (2000,
2001abcd)
Gwynn et al. (2000);
Gwynn and Thurston (2001)
Lipfertetal. (2001)
Lippmann et al. (2000)
Moolgavkar (2000a)


Moolgavkar (2000b)


Samet et al. (2000abc)
8 Canadian Cities
Pittsburgh, PA
19 U.S. cities with
ozone data
Santa Clara County,
CA
Montreal, PQ, Canada
Buffalo, NY
Philadelphia, PA -
Camden, NJ
Detroit, MI
Los Angeles, CA
Chicago, IL
Phoenix, AZ
Los Angeles, CA
Chicago, IL
Phoenix, AZ
19 U.S. cities with
co-pollutant models
PM2 5, PM10.2.5, PM10
PM2 5, PM10.2.5, PM10
PM10
PM25,PM10.25,PM10,
COH, NO3', SO4=
Estimated PM2 5,
Sutton sulfate, COH
PM25,PM10, COH,
SO;, H+
PM2.5, PM10.2.5, PM10
PM2 5, PM10.2.5, PM10
PM2 5, PM10.2.5, PM10
PM10
PM10
PM2 5, PM10.2 .5, PM10
PM10
PM10
PM10
O3, CO, NO2, SO2
O3, CO, NO2, SO2
03
O3, CO, NO2
O3, CO, NO2, SO2, NO
O3, CO, NO2, SO2
O3 (only particle
co-pollutant)
O3, CO, NO2, SO2
O3, CO, NO2, SO2
O3, CO, NO2, SO2
O3, CO, NO2, SO2
O3, CO, NO2, SO2
O3, CO, NO2, SO2
O3, CO, NO2, SO2
O3, CO, NO2, SO2
Studies in Latin America
Borja-Abuto et al. (1998)
Castillejos et al. (2000)
Cifuentes et al. (2000)
Loomisetal. (1999)
Mexico City, D.F.,
Mexico
Mexico City, D.F.,
Mexico
Santiago, Chile
Mexico City, D.F.,
Mexico
PM25
PM2 5, PM10.2 .5, PM10
PM2.5, PM10.2.5, PM10
PM25
O3, NO2
03, N02
O3, CO, NO2, SO2

Studies in Europe
Atkinson etal. (2001)
Katsouyanni et al. (2001)
Sunyer and Basagna (2001)
8 European cities in
APHEA2
29 European cities
Barcelona, Spain
Several, converted
to PM10
PM10, Black Smoke (BS)
PM10
O3, CO, NO2, SO2
O3, NO2, SO2
O3, CO, NO2
Studies in Asia
Kwon etal. (2001)
Seoul, South Korea
PM10
O3, CO, NO2, SO2
April 2002
8-199
DRAFT-DO NOT QUOTE OR CITE

-------
 1      mortality in Los Angeles is also greatly attenuated, but there is substantial variance deflation.
 2      Perhaps CO is either a surrogate or a potential confounder of PM10 in Los Angeles.
 3
 4      Fairley et al. (1999) total mortality in Santa Clara County (San Jose), CA.
 5           This study is noteworthy because it is, as far as we are aware, the only study using particle
 6      nitrates as an exposure index.  The PM10-nitrate component is very stable and shows variance
 7      inflation from NO2 and PM2 5, as might be expected.  The extent to which nitrate is a component
 8      of PM25 rather than PM10 in this study is unknown.  In many western cities, PM25 is much more
 9      alkaline than in the eastern U.S., so that nitrates are less likely to be displaced to the coarse PM10
10      fraction.  In the eastern U.S., where particle acidity is greater,  there may be a greater
11      displacement of nitrates from fine particles where sulfates are a much larger fraction of particle
12      mass than in the west, and consequently nitrates are more likely to reside in the coarse fraction in
13      the east.  It is therefore uncertain that the nitrate component of the atmosphere can account for
14      the large adverse health effects of PM10 observed in many cities in the northeast and industrial
15      midwest.  Clearly, it would be desirable to have more epidemiology  studies with the nitrate
16      component (size-stratified and measured  in a  manner to avoid evaporated losses) used as a PM
17      component in models, notwithstanding technical difficulties that might be encountered in
18      measuring the samples. As shown in Table 8-34, the PM2 5 effect size estimate is almost
19      eliminated by including PM10-nitrates , and the estimated PM25 effect size variance inflated by
20      including the criteria gaseous pollutants.  As shown in Figure  8-24, the PM10-nitrate and PM2 5
21      effect size estimates are stable against gaseous pollutants. It is unlikely that the gaseous
22      pollutants are confounders of PM10-nitrate.
23
24      Other studies
25           The results from other studies are also summarized in Table 8-34 and Figures 8-19 to 8-26.
26      There are numerous examples of effect size instability and variance inflation. The only other
27      cases of substantial variance deflation (i.e., better prediction of PM effect size by inclusion of
28      co-pollutants) are in Burnett et al. (2000), Lippmann et al. (2000), and Goldberg et al.  (2000a).
29      In Burnett et al.  (2000), the model with sulfates as a surrogate for PM2 5, three gases, and four
30      metals in PM2 5 give a better estimate of the sulfate effect than do the models with the gravimetric
31      PM2 5  index. In Lippmann et al. (2000) the models for total mortality with fine or coarse particles

        April 2002                                8-200        DRAFT-DO NOT QUOTE OR CITE

-------
1 give better predictions of the particle effect when the other size component is included than do
2 any of the gaseous co-pollutant models, regardless of the single day lags used for the various PM
3 or gaseous co-pollutants (see Table 8-36).
4
5
TABLE 8-36. SINGLE-DAY LAGS USED IN CO-POLLUTANT MODELS IN
(Lippmann et al., 2000, Tables 13-14)
Pollutant
Endpoint
Total Mortality
Circulatory Mortality
Respiratory Mortality
Pneumonia Admissions
COPD Admissions
Ischemic Heart Disease (IHD)
Admissions
Dysrhythmia Admissions
Heart Failure Admissions
Stroke Admissions
PM10
1
1
0
1
3
2
1
0
1
PM25
3
1
0
1
3
2
1
1
0
PM10_2.5
1
1
2
1
3
2
0
0
1
03
0
0
0
3
3
3
3
3
3
CO
1
1
1
3
3
3
3
3
3
NO2
1
1
1
3
3
3
3
3
3
SO2
3
3
3
3
3
3
3
3
3
 1     8.4.2.3 Assessment of Confounding in Multi-City Studies:  Pooling Effects
 2           One approach to evaluating confounding used in a number of multi-city studies is to pool or
 3     combine the results of individual within-city studies using either standard analytic techniques
 4     such as inverse variance averaging (Atkinson et al., 2001; Katsouyanni et al., 2001) or Bayesian
 5     second-stage meta-analysis methods (Samet et al., 2000a) which may be thought of as another
 6     kind of averaging. The argument is that if the pooled or combined PM effect size estimates for
 7     the single-pollutant models across a number of cities differing greatly in PM and co-pollutant
 8     distributions and correlations are similar to those obtained from PM effect size estimates of
 9     multi-pollutant models across the same cities, then it is unlikely that the co-pollutants are
10     confounding the PM effect.
       April 2002
8-201
DRAFT-DO NOT QUOTE OR CITE

-------
 1           The basis of this argument is not self-evident.  Examination of the results of Samet et al.
 2      (2000a) for the 20 largest U.S. cities discussed in Section 8.2.2.2 shows that there are a variety of
 3      different patterns of change in PM10 effect size associated with including gaseous co-pollutants in
 4      a co-pollutant model.  Results for multi-pollutant models in the NMMAPS Part n study (Samet
 5      et al., 2000a) for the 20 largest U.S. cities are shown in Section 8.4.2.2, Figure 8-21.  The
 6      Bayesian posterior distribution of the estimates was shown earlier in Figure 8-3. It should be
 7      noted that the posterior distribution for the mean PM10 effect has about the same standard
 8      deviation for all of the co-pollutant models, as might be expected by examining Figure 8-21. The
 9      posterior distribution for the mean PM10 effect size estimate remains relatively unchanged from
10      the single-pollutant model when O3 is included as a co-pollutant, and tends to decrease
11      substantially (by about 30%) when either CO or NO2 are added as co-pollutants in addition to O3.
12      Adding SO2 causes a smaller reduction the estimated PM10 effect.  The estimated PM10 effect
13      follows a similar pattern of association with the gaseous criteria pollutants in many large U.S.
14      cities, but shows a very different pattern in other cities, either being stable with respect to co-
15      pollutants or showing increasing effects of PM10 with the inclusion of CO, NO2, or SO2 in a
16      multi-pollutant model for apparently dissimilar cities including Seattle, WA, Phoenix, AZ,
17      Dallas-Form Worth, TX, and Philadelphia, PA. One can argue that CO and NO2 are  often
18      closely and positively associated with PM10, especially if PM10 is dominated by the fine particle
19      fraction, predominantly from combustion, thus reducing the magnitude and significance of the
20      estimated PM10 effect. This explanation is less convincing in cities such as Phoenix,  AZ, where
21      excess mortality is better associated with coarse particles than with fine particles (Mar et al.,
22      2000; Smith et al., 2000; Clyde et al., 2000).
23           Moolgavkar (2000b) finds different effects of PM10 on circulatory mortality in Phoenix,
24      where the single-pollutant and multi-pollutant models  largely agree, disagreeing slightly more in
25      magnitude and significance at lag days 0 and  1, 4 and 5, than at lag days 2 and 3. In general,
26      Moolgavkar finds different temporal patterns of the effect of PM10 for single-pollutant and
27      co-pollutant models among Phoenix, AZ, Chicago, IL, and Los Angeles, CA. There  are many
28      possible reasons for city-to-city variations in these relationships.  One possibility is differences in
29      the mix of fine and coarse particles (which may be quite different in those cities). The combined
30      estimates across cities using the one-day lags for all cities or all regions may overlook different
        April 2002                                8-202       DRAFT-DO NOT QUOTE OR CITE

-------
 1      delays associated with different causes for total or cause-specific mortality in different cities,
 2      suggesting that caution be used when pooling data from different places.
 3           Analogous differences in the stability of the estimated PM coefficients have also been
 4      noted in other city-specific studies discussed in Section 8.4.2.2. Does the occasional instability
 5      of PM coefficients in co-pollutant models across different cities reflect real differences, or is it
 6      merely another kind of statistical variability that might be explained in a second-stage
 7      regression?  Second-stage regression approaches are discussed next in Section 8.4.2.4.
 8
 9      8.4.2.4  Assessment of Confounding in Multi-City Studies:  Regression
10      8.4.2.4.1 Second-Stage Regression and Identification of Effects Modifiers
11           The approach used by Atkinson et al. (2001); Janssen et al. (2002); Katsouyanni  et al.
12      (2001); Levy et al. (2000); and Samet et al. (2000b) is to accept the estimated PM effect-size
13      estimates as samples from a distribution of possible effect sizes in different cities (a "meta-
14      population") and to fit a weighted regression model of the estimated effect sizes on various
15      community-wide indices. The community-wide indices for these studies have included: (a) the
16      mean or median levels of co-pollutants; (b) the median or trimmed mean of the correlations of
17      the PM10 concentrations at different sites across the city (as an index of spatial measurement
18      error); (c) characteristics of particles such as the observed or estimated PM25/PM10 ratio;
19      (d) some characteristics of the distribution of meteorological variables (e.g., mean maximum
20      annual temperature); (e) some community-wide  surrogates for possible sources (e.g., density of
21      vehicle miles traveled, percent of population using public transportation, or PM25/NO2 ratio as
22      indicators of motor vehicle use);  (f) factors possibly affecting exposure (e.g., percentage of
23      residences with air conditioning); and (g) sociodemographic characteristics possibly affecting
24      exposure and susceptibility to particles (e.g., average education level, percentage of residents
25      >64 years of age, measures of immigration or emigration from the community).
26           Samet et al. (2000b) found  that estimated PM10 effects on total mortality (a) increased with
27      increasing mean O3 (not significant in any model), increasing mean NO2 (significant in three- and
28      four-variable models, only marginally significant in a five-variable model), and increasing
29      percentage without a high school diploma, but (b) decreased with increasing mean PM10
30      (significant only in the best five-pollutant model) and the median PM10 cross correlation.
        April 2002                                8-203        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Janssen et al. (2002) found variations among PM10 effects on hospital admissions in single-
 2      pollutant models to be associated with differences in: the community-wide prevalence of air
 3      conditioners; sources of PM10, population density; and density of vehicle traffic (mean daily
 4      vehicle miles of urban travel per square mile).  The same 14 cities used in the hospital
 5      admissions studies in Samet et al. (2000b) were divided into two groups, five in which PM10
 6      concentrations peaked in the winter (Boulder and Colorado Springs, CO; Provo-Orem, UT;
 7      Seattle and Spokane, WA) and nine others where PM10 concentrations peaked in the summer
 8      (Birmingham, AL; Canton and Youngstown, OH; Chicago, IL; Detroit, MI; Minneapolis, MN;
 9      Nashville, TN; New Haven, CT; Pittsburgh, PA). There was a statistically significant negative
10      relationship between the percentage of homes with central  air conditioning and the regression
11      coefficient (excess relative risk) for cardiovascular disease, this being lower in winter-peaking
12      cities than in summer-peaking cities.  The relationship between exposure and ventilation rate is
13      discussed further in Section 8.4.2.5.  Ventilation rate also effects personal exposure to gaseous
14      co-pollutants. Additional studies similar to Jannsen et al. (2002) would likely help clarify further
15      the associations between particles and gases and may offer a useful alternative method for
16      assessing multi-pollutant models across different cities.
17           Katsouyanni et al. (2001) found that PM10 effects on mortality increased with mean NO2
18      level, mean temperature, and percentage of population with age > 65 years. The estimated PM10
19      effects decreased with increasing PM10/NO2 ratio, increasing relative humidity, and increasing
20      age-adjusted mean mortality rate in the 29 cities in the APHEAII study in Europe.
21           Atkinson et al. (2001) studied respiratory mortality in eight European cities and found a
22      significant positive relationship between asthma mortality at ages 0 to 14 years and the percent of
23      population > age 65, a significant negative relationship with community smoking prevalence,  and
24      a negative relationship with relative humidity.  In the analyses for age 65+ total respiratory
25      mortality,  and age 65+ mortality from asthma and COPD, effect size increased significantly with
26      larger mean O3.
27           These studies, while quite informative, do not address the core issue in assessment of
28      potential confounders: in any single city, it is often difficult to disentangle the components of the
29      mixture of air pollutants that actually exists. The problem  derives from the typically co-linear
30      relationship of particles and gaseous co-pollutants, where co-pollutants may have relatively high
31      linear correlation coefficients among themselves. The assessment of confounding depends on the

        April 2002                                8-204        DRAFT-DO NOT QUOTE OR CITE

-------
 1      correlations among the pollutants, not on their absolute levels. It would be possible for
 2      co-pollutants to have exactly the same correlations with PM even if the absolute concentrations
 3      of the co-pollutants differed greatly from one city to the next. For this reason, even if the
 4      co-pollutant concentrations are very low, co-linearity could still be manifested in a
 5      multi-pollutant model because the co-pollutant would increase and decrease in step with the PM
 6      index due to common meteorological conditions. For example, in a recent study of respiratory
 7      symptoms in Port Alberni, BC, Canada, the levels of SO2 and other pollutants are low, and are
 8      less likely to be the cause of the observed effects (so not a confounder), but may complicate a
 9      multi-pollutant model because of their co-linearity with PM.
10           Another issue is whether the mean or median PM or co-pollutant concentration is the best
11      covariate in a second-stage model.  An indicator of high-level concentrations might be
12      informative, e.g., the mean or median of 95th percentiles for each year (approximately the 3rd
13      largest of 61 annual  every-6th-day observations for PM and the 18th largest for 365 daily measures
14      of the co-pollutants).
15
16      8.4.2.4.2 Regression of Effect Size Coefficients on Co-Pollutant vs. PM Coefficients in a
17               Multi-City Study
18           Several authors (Samet et al., 2000b;  Schwartz, 1999, 2000a) have applied two-stage
19      regression techniques in an effort to identify the  extent to which a pollutant is directly associated
20      with human health effects, as opposed to acting indirectly through its association with other
21      pollutants.  Noting that relationships between co-pollutants differ by city, it has been proposed
22      that such differences can facilitate the separation of direct and indirect effects through use of a
23      second-stage meta-regression.  The second-stage meta-regression approach regresses the city-
24      specific PM regression coefficients (which may have been adjusted for individual-level or time-
25      varying covariates in a previous stage) against the city-specific correlations between PM and a
26      selected co-pollutant.  Typically the second-stage regression model is limited to a single
27      independent variable, and the regression results are presented graphically along with the  data
28      points. For the results of the second-stage regression, a non-zero slope is taken as evidence of
29      confounding by the co-pollutant, while a non-zero intercept is taken as evidence of a true,
30      unconfounded association with PM.
31           Marcus and Kegler (2001) demonstrate that the use of the intercept as an  indicator  an
32      unconfounded association will only work if the second-stage model has been correctly specified.
        April 2002                                8-205        DRAFT-DO NOT  QUOTE OR CITE

-------
 1      Their counter-example is simply the case where the PM association is independently confounded
 2      by two co-pollutants and only one of these confounders is included in the second-stage regression
 3      model. In this case, the PM association would still be confounded by the omitted co-pollutant
 4      and a non-zero intercept from the mis-specified model would not be valid evidence of an
 5      unconfounded association with PM.
 6           The counter-example is an illustration of residual confounding which is already well-
 7      understood by epidemiologists. Residual confounding may arise (1) when a regression model
 8      fails to include all of the potential confounders or (2) when a regression model includes a poorly
 9      measured covariate that captures only a portion of the confounding by that covariate. The
10      approach to the evaluation of confounding proposed by Schwartz and Samet et al. could be
11      improved by the simple expedient of use of a multi-variate second-stage regression,  in which the
12      model includes several gaseous co-pollutants. Even a multi-variate model would not fully
13      resolve the issue of residual confounding in the case of (a) poorly measured covariates or (b)
14      other omitted covariates. In either of these cases, the use of a non-zero intercept as evidence of a
15      true, unconfounded association with PM would be incorrect.
16           However, the most important aspect of the approach proposed by Schwartz and Samet et al.
17      was not their use of the intercept, but rather their use of the second-stage slope as an indication of
18      both the presence and magnitude of confounding by the modeled co-pollutant. The simulations
19      illustrated in the article by Marcus and Kegler clearly demonstrate that a confounding
20      co-pollutant would produce a non-zero slope, even in the presence of omitted confounders.
21      In addition to confounding, a non-zero slope might, under certain circumstances, also be
22      evidence of effect modification of a true PM association by the modeled co-pollutant.
23      Conversely, the absence of a slope in the second-stage regression could be evidence that the
24      particulate matter association is neither confounded nor modified by the modeled co-pollutant, or
25      that positive slopes could be obscured by negative slopes in relationships among co-pollutants,
26      suggested by the observation that the longitudinal correlations between PM10 and O3 or between
27      NO2 and O3 are often negative.
28
29
30
        April 2002                                8-206       DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.4.2.5 Assessment of Confounding Based on Exposure
 2      8.4.2.5.1  Review of Sarnat et al. (2001):  Is there significant personal exposure to gaseous
 3               co-pollutants?
 4           A direct method for evaluating whether a putative causal factor is confounded by another
 5      factor is based on the requirements that a confounder be (1) associated with the health outcome
 6      or disease, and (2) associated with exposure to the putative causal factor. If individuals are not
 7      exposed to a potential confounder, then it cannot be a confounder of another agent to which the
 8      individual is exposed, although there is no guarantee that the putative causal factor causes the
 9      outcome unless there is evidence of biological effects at the levels to which the individual is
10      exposed.  The most likely potential confounders in air pollution epidemiology studies are the
11      gaseous criteria pollutants,  specific size components or  chemical components of particles, and
12      meteorological variables associated with exposure to PM or other pollutants. Many of these
13      potential confounding factors have been shown to cause adverse cardiovascular or respiratory
14      effects after exposure to elevated levels in laboratory animal studies, in in vitro experimental
15      studies, or as small physiological or functional changes  in human  adult volunteers. There is also
16      evidence for toxic effects from either short-term or long-term exposures to other ambient air
17      pollutants with which the criteria pollutants are associated. Finally, while extremes of
18      temperature and humidity are known to be independently associated with increases in mortality
19      and morbidity, they are also associated with concentrations and possibly even exposures (e.g., by
20      closing the windows and using air conditioners) to these pollutants, as well  as ambient particles.
21      Thus, the gaseous co-pollutants and other environmental variables cannot be totally precluded as
22      confounders on the basis of lack of effects independent  of PM.
23           The question raised in two important papers by Sarnat et al.  (2000, 2001) addresses the
24      exposure  aspect of confounding, i.e., are the  gaseous pollutants confounders or surrogates of PM
25      effects?  The first study (Sarnat et al., 2000)  enrolled 15 non-smoking elderly participants (age
26      65+, average 75 years) in Baltimore, MD, who wore multi-pollutant personal samplers during the
27      summer of 1998 and the winter of 1999. Selection of participants was non-random, as they all
28      were healthy (i.e., asymptomatic)  non-smokers, living in private residences, all with central air
29      conditioning but one (denoted SA4). The participants came from  a range of socio-economic
30      backgrounds and locations  within Baltimore  (details not reported). The ambient pollutant
31      concentrations were measured at seven state  or federal monitoring sites (shown in Chang et al.,
32      2000, Figure 2), five in the city of Baltimore and two in suburban counties,  all within nine miles
        April 2002                                8-207        DRAFT-DO NOT QUOTE OR CITE

-------
 1      of the central business district site denoted CBD1. Longitudinal within-subject correlations of
 2      personal and ambient PM2 5 were high in the summer (median Spearman r = 0.74) and low in the
 3      winter (median r = 0.25), and residential ventilation was an important determinant of this
 4      association (high in well-ventilated environments, low in poorly-ventilated environments). The
 5      highest association, on average, was between personal and ambient SO4=, given that SO4= has few
 6      indoor sources.  There were successively smaller mean correlations between personal and
 7      ambient concentrations for PM2 5, PM10, O3, NO2, and PM10_2 5 in the summer, and PM10, PM2 5,
 8      PM10_2 5, O3, NO2, SO2 in the winter, as shown in Table 3-6 of Sarnat et al. (2000). The lower
 9      correlations between personal and ambient concentrations of the gaseous pollutants may reflect
10      both the much greater spatial heterogeneity of the gaseous pollutants (except possibly for ozone),
11      and the fact that many gaseous pollutant concentrations were below the seasonal detection limit
12      in Baltimore, producing negative median concentrations for some participants.
13           The study design should be considered in some detail.  Each study period was divided into
14      three 12-day segments (denoted A, B, C) , with participants providing n = 9 to 12 days of
15      personal exposure data from which the longitudinal correlation coefficients for each participant
16      were calculated. The participants in each successive 12-day block or group were different
17      individuals with different residence locations, possible patterns of behavior, and household
18      characteristics that may have affected exposure. The number of participants (N) in each block is
19      shown in Table 8-37. One might hypothesize no difference among blocks in this small sample.
20           The results presented in Sarnat et al. (2000) have been aggregated over these three waves or
21      blocks. In spite of the admittedly very small numbers, one might ask whether the aggregation or
22      pooling across blocks is appropriate, given that the participants in each block are independent of
23      each other.  The evaluate this, we transformed the correlation data to more nearly normally
24      distributed observations with constant variance, as if the Spearman correlation coefficient r was a
25      Pearson coefficient, using the Fisher Z-transformation, Z = 0.5 (ln(l + r ) - ln(l - r )). We then
26      tested the hypothesis that the mean values of the transformed personal  to ambient correlations
27      were the same in blocks A, B, C on average, using a standard analysis of variance test.
28           There were no significant differences among the correlation between summertime personal
29      exposure and ambient air pollution among blocks A, B, C for pollutants that are believed to have
30      a reasonably uniform spatial distribution in summer, including SO4=, PM25, PM10, and O3. The
31      PM10_2 5 difference is also very non-significant if SC5 is included in the data, but becomes

        April 2002                                8-208        DRAFT-DO NOT QUOTE OR CITE

-------
            TABLE 8-37. NUMBER OF PARTICIPANTS, N, IN EACH BLOCK FOR THE
                            EXPOSURE STUDY IN SARNAT ET AL. (2000)
Summer 1998
Block
A
B
C
Total
Approximate Dates
June 3 0 - July 12
July 14 - July 25
July 27 - August 7
June 30 - August 7
Number N
O
6
5
14*
Winter 1999
A
B
C
Total
February 2 - February 13
February 16 - February 27
March 2 -March 13
February 2 - March 13
4
4
6
14*
         One of the 15 participants was excluded because of high exposure to environmental tobacco smoke outside
         the residence.
 1      statistically significant when PM10_2 5 from this participant is excluded.  Differences among the
 2      correlation between summertime personal exposure and ambient NO2 among blocks A, B, C is
 3      nearly significant, even though there is a larger within-block variance for the other pollutants,
 4      consistent with NO2 having a somewhat non-uniform spatial distribution.
 5           We note significant between-block differences among the correlation between wintertime
 6      personal exposure and ambient air pollution among blocks A, B, C for pollutants that had a
 7      reasonably uniform spatial distribution in summer, including SO4= and PM10 in spite of the small
 8      numbers and large within-block variance of personal correlations.  There was also greater
 9      evidence for between-block differences in the wintertime correlation between total personal
10      exposure and ambient PM2 5 concentration (P = 0.11 for Z) than  for the summertime correlation
11      (P = 0.24). There was little indication of wintertime temporal variability among the personal -
12      ambient correlations for NO2, SO2, and  O3, the ambient concentrations of which were often
13      below the wintertime detection limit.  The PM10_2 5  interblock difference in person al vs. ambient
14      correlation is also very non-significant.  We believe it is useful to recognize that large differences
        April 2002                                8-209       DRAFT-DO NOT QUOTE OR CITE

-------
 1      among people and temporal blocks of personal-ambient correlation coefficients for gases and
 2      particles may require personal exposure studies with a sufficient number of participants and of
 3      longer duration to establish those relationships with a high degree of statistical certainty. This
 4      analysis would be meaningful only if one block was high and significant and another was low
 5      and insignificant.
 6           The correlation of personal PM25 exposure to ambient concentrations of other pollutants
 7      was also reported in Table 7 of Sarnat et al. (2000). The median correlation of personal PM25
 8      exposure with ambient PM25 was higher than the correlation of personal exposure to PM2 5 with
 9      any of the ambient concentrations of PM10_25, O3, NO2 in summer, but actually smaller in winter
10      with PM10_25 and NO2.  The correlation of winter PM25 exposure and ambient O3 is very negative.
11           It would also have been useful to have ambient measurements at locations near the
12      participants' residences, in  order to determine whether the difference in strength of association is
13      related to the heterogeneity in the spatial and temporal distribution of co-pollutant gases  and
14      coarse particles relative to central site monitors versus the comparative (but not absolute)
15      homogeneity of PM25 measurements, a "measurement error" problem reviewed in
16      Sections 8.4.5.3 and 8.4.7.2.  Finally, it would have also been desirable to have reported  the
17      correlations between personal exposure to each of the gaseous co-pollutants versus the ambient
18      concentration of PM25 for each participant, analogous to Table 6.  This would have allowed a
19      more direct comparison of the hypothesis in Sarnat et al. (2001) reviewed next in
20      Section 8.4.2.5.2.
21
22      8.4.2.5.2  Review of Sarnat et al. (2001): Are gaseous pollutants confounders or surrogates?
23           The Sarnat et al. (2001) paper extends the results in Sarnat et al. (2000) to additional
24      cohorts and participants: (a) 21 healthy children, ages 9 to 13 years; (b)  15 individuals with
25      COPD, average age 65 years; and (c) a total of 20 older healthy adults of average age 75  years,
26      6 more than in the earlier paper. All participants were non-smokers who lived nonsmoking
27      private residences. Fourteen of the healthy adults participated in both summer and winter
28      sampling campaigns described above. The COPD cohort consisted of individuals with
29      physician-diagnosed COPD, with an average age younger than the healthy adult cohort.  The
30      sampling plan is shown in Table 1 of Sarnat et al. (2001). Although the participants lived in
        April 2002                                8-210        DRAFT-DO NOT QUOTE OR CITE

-------
 1      various parts of Baltimore city and County and had a range of socio-economic backgrounds, they
 2      were not selected as a representative sample of susceptible sub-populations.
 3           The Sarnat et al. publication reports on personal monitoring of 56 subjects for fine PM, O3,
 4      NO2, and SO2 in comparison with ambient concentrations of these substances. For fine PM, the
 5      personal measurements are associated with the ambient fine PM central-site measurements, O3
 6      (negative in winter), NO2, CO (winter only), and SO2 (winter only and negative). For the gaseous
 7      co-pollutants, the various personal measurements are not positively associated with the ambient
 8      central-site measurements of the same gas. The authors conclude that the ambient, central-site
 9      measurements of the gaseous co-pollutants may be surrogates for specific constituents of fine PM
10      rather than confounders.
11           Among the combined sample of 56 participants, the highest median correlation between
12      personal exposure and ambient concentration was for particle sulfates, a component of
13      predominately ambient origin.  Sulfates had a summertime median Spearman correlation of 0.88,
14      13 correlations being significant out of 14 for older healthy adults, and a wintertime median
15      correlation of 0.71, 16 out of 29 being significant correlations including 14 healthy adults.  The
16      median Spearman correlation between total personal exposure and ambient concentration was
17      also high for PM25, with a median Spearman correlation of 0.65, 13 significant correlations
18      among 24 healthy older adults  and children combined, and a wintertime median correlation of
19      0.22, with 10 of 44 significant  correlations using data combined from the three cohorts. Among
20      the gaseous co-pollutants, the personal-ambient correlation was highest for NO2, with 7 out of
21      44 significant correlations using data combined from the three cohorts.
22           The ambient pollutants are correlated as expected, with high positive summertime
23      correlations seen between the regionally more correlated pollutants PM2 5 and O3, between two
24      combustion products NO2 and  CO, and a positive significant correlation between ambient PM2 5
25      and NO2 shown in Table 8-38.  There are high negative wintertime correlations between the
26      regionally correlated pollutant  O3 and the combustion products PM2 5, NO2, or CO, and high
27      positive correlations among the combustion products PM2 5, NO2, and CO.
28           Total personal exposure to PM2 5 and exposure to estimated ambient PM2 5 is not
29      significantly correlated with personal exposure to gases, except for NO2 exposure in summer.
30      Sarnat et al. (2001) expressed this as a linear regression, with personal PM2 5 the dependent
31      variable and personal exposure to NO2, O3, or  SO2 as the independent variable, finding that:

        April 2002                                8-211       DRAFT-DO NOT QUOTE OR CITE

-------
               TABLE 8-38. CORRELATIONS AMONG AMBIENT POLLUTANTS IN
               BALTIMORE. SUMMERTIME CORRELATIONS IN UPPER RIGHT,
        WINTERTIME IN LOWER LEFT.  STATISTICALLY SIGNIFICANT SPEARMAN'S
                CORRELATIONS ARE SHOWN AS UNDERLINED BOLD VALUES.
Pollutant
PM25
03
NO2
CO
SO2
PM25
1
-0.72
0.75
0.69
-0.17
03
0.67
1
-0.71
-0.67
0.41
NO2
0.37
0.02
1
0.76
-0.17
CO SO2
0.15 —
-0.06 —
0.75 —
1
-0.12 1
        Source: Based on Table 3 in Sarnat et al. (2001).
 1                (Personal exposure to PM2 5) = 18.65 + 0.18 (Personal exposure to NO2)      (8-4)
 2
 3     in summer, with both slope and intercept terms statistically significant and NO2 measured in
 4     units of parts per billion (ppb). It is likely that the relative measurement error of PM25 exposure
 5     is smaller—possibly much smaller - than that of personal exposure to the gaseous pollutants, as
 6     shown from Tables 1, 4, and 5 in Sarnat et al. (2000). Among the 14 healthy older adults in the
 7     earlier study, only 3 had mean summer exposure concentrations for O3 greater than the
 8     summertime limit of detection (LOD) of 6.6 ppb (SC2, SC4, SC5, all in block  C), only 3/14 of
 9     the mean NO2 personal exposures below the LOD (SA1, SA2, SB3) of 5.5 ppb, but all of the
10     mean PM25 exposure concentrations were 5 to 12 times larger than the summertime LOD of 2.6.
11     None of the wintertime mean O3 exposure concentrations is larger than the LOD, and 5 of the
12     6 mean O3 exposures in block C are negative. Only 2 of 14 wintertime mean NO2 exposures is
13     smaller than the LOD (WAI, WB1) of 11.7 ppb, whereas all of the PM25 personal exposures
14     means are above the LOD, some by a factor of 12 to 13.
15          Figure 2 in Sarnat et al. (2001) shows box plots of the distribution of Spearman correlations
16     between personal and ambient concentrations for individual participants.  In the summer, the
17     median correlation between personal O3 and ambient O3 is quite low (left-hand box) and only one
18     correlation is statistically significant, much lower than the median correlation between personal

       April 2002                               8-212       DRAFT-DO NOT QUOTE OR CITE

-------
 1      O3 and ambient PM2 5 (right-hand box) where five correlations are statistically significant.
 2      However, while the summertime median correlation between personal NO2 and ambient NO2 is
 3      also quite low (left-hand bar) and only three correlations are statistically significant, it is not
 4      much lower than the median correlation between personal NO2 and ambient PM2 5 for which four
 5      correlations are significant. The median correlations are, of course, much higher for ambient
 6      PM2 5 vs. ambient O3 or NO2. Thus, ambient PM2 5 may be a better proxy for personal exposure
 7      to O3 than is ambient O3. However, personal exposure to NO2 is almost as well correlated with
 8      ambient NO2 as with ambient PM25. If it is believed that exposure to NO2 also causes adverse
 9      health effects, along with exposure to PM2 5, then it is not clear that ambient PM2 5 is merely a
10      proxy for ambient NO2.
11           In the winter, the median correlation between personal O3 and ambient O3 is quite low
12      (left-hand box) and no correlations are statistically significant, whereas the wintertime median
13      correlation between personal O3 and ambient PM25 (right-hand box) where seven correlations are
14      statistically significant and negative. While the wintertime median correlation between personal
15      NO2 and ambient NO2 is also quite low (left-hand bar), six correlations are positive and
16      statistically significant, the median personal-ambient NO2 correlation is not much lower than  the
17      median correlation between personal NO2 and ambient PM2 5 for which four correlations are
18      significant.  The median correlations are, of course, much higher for ambient PM2 5 vs. ambient
19      O3 or NO2. Thus, ambient PM2 5 may be a better proxy for personal exposure  to O3 than is
20      ambient O3. However, personal exposure to NO2 is about as well correlated with ambient NO2 as
21      with ambient PM25. If it is believed that exposure to NO2 also causes adverse health effects,
22      along with exposure to PM2 5, then it is not clear that ambient PM2 5 is merely  a proxy for ambient
23      NO2. Wintertime personal exposure to SO2 tends to be negatively associated with both  ambient
24      SO2 and ambient PM2 5, with similar median correlations, but with one significantly negative
25      correlation against ambient SO2 vs. four significantly negative correlations against ambient
26      PM2 5. Thus, ambient PM2 5 may be a surrogate for personal SO2 exposure.
27           CO personal exposures have little correlation with ambient PM2 5 in summer, with only  two
28      statistically significant correlations, but may be more strongly associated with ambient PM2 5 in
29      winter, showing five positive and one negative significant correlation and a larger positive
30      median correlation. However,  there is no distribution of correlations of personal vs. ambient  CO
31      with which to compare these findings.

        April 2002                                8-213        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Personal exposures to O3 are more positively correlated with personal exposures to PM2 5 in
 2      summertime, and more negatively correlated in wintertime, than are personal O3 exposures with
 3      ambient O3, and the associations of personal O3 with ambient PM25 even more so, suggesting that
 4      ambient PM2 5 may be a good surrogate for personal O3. On the other hand, summertime
 5      personal NO2 is no more closely associated with personal PM2 5 than with ambient NO2, and only
 6      slightly less so than with ambient PM2 5, so that even if personal PM2 5 measurements were
 7      available, one would expect them to be no more informative about NO2 personal exposures than
 8      would ambient NO2. Wintertime personal NO2 is not associated with personal PM2 5, and about
 9      equally associated with ambient NO2 and with ambient PM2 5, so that even if personal PM25
10      measurements were available, one would expect them to be completely uninformative about NO2
11      personal exposures compared to ambient NO2.
12           Table 9 in Sarnat et al. (2001) contains extensive results on the wintertime linear
13      relationships of personal total PM25 exposure, exposure to estimated PM25 of ambient origin,
14      personal SO4= exposure, and personal elemental carbon (EC) exposure, versus ambient
15      concentrations of the gaseous co-pollutants.  Many of these are statistically significant: negative
16      relationships of exposure to all of these particle components vs. ambient O3, positive
17      relationships of personal exposure to EC and to estimated PM25 of ambient origin versus NO2 or
18      CO, and negative relationships of exposure to SO4= and estimated PM2 5 of ambient origin versus
19      ambient SO2.  However, as noted above, and as done by the authors in Figure 2, comparison of
20      personal exposures of the gaseous co-pollutants to these particle components might be more
21      useful in evaluating the question proposed by the authors:  are gases confounders or surrogates of
22      fine particles or fine particle components?
23
24      8.4.2.5.3  Confounding of co-pollutant effects arising from the spatial distribution of particles
25               and gases.
26           We focus here mainly on the question of co-pollutant exposures as a potential confounder
27      of PM exposure, identified in the previous sections. There is a component of total exposure to
28      NO2, CO, and other pollutants derived from ambient air most easily detected in the vicinity of
29      strong sources, such as the large number of automobiles and heavy-duty vehicles on major
30      highways or trunk roads near the exposed populations. The link to human health effects, if any,
31      would require satisfying three steps:  (a) co-pollutant concentrations are high near certain line or
32      point sources, and decrease with increasing distance from the source much more rapidly than
        April 2002                                8-214        DRAFT-DO NOT QUOTE OR CITE

-------
 1      does the PM10 or PM2 5 concentration decrease with increasing distance; (b) humans residing near
 2      strong sources are more highly exposed to co-pollutants than those living farther away; and
 3      (c) increased risks of adverse health effects attributable to the co-pollutants occur in proximity to
 4      stronger sources of the co-pollutants. For this reason, one may attribute some the difficulties in
 5      interpreting the findings of multi-pollutant PM epidemiology models including co-pollutants as
 6      spatial "measurement errors" associated with the non-uniform distribution of the co-pollutants in
 7      an urban area. Thus, there may be a reduced likelihood that central site co-pollutant monitors
 8      will accurately characterize population exposure compared to the ability of central site PM
 9      monitors to characterize PM population exposure.
10           Recent reviews about traffic-oriented concentration gradients and exposures to other
11      pollutants have been published by van Wijnen and van der Zee (1998) and by Monn (2001).
12      Location effects are  particularly noticeable at the neighborhood level if there is a strong line
13      source (arterial street or freeway,  for example) or point source (fossil-fuel-burning power plant,
14      for example) near the micro-environment. Some studies provide quantitative relationships
15      between distance from a heavily traveled roadway and concentrations of various pollutants (van
16      Wijnen and van der Zee, 1998).
17           The spatial correlations for PM2 5 were generally higher than for those of PM10 in the
18      PTEAM Riverside study (Clayton et al., 1993; Wallace, 1996) and in Philadelphia (Burton et al.,
19      1996; Wilson and Suh, 1997). However, larger spatial variations may occur for particles with
20      important local sources, such as highways carrying a large number of diesel trucks.  Where PM10
21      is dominated by coarse particles, substantial variations (±20%) occurred between pairs of
22      monitors within 4 to 14 km in California's San Joaquin Valley (Blanchard et al., 1999).  Several
23      European studies have found modest variations in ambient PM10 concentrations for residences
24      and housing close to roadways (Kingham et al., 2000, for Huddersfield, U.K.; Monn et al., 1997,
25      for Zurich), but large differences in NO2 concentrations occurred within a few meters of a Swiss
26      street during summer (Monn  et al., 1997). The Monn results are shown as Figure 8-27.  The
27      location and behavior of participants in a personal exposure study,  as well as the spatial aspects
28      of socioeconomic differences in exposure (Rotko et al., 2000), may be important in defining
29      differences in exposure among sub-populations in an epidemiology study.
30           Numerous studies on personal exposure to airborne particles  are discussed in  detail in
31      Chapter 5. There is  little doubt that elevated ambient concentrations of sulfates and fine particles

        April 2002                                8-215        DRAFT-DO NOT QUOTE OR CITE

-------
                                     50
                                     40
                                  o
                                  to
                                  §  30
                                  O
                                     20
                                          . N02 Summer
                                            PM10 Summer
                                             20
    40
                                                            J_
60   80
       Figure 8-27. Concentration of PM10 and NO2 versus distance.
       Source: Monn et al. (2000).
 1     are closely related to elevated personal exposures to fine particles of ambient origin and to
 2     elevated personal exposure to sulfates. In general, concentrations of ambient fine particles and
 3     sulfates are more closely correlated to distance from a major highway or other sources than are
 4     PM10_2 5, NO2, and CO, whose concentrations decrease with increasing distance. Rotko et al. also
 5     reported that PM2 5 was more uniformly distributed in Helsinki than was NO2.
 6           Janssen et al. (2001) evaluated personal indoor and outdoor NO2 and PM2 5 concentrations
 7     at 24 schools located within 400 m of 22 different stretches of freeway in the Netherlands.
 8     Indoor PM2 5 exposure was correlated with the distance from the school to the freeway and was
 9     moderately correlated with the truck traffic volume, but not with  the total or car traffic volume.
10     Indoor NO2 concentration was significantly associated with car traffic volume and with percent
11     of time downwind, but not with distance from the freeway or with truck traffic volume. Outdoor
12     NO2 concentration was significantly correlated only with percent of time downwind from the
13     freeway. PM25 concentrations indoors and outdoors were both significantly correlated with truck
14     traffic and distance from the freeway, but not with car traffic (at P < 0.05) or downwind
15     percentage.
       April 2002
8-216
 DRAFT-DO NOT QUOTE OR CITE

-------
 1           Traffic-oriented health effects were reviewed by Wjst. et al. (1993) and by van Wijnen and
 2      van der Zee (1998).  New studies have appeared since then, including those by Venn et al. (2001)
 3      and Roemer and van Wijnen (2001).  The study by Roemer and van Wijnen is notable for three
 4      reasons: (i) the endpoint is total mortality, as a more serious outcome than respiratory symptoms
 5      in children; (ii) the populations in the study were divided into a "traffic" population living along
 6      roads with traffic volume greater than 10,000 vehicles per day (about 10 percent of the total
 7      population of Amsterdam) and a background population; (iii) measured air pollution
 8      concentrations of BS, PM10, NO2, NO, CO, SO2, and O3 (8-hour mean) were available for the
 9      background populations, and BS, NO2, NO, CO, SO2, and O3 for the "traffic" populations. All of
10      the pollutant concentrations except for O3 were higher for the "traffic" sites than for the
11      background sites.  The excess risk rates in  Table 3 in Roemer and van Wijnen (2001) for Black
12      Smoke (lags 1 and 2) were much higher and, for NO2 (lag 1) somewhat higher, than in the traffic
13      population. However, the statistically significant risk rates for the total population using the
14      "traffic" air pollution sites was lower than  those using the background sites. No results were
15      reported for risk rates for the "traffic" population using traffic monitor sites, but the background
16      monitor concentrations were moderately correlated with those at the traffic monitoring sites.
17
18      8.4.2.6  Assessment of Confounding by Factor Analysis
19           How can one assess confounding in a single-city study if PM and its gaseous co-pollutants
20      are inextricably mixed in the urban atmosphere? One possibility is to make use  of the correlation
21      structure of the data by extracting its principal components, or factors if the principal components
22      are rotated to provide a clearer picture of the main components. The principal components (p.c.)
23      or factors are linear combinations of the pollutant concentrations and are exactly independent for
24      p.c. and  nearly so for rotated factors.  An important advantage of this method is that there is no
25      problem of instability when all or most p.c. or factors are used in a multi-p.c. model. Several
26      variations of this approach have been used in PM epidemiology studies.  Ozkaynak et al. (1996)
27      studied the relationship between mortality, air pollution, and weather using factor analysis
28      methods, where the factors were constructed from a particle index CoH, CO, and weather
29      variables.  The analyses in Laden et al. (2000) and Tsai et al. (1999, 2000) used factors based on
30      chemical elements in fine particles for each day for which there was a fine particle filter sample.
31      Laden et al. (2000) identified up to seven sources in six cities, with some differences in the target

        April 2002                                8-217        DRAFT-DO NOT QUOTE OR CITE

-------
 1      elements for sources across the cities. Three sources were present in all cities (mobile sources,
 2      coal combustion, crustal particles). Poisson regression models were fitted to mortality data with
 3      all source factors included simultaneously. Positive and statistically significant effects were
 4      found for the motor vehicle and coal combustion sources in most cities, but not the crustal
 5      source.  These studies did not include gaseous pollutants.
 6           Mar et al. (2000) developed a factor-analytic model for Phoenix, AZ, based on
 7      12 components of particles and three gaseous pollutants (CO, NO2,  SO2) measured on the same
 8      day.  They reported relative risks of total and cardiovascular mortality for five factors,
 9      representing: motor vehicle exhaust and resuspended road dust; soil; vegetative burning; local
10      sources of SO2; and regional  sulfate. The results reported in Tables 9 of Mar et al. (2000) are for
11      single-factor models.  The authors state that regression analysis with all of the factors included in
12      a multisource model produced similar results.
13           This is a promising approach to analyzing multi-pollutant models and, in principle, could
14      be used in other studies with particles and gaseous criteria pollutants, even if little or no chemical
15      composition data were available. At the very least, evaluating the principal components of a
16      particle-gas mixture might help to identify which combinations of particles and gases are most
17      difficult to separate statistically in a regression analysis.
18
19      8.4.2.7  Simulation Analysis of Confounding
20           Since no single model specification can a priori be designated as "correct" in addressing
21      confounding effects of co-pollutants, discrepancies in results among studies, even for the same
22      dataset, are to be expected. While any assessment of relative "adequacy" of these alternative
23      model specifications is difficult with observational data, the implication of "inadequate" model
24      specifications may be studied through simulations using synthetic data in which the "correct"
25      model is known.  Chen et al.  (1999) conducted such simulations using a synthetic data set in
26      which the causal variables are known, and the effects of model misspecification were studied in
27      the presence of two variables (xx and x2), with varying levels of correlation, in a Poisson model.
28      They considered three situations: (1) model under/it, in which mortality was generated with both
29      Xj  and x2, but regressed only on xt; (2) model overfit, in which mortality was generated with only
30      xl3 but regressed on both x1 and x2; and (3) model misfit., in which mortality was generated with
31      either xt or x2 but regressed on the other variable.  They observed that the confounding of

        April 2002                                 8-218        DRAFT-DO NOT QUOTE OR CITE

-------
 1      covariates in an overfltted model does not bias the estimated coefficients but does reduce their
 2      significance; and that the effect of model underfit or misfit leads not only to erroneous estimated
 3      coefficients but also to erroneous significance. Based on these observations, Chen et al.
 4      suggested that "models which use only one or two air quality variables (such as PM10 and SO2)
 5      are probably unreliable, and that models containing several correlated and toxic or potentially
 6      toxic air quality variables should also be investigated...". While conceptually useful, this
 7      simulation study ignored one factor that is crucial in evaluating the implication of confounding,
 8      the relative error.  For example, including several correlated  pollutants in a regression model may
 9      lead to erroneous inferences, unless one considers the relative error associated with each of the
10      pollutants.
11
12      8.4.2.8 Discussion
13           In the preceding several  section, a number of methods for evaluating the potential for
14      gaseous co-pollutants to confound particle effects were discussed. Multi-pollutant models may
15      be sensitive to multi-colinearity (high correlations among particle and gaseous pollutant
16      concentrations), and to so-called "measurement errors", possibly associated with spatial
17      variability. Combining multi-pollutant models across several cities may not improve the
18      precision of the mean PM effect size estimate combined, if the differences among the cities is as
19      large or larger in the multi-pollutant models as in the single-pollutant PM model.  Second-stage
20      regressions have been useful in identifying effect modifiers in the NMMAPS and APHEA 2
21      studies, but may not, in general, provide a solution to the problem that confounding of effects is a
22      within-city phenomenon. Furthermore, the correlations among pollutants may change from
23      season to season and from place to place, suggesting that confounding as indicated by co-
24      linearity is not always the same.
25           Two promising approaches are also discussed, the first based on personal exposures to
26      particles and gases of three panels of participants in Baltimore, MD (Sarnat et al., 2000, 2001).
27      This directly addresses the premise that if individuals are not exposed to a potential confounder,
28      then it cannot really be a confounder of the presumed causal  effect. While the results in this
29      paper support the conclusion that personal exposure to sulfates, fine particles, and PM10 are well
30      correlated with the corresponding fixed site ambient concentrations, the correlations are much
31      lower for PM10_2 5, O3, and NO2.  There is however a great deal of variation from one of three

        April 2002                                8-219       DRAFT-DO NOT QUOTE OR CITE

-------
 1      two-week panels from one season to the next. The sample size is small (N = 56), but did detect
 2      marginally significant associations between personal and ambient NO2 for the personal-ambient
 3      correlation, although much lower than for particles. There were, however,  a number of
 4      residences in which personal and ambient NO2 were highly correlated. This has been known to
 5      happen in other studies when the residences are close to a major road, which describes several
 6      members in the three cohorts (health elderly adults, adults with COPD, children 9-13 years.)
 7           The other promising approach is to use principal component or factor analysis to determine
 8      which combinations of gaseous criteria pollutants and PM size fractions or chemical constituents
 9      together cannot be easily disentangled, and which pollutants are substantially independent of the
10      linear combinations of the others. For example, Mar et al. (2000) shows independent effects of
11      regional sulfate, motor vehicle-related particles, particles from vegetive burning, and PM10_25 for
12      cardiovascular mortality in Phoenix.
13
14      8.4.3  Role of Particulate Matter Components
15           In the 1996 PM AQCD, extensive epidemiologic evidence substantiated very well positive
16      associations between ambient PM10 concentrations and various health indicators, e.g., mortality,
17      hospital admissions, respiratory symptoms, pulmonary function decrements, etc..  A somewhat
18      more limited number of studies were then available which substantiated mortality and morbidity
19      associations with various fine particle  indicators (e.g., PM2 5, sulfate, FT, etc.); and only one, the
20      Harvard Six Cities analysis by Schwartz et al. (1996a), evaluated relative contributions of the
21      fine (PM2 5) versus the coarse (PM10_2 5) fraction of PM10, with PM25 appearing to be associated
22      more strongly with mortality effects than PM10_2 5. Lastly, only a very few studies seemed to be
23      indicative of possible coarse particle effects,  e.g., increased asthma risks associated with quite
24      high PM10 concentrations in a few locations where coarse particles strongly dominated the
25      ambient PM10 mix.
26
27      8.4.3.1  Fine- and Coarse-Particle Effects on Mortality
28           A greatly enlarged and still rapidly growing number of new studies published since the
29      1996 PM AQCD provide much new evidence further substantiating ambient PM associations
30      with increased human mortality and morbidity. As indicated in Table 8-1, most newly reported
31      analyses, with few exceptions, continue to show statistically significant associations between

        April 2002                                8-220        DRAFT-DO NOT QUOTE OR CITE

-------
 1      short-term (24-h) PM concentrations and increases in daily mortality in many U.S. and Canadian
 2      cities (as well as elsewhere). Also, the reanalyses of Harvard Six City and ACS study data
 3      substantiate the original investigator's findings of long-term PM exposure associations with
 4      increased mortality as well.
 5
 6      8.4.3.1.1 Effects on Total Mortality
 1           The effects estimates from the newly reported studies are generally consistent with those
 8      derived from the earlier 1996 PM AQCD assessment, which reported risk estimates for excess
 9      total (nonaccidental) deaths associated with short-term PM exposures as generally falling within
10      the range of ca. 1.5 to 8.5% per 50 //g/m3 PM10 (24-h) increment and ca. 2.5 to 5.5% increase per
11      25 Mg/m3 PM2 5 (24-h) increment.
12           Several new PM epidemiology studies which conducted time-series analyses in multiple
13      cities were noted to be of particular interest, in that they provide evidence of effects across
14      various geographic locations (using standardized methodologies) and more precise pooled effect
15      size estimates with narrow confidence bounds, reflecting the typically much stronger power of
16      such multi-city studies over individual-city analyses to estimate a mean effect.  Based on pooled
17      analyses across multiple cities, the percent total (non-accidental) excess deaths per 50 //g/m3
18      PM10 increment were estimated in different multi-city analyses to be: (a) 2.3% in the 90 largest
19      U.S. cities;  (b) 3.4% in 10 large U.S. cities; (c) 3.5% in the 8 largest Canadian cities; and
20      (d) 2.0% in European cities.
21           Many new individual-city studies found positive associations (most statistically significant
22      at p < 0.05) for the PM2 5 fraction, with effect size estimates typically ranging from ca. 2.0 to ca.
23      8.5% per 25 //g/m3 PM25 for U.S. and Canadian  cities.  Of the 10 or so new analyses that not
24      only evaluated PM10 effects but also made an effort to compare fine versus coarse fraction
25      contributions to total mortality, only two are multi-city analyses yielding pooled effects
26      estimates:  (a) the Klemm and Mason (2000) recomputation of Harvard Six Cities data,
27      confirming the original published findings by Schwartz et al. (1996a); and (b) the Burnett et al.
28      (2000) study of the 8 largest Canadian cities. Both of these studies found roughly comparable,
29      statistically significant excess risk estimates for PM2 5, i.e., approximately 3% increased total
30      mortality risk per 25 //g/m3 PM2 5 increment.
        April 2002                                8-221        DRAFT-DO NOT QUOTE OR CITE

-------
 1           With regard to possible coarse particle short-term exposure effects on mortality, in those
 2      new studies which evaluated PM10_2 5 effects as well as PM2 5 effects, the coarse particle (PM10_2 5)
 3      fraction was also consistently positively associated with increased total mortality, albeit the
 4      coarse fraction effect size estimates were generally less precise than those for PM2 5 and
 5      statistically significant at p < 0.05 in only a few studies (Figure 8-6). Still, the overall picture
 6      tends to suggest that excess total mortality risks may well reflect actual coarse fraction particle
 7      effects, in at least some locations. This may be most consistently the case in arid areas, e.g., in
 8      Mexico City, Santiago, Chile, or in the Phoenix area (as shown in Mar et al., 2000). On the other
 9      hand,  significant (or nearly significant) elevations in coarse PM-related total mortality risks have
10      also been detected for Steubenville, PH (an eastern U.S. urban area in the Harvard Six City
11      Study), as shown by Schwartz et al. (1996a). These results may reflect contamination of later-
12      resuspended coarse PM by metals in fine PM emitted from smelters (Phoenix) or steel mills
13      (Steubenville) that was earlier deposited on nearby soils. Excess total mortality risks associated
14      with short-term (24-h) exposures to coarse fraction particles capable of depositing in the lower
15      respiratory tract generally fall in the range of 0.5 to 6.0% per 25 //g/m3 PM10_25 increment for U.S.
16      and Canadian cities.
17           Three new papers provide particularly interesting new information on relationships between
18      short-term coarse particle exposures and total elderly mortality (age 65 and older), using
19      exposure TEOM data from the EPA ORD NERL monitoring site in Phoenix, AZ. Each used
20      quite different models but each reported statistically significant relationships between mortality
21      and coarse PM, specifically PM10_2 5, an indicator for the thoracic fraction of coarse-mode PM.
22           Smith et al. (2000), using a three-day running average as the exposure metric, performed
23      linear regression of the square root of daily mortality on the long-term trend, meteorological and
24      PM-based variables.  Two mortality variables were used, total (non-accidental) deaths for the city
25      of Phoenix and  the same for a larger, regional area. Using a linear analysis, effects based on
26      coarse PM were statistically significant for both regions, whereas effects based on fine PM
27      (PM2 5) were not.  However, when the  possibility of a nonlinear response was taken into account,
28      no evidence was found for a nonlinear effect for coarse PM, but fine PM was found to have a
29      statistically significant effect for concentration  thresholds of 20 and 25 //g/m3. There was no
30      evidence of confounding between fine and coarse PM, suggesting that fine and coarse PM  are
31      "essentially separate pollutants having distinct  effects". Smith et al.  (2000) also  observed a

        April 2002                                8-222        DRAFT-DO NOT QUOTE OR CITE

-------
 1      seasonal effect for coarse PM, the effect being statistically significant only during spring and
 2      summer.  Based on a principal component analysis of elemental concentrations, crustal elements
 3      are highest in spring and summer and anthropogenic elements lowest, but Smith et al. (2000) felt
 4      that the implication that crustal, rather than anthropogenic elements, were responsible for the PM
 5      mortality was counterintuitive.
 6           Clyde et al. (2000) used a more conventional model, a Poisson regression of log deaths on
 7      linear PM variables; but they employed Bayesian model averaging to consider a wide variety of
 8      variations in the basic model.  They considered three regions: the Phoenix metropolitan area;
 9      a small subset of zip code to give a region presumably with uniform PM25; and a still smaller zip
10      code region surrounding the monitoring site (thought to be uniform as to PM10 concentrations).
11      The models considered lags of 0, 1, 2, or 3 days but only for single day PM variables (no running
12      averages  as used by Smith et al., 2000). A PM effect with a reasonable probability was found
13      only in the uniform PM2 5 region and only for coarse PM.
14           Mar et al. (2000) used conventional Poisson regression methods and limited their analyses
15      to the smallest area (called Uniform PM10 by Clyde et al.). They reported modeling data for lag
16      days 0 to 4.  Coarse fraction PM was marginally significant on lag day 0.  No direct fine particle
17      measures were statistically significant on day 0.  A regional sulfate factor determined from
18      source apportionment, however, was statistically significant. No correlations were reported for
19      the source apportionment factors, but the correlation coefficient between sulfur (S) in PM25 (as
20      measured by XRF) with coarse fraction PM was only 0.13, suggesting separate and distinct
21      effects for regional sulfate and coarse fraction PM.
22           The above three studies of PM- total mortality relationships in Phoenix tend to suggest a
23      statistical association of coarse fraction PM with total elderly mortality in addition to and
24      different from any relationship with fine PM, fine PM components, or source factors for fine PM.
25           With regard to long-term PM exposure effects on total (non-accidental) mortality, the
26      newly available evidence from the HEI Reanalyses of Harvard Six Cities and ACS data (and
27      extensions, thereof), substantiate well associations attributable to chronic exposures to inhalable
28      thoracic particles (indexed by PM15 or PM10) and the fine fraction of such particles (indexed by
29      PM2 5 and/or sulfates).   Statistically significant excess risk for total mortality was shown by the
30      reanalyses to fall in the range of 4-18% per 20 //g/m3 PM15/10 increment and 14-28% per
31      20 //g/m3 PM2 5 increase, thus suggesting likely stronger associations with fine versus coarse

        April 2002                                8-223        DRAFT-DO NOT QUOTE OR CITE

-------
 1      fraction particles. Significant fine PM associations with total mortality were also found in the
 2      latest reported AHSMOG results for males, but not in females.
 3           Other recent studies on the relation of mortality to particle composition and source (Laden
 4      et al., 2000; Mar et al., 2000; Ozkaynak et al., 1996;  Tsai et al., 2000) suggest that particles from
 5      certain sources may have much higher potential for adverse health effects than others, as
 6      delineated by source-oriented evaluations involving factor analyses. Laden et al. (2000)
 7      conducted factor analyses of the elemental composition of PM25 for Harvard Six Cities study
 8      data for 1979-1988. In the analysis for all six cities combined, the excess risk for daily mortality
 9      was estimated to be 3.4% (CI, 1.7 to 5.2) per 10 //g/m3 increment  in a mobile source factor;  1.1%
10      (CI, 0.3 to 2.0) per 10 //g/m3 for a coal source factor, and -2.3% (CI, -5.8 to 1.2) per 10 //g/m3
11      for a crustal factor.  There was large variation among the cities and some suggestion of an
12      association with a fuel oil factor identified by V or Mn, but it was not statistically significant.
13           Mar et al.  (2000) applied factor analysis to evaluate mortality in relation to 1995-1997 fine
14      particle elemental components and gaseous pollutants (CO, NO2,  SO2) in an area of Phoenix, AZ,
15      close to the air pollution monitors.  The PM25 constituents included sulfur, Zn, Pb,  soil-corrected
16      potassium, organic and elemental carbon, and a soil component estimated from oxides of Al, Si,
17      Ca,  Fe, and It. Based on models fitted using one pollutant at a time, statistically significant
18      associations were found between total mortality and PM10, CO (lags 0 and 1), NO2 (lags 0, 1, 3,
19      4), S (negative), and soil (negative). Statistically significant associations were also found
20      between cardiovascular mortality and CO (lags 0 to 4), NO2 (lags  1 and 4), SO2 (lags 3 and 4),
21      PM25 (lags 1,3,4), PM10 (lag 0), PM10_2 5 (lag 0), and elemental, organic, or total carbon.
22      Cardiovascular  mortality was  significantly related to a vegetative burning factor (high loadings
23      on organic carbon and soil-corrected potassium), motor vehicle exhaust/resuspended road dust
24      factor (with high loadings on Mn, Fe, Zn, Pb, OC, EC, CO, and NO2), and a regional sulfate
25      factor (with a high loading on S). However, total mortality was negatively associated with a soil
26      factor (high loadings on Al, Fe, Si) and a local SO2 source factor,  but was positively associated
27      with the regional sulfate factor.
28           Tsai et al. (2000) analyzed daily time series of total and cardiorespiratory deaths, using
29      short periods of 1981-1983 data for Newark, Elizabeth, and Camden, NJ. In addition to
30      inhalable particle mass (PM15) and fine particle mass (PM2 5), the study evaluated data for metals
31      (Pb, Mn, Fe, Cd, V, Ni, Zn, Cu) and for three fractions of extractable organic matter. Factor

        April 2002                                8-224        DRAFT-DO NOT QUOTE OR CITE

-------
 1      analyses were carried out using the metals, CO, and sulfates.  The most significant sources or
 2      factors identified as predictors of daily mortality were oil burning (targets V, Ni), Zn and Cd
 3      processing, and sulfates.  Other factors (dust, motor vehicles targeted by Pb and CO, industrial
 4      Cu or Fe processing) were not significant predictors. In Newark, oil burning sources and sulfates
 5      were positive predictors, and Zn/Cd a negative predictor for total mortality. In Camden oil
 6      burning and motor vehicle emissions predicted total mortality, but copper showed a marginal
 7      negative association. Oil burning, motor vehicle emissions, and sulfates were predictors of
 8      cardiorespiratory mortality in Camden.  In Elizabeth, resuspended dust indexed by Fe and Mn
 9      showed marginal negative associations with mortality, as did industrial sources traced by Cu.
10           The set of results from the above factor analyses studies do not yet allow one to identify
11      with great certainty a clear set of specific high-risk chemical components of PM. Nevertheless,
12      some commonalities across the studies seem to highlight the likely importance of mobile source
13      and other fuel combustion emissions (and apparent lesser importance of crustal particles) as
14      contributing to increased total or cardiorespiratory mortality.
15
16      8.4.3.1.2 Effects on Cause-Specific Mortality
17      Cardiovascular- and Respiratory-Related Mortality
18           Numerous new studies have evaluated PM-related effects on cause-specific mortality.
19      Most all report positive, often statistically significant (at p < 0.05), short-term (24-h) PM
20      exposure associations with cardiovascular (CVD)- and respiratory-related deaths.  Cause-specific
21      effects estimates appear to mainly fall in the range of 3.0 to 7.0% per 25 //g/m3 24-h PM25 for
22      cardiovascular or combined cardiorespiratory mortality and 2.0 to 7.0% per 25 //g/m3 24-h PM2 5
23      for respiratory mortality in U.S. cities.  Effect size estimates for the coarse fraction (PM10_25) for
24      cause-specific mortality generally fall in the range of ca. 3.0 to 8.0% for cardiovascular and ca.
25      3.0 to 16.0% for respiratory causes per 25 //g/m3 increase in PM10_25.
26           Also of particular interest, the above noted study by Mar et al. examined the associations of
27      a variety of PM indicators with cardiovascular mortality (for age >65), again in the zip code area
28      near the Phoenix monitoring site.  For this end point, coarse PM was statistically significant on
29      lag day 0 but not on subsequent lag days. PM25 and a number of fine PM indicators were
30      statistically significant on lag day  1 but not on lag day 0. This suggests a distinct and separate
31      relationship of PM2 5 and PM10_2 5.  As in the case of total mortality, the only fine PM indicator

        April 2002                                8-225        DRAFT-DO NOT QUOTE OR CITE

-------
 1      found to be statistically significant on lag day 0 was regional sulfate. However, the low
 2      correlation coefficient between S in PM2 5 and PM10_2 5 (r = 0.13) suggests that the two
 3      relationships represent different sets of deaths. Thus, there is some evidence suggesting that the
 4      risk of cardiovascular mortality , as well as that of total mortality, may be statistically associated
 5      with PM10_2 5 and that this relationship may be independent of any relationships with fine particle
 6      indicators.
 7
 8      Long- Term PM Exposure and Lung Cancer
 9           Of particular interest with regard to PM-related effects on cause-specific mortality is a
10      growing body of evidence linking long-term PM exposure with increased risk of lung cancer.
11      Historical evidence has included studies of lung cancer trends, studies of occupational groups,
12      comparisons of urban and rural populations, and case-control and cohort studies using diverse
13      exposure metrics (Cohen and Pope, 1995).  Table 8-39 (derived from Cohen, 2000) indicates
14      that, despite possible problems with respect to potential errors in exposure and other risk factor
15      measurement errors, numerous past ecological and case-control studies of PM and lung cancer
16      have generally indicated a lung cancer RR greater than 1.0 to be associated with living in areas
17      indicated as having higher PM exposures.
18           Prospective cohort studies offer a potentially more powerful approach to evaluate the
19      apparent association between PM exposures and the development of lung cancer. The  1996 PM
20      AQCD (U.S. Environmental Protection Agency, 1996a) summarized three of these more
21      elaborate studies that carefully evaluated the effects of PM air pollution exposure on lung cancer
22      using the prospective cohort design. In the Adventist Health Smog Study (AHSMOG), Abbey
23      et al. (1991) followed a cohort of Seventh Day Adventists, whose extremely low prevalence of
24      smoking and uniform, relatively healthy dietary patterns reduce the potential for confounding by
25      these factors.  Excess lung cancer incidence was observed in females in relation to both particle
26      (TSP) and ozone exposure after 6 years follow up time.  Dockery et al.  (1993) reported the
27      results ofa  14-to  16-year prospective follow-up of 8,111 adults living in six U.S. cities that
28      evaluated associations between air pollution and mortality. After controlling for individual
29      differences  in age, sex, cigarette smoking, BMI, education, and occupational exposure, Dockery
30      et al. (1993) found an elevated but non-significant risk for lung cancer (RR = 1.37; 95%CI = 0.81
31      to 2.31) for a difference in PM2 5 pollution equal to that of the most polluted versus the least

        April 2002                                 8-226       DRAFT-DO NOT QUOTE OR CITE

-------
               TABLE 8-39. SUMMARY OF PAST ECOLOGIC AND CASE-CONTROL
               EPIDEMIOLOGIC STUDIES OF OUTDOOR AIR AND LUNG CANCER
Study Type Authors
Ecologic Henderson et al. ,
1975
Buffleretal.,
1988
Archer, 1990
Case-Control Pike et al., 1979
Vena, 1982
Jedrychowski,
etal., 1990
Katsouyanni,
etal., 1990
Barbone etal.,
1995
Nyberg et al.,
2000
Locale
Los Angeles, CA
Houston, TX
Utah
Los Angeles
Buffalo, NY
Cracow, Poland
Athens, Greece
Trieste, Italy
Stockholm,
Sweden
Exposure Classification
High PAH Areas
TSP by Census Tract
TSP by county
BAP Geo. Areas
TSP Geo. Areas
TSP and SO2
Geo. Areas
Soot Concentration
Geo. Areas
High Particle
Deposition Areas
High NO2 Areas
Rate Ratio (95% CI)
1.3@96-116ug/m3TSP
(CI: N/A)
1.9@16ug/m3TSP
(CI: N/A)
1.6@85ug/m3TSP
(CI: N/A)
1.3 @96-116ug/m3TSP
1. 7 @ 80-200 ug/m3 TSP
(CI: 1.0-2.9)
1.1@TSP> 150 ug/m3
(CI: N/A)
I.l@sootupto400
ug/m3
(CI: N/A)
1.4@>0.3g/m2/day
(CI: 1.1-1.8)
1.3
(CI: 0.9-1.9)
        Source: Cohen (2000).
 1     polluted city. Pope et al. (1995) similarly analyzed PM2 5 and sulfate (SO4) air pollution as
 2     predictors of mortality in a prospective study of 7-year survival data (1982 to 1989) for about
 3     550,000 adult volunteers obtained by the American Cancer Society (ACS).  Both the ACS and
 4     Harvard studies have been subjected to much scrutiny, including an extensive independent audit
 5     and re-analysis of the original data (Krewski et al., 2000) that confirmed the originally published
 6     results. The ACS study controlled for individual differences in age, sex, race, cigarette smoking,
 7     pipe and cigar smoking, exposure to passive cigarette smoke, occupational exposure, education,
 8     BMI, and alcohol use. Lung cancer mortality was significantly associated with particulate air
 9     pollution when SO4= was used as the index,, but not when PM2 5 mass was used as the index for a
10     smaller subset of the study population that resided in metropolitan areas where PM2 5 data were
11     available from the Inhalable Particle (IP) Network.  Thus, while these prospective cohort studies
       April 2002
8-227
DRAFT-DO NOT QUOTE OR CITE

-------
 1      have also indicated that long-term PM exposure is associated with an increased cancer risk, the
 2      effect estimates were generally not statistically significant, quite possibly due to inadequate
 3      statistical power by these studies at that time (e.g., due to inadequate population size and/or
 4      follow-up time for long-latency cancers).
 5          The AHSMOG investigators have re-examined the association between long-term PM
 6      exposure and increased risk of both lung cancer incidence and lung cancer mortality in
 7      nonsmokers using longer-term follow-up of this cohort and improved analystical approaches.
 8      Beeson et al. (1998) considered this cohort of some 6,338 nonsmoking, non-Hispanic, white
 9      Californian adults, ages 27-95, that was followed from 1977 to 1992 for newly diagnosed
10      cancers. Incident lung cancer in males was positively and significantly associated with IQR
11      increases for mean concentrations of PM10 (RR = 5.21; 95% CI =1.94-13.99). For females in the
12      cohort,  incident lung cancer was positively associated with Inter-Quartile Range (IQR) increases
13      for SO2(RR = 2.14; CI, 1.36-3.37) and IQR increases for PM10 exceedance frequencies of 50
14      ug/m3 (RR =1.21; 95% CI = 0.55-2.66) and 60 ug/m3 (RR = 1.25; 95% CI = 0.57-2.71). Thus,
15      increased risks of incident lung cancer were deemed by the authors to be associated with elevated
16      long-term ambient concentrations of PM10 and SO2 in both genders. The higher PM10 risk effect
17      estimate for cancer in  males appeared to be partially due to gender differences in long-term air
18      pollution exposures.   Abbey et al. (1999) also related long-term ambient concentrations of PM10,
19      SO4=, SO2, O3, and NO2 to 1977-1992 mortality in the AHSMOG cohort. After adjusting for a
20      wide range of potentially confounding factors, including occupational and indoor sources of air
21      pollutants, PM10 showed a strong association with lung cancer deaths in males (PM10 IQR
22      RR=2.38; 95% CI: 1.42 - 3.97). In this cohort, males spent more  time outdoors than females,
23      thus having higher estimated air pollution exposures than the cohort females. Ozone showed an
24      even stronger association with lung cancer mortality for males, and SO2 showed strong
25      associations with lung cancer mortality for both sexes. The authors reported that other pollutants
26      showed weak or no association with mortality. Therefore, increases in both lung cancer
27      incidence and lung cancer mortality in the extended follow-up analysis of the AHSMOG study
28      were found to be most consistently associated with elevated long-term ambient concentrations of
29      PM10 and SO2, especially among males.
30          A recent follow-up analysis of the major ACS study by Pope et al. (2002) responds to a
31      number of criticisms previously noted for the earlier ACS analysis (Pope et al., 1995) in the  1996

        April 2002                                8-228        DRAFT-DO NOT QUOTE OR CITE

-------
 1      PM AQCD (U.S. Environmental Protection Agency, 1996a), most notably by including
 2      examinations of other pollutants, better occupational indices, and diet information, while also
 3      addressing possible spatial auto-correlations due to regional location. The recent extension of the
 4      ACS study includes approximately 500,000 adult men and women drawn from ACS-CPS-II
 5      enrollment and follow-up during 1982-1998. This new analysis of the ACS cohort substantially
 6      expands the prior analysis, including:  (1) a more than doubling of the follow-up time to 16 years
 7      (and a more than tripling of the number of deaths in the analysis); (2) substantially expanded
 8      exposure data, including gaseous co-pollutant data and new PM2 5 data collected in 1999-2001;
 9      (3) improved control of occupational exposures; (4) incorporation of dietary variables that
10      account for total fat consumption, as well as consumption of vegetables, citrus and high-fiber
11      grains; and  (5) utilization of recent advances in statistical modeling, including the incorporation
12      of random effects and non-parametric spatial smoothing components in the Cox proportional
13      hazards model.
14           In this extended ACS analysis, it was found that long-term exposure to air pollution, and
15      especially to PM25, is associated with increased annual risk of mortality. With the longer 15-year
16      follow-up period and with improved metrics of PM25 exposures, this study for the first time
17      detected a statistically significant association between living in a city with higher PM25 and
18      increased risk of dying of lung cancer. Each 10 ug/m3 elevation in annual average fine PM was
19      associated with a 13 percent (95% CI=4%-23%) increase in lung cancer mortality. Coarse
20      particles and gaseous pollutants were generally not significantly associated with excess lung
21      cancer mortality. SO4= was significantly associated with mortality and lung cancer deaths in this
22      extended data set, yielding RR's consistent with (i.e., not significantly different from) the SO4=
23      RR's reported in the previously published 7-year follow-up (Pope et al, 1995).  However, while
24      PM2 5 was specific to the causes most biologically plausible to be influenced by air pollution in
25      this analysis (i.e., cardio-pulmonary and cancer), SO4= was significantly associated with every
26      mortality category in this new analysis, including that for "all-other causes",  This suggests that
27      the PM2 5 associations found are more biologically plausible than the less specific SO4=
28      associations found. The PM2 5 cancer risk appears greatest for non-smokers and among those
29      with lower socio-economic status (as indicated by lower educational attainment).
30           Overall, these new cohort studies confirm and strengthen the published older ecological and
31      case-control evidence indicating that living in an area that has experienced higher PM exposures

        April 2002                                8-229        DRAFT-DO NOT QUOTE OR CITE

-------
 1      can cause a significant increase in the RR of lung cancer incidence and associated mortality.
 2      In particular, the new ACS cohort analysis more clearly indicates that living in a city with higher
 3      PM25 levels is associated with an elevated risk of lung cancer amounting to an increase of some
 4      10 to 15% above the lung cancer risk in a cleaner city.
 5           With regard to specific ambient fine particle constituents that may significantly contribute
 6      to the observed ambient PM-related increases in lung cancer, PM components of diesel engine
 7      exhaust represent one class of likely important contributors. Diesel emission PM typically
 8      comprises a noticeable fraction of ambient fine particles in many urban areas, having been
 9      estimated to comprise from approximately 5 to 35% of ambient PM25 in some U.S. urban areas
10      (see Chapter 3). Also, as discussed in a separate Health Effects Assessment of Diesel Engine
11      Exhaust (U.S. Environmental Protection Agency, 2002), extensive epidemiologic and toxicologic
12      evidence links diesel emissions (including fine PM components) to increased risk of lung cancer.
13
14      8.4.3.1.3  Shortening-of-Life Associated With Long-Term Ambient Particulate
15               Matter Exposure
16           The public health burden of mortality associated with exposure to ambient PM depends not
17      only on the increased risk of death, but also on the length of life shortening that is attributable to
18      those deaths. However, the 1996 PM AQCD concluded that confident quantitive determination
19      of years of life lost to ambient PM exposure was not yet possible; life shortening may range from
20      days to years (U.S. Environmental Protection Agency, 1996a). Now, some newly available
21      analyses provide further interesting insights with regard to potential life-shortening associated
22      with chronic PM exposures.
23
24      8.4.3.1.3.1  Life-Shortening Estimates Based on Semi-Individual Cohort Study Results
25           Brunekreef (1997) reviewed the available evidence of the mortality effects of long-term
26      exposure to PM air pollution and, using life table methods, derived an estimate of the reduction
27      in life expectancy implied by those effect estimates. Based on the results of Pope et al. (1995)
28      and Dockery et al. (1993), a relative risk of 1.1 per 10 //g/m3 exposure over 15 years was
29      assumed for the effect of PM air pollution on men 25-75 years of age. A 1992 life table for men
30      in the Netherlands was developed for 10 successive five-year categories that make up the
31      25-75 year old age range.  Life expectancy of a 25 year old was then calculated for this base case
32      and compared with the calculated life expectancy for the PM-exposed case, where the  death rates
        April 2002                               8-230       DRAFT-DO NOT QUOTE OR CITE

-------
 1      were increased in each age group by a factor of 1.1. A difference of 1.11 years was found
 2      between the "exposed" and "clean air" cohorts' overall life expectancy at age 25. Looked at
 3      another way, this implies that the expectation of the lifespan for persons who actually died from
 4      air pollution was reduced by more than 10 years, since they represent a small percentage of the
 5      entire cohort population. A similar calculation by the authors for the 1969-71 life table for U.S.
 6      white males yielded an even larger reduction of  1.31 years for the entire population's life
 7      expectancy at age 25. Thus, these calculations imply that relatively small differences in long-
 8      term exposure to ambient PM can have substantial effects on life expectancy.
 9
10      8.4.3.1.3.2 Potential Effects of Infant Mortality on Life-Shortening Estimates
11           Deaths among children can logically have the greatest influence on a population's overall
12      life expectancy, but the Brunekreef (1997) life table calculations did not consider any possible
13      long-term air pollution exposure effects on the population aged <25 years.  As discussed above,
14      some of the older cross-sectional studies and the more recent studies by Bobak and Leon (1992),
15      Woodruff et al. (1997), Bobak and Leon (1999), and Loomis et al. (1999) suggest that infants
16      may be among sub-populations notably affected by long-term PM exposure.  Thus, although it is
17      difficult to quantify, any premature mortality that does occur among children due to long-term
18      PM exposure  (as suggested by these new studies) would significantly increase the overall
19      population life shortening over and above that estimated by Brunekreef (1997) for long-term PM
20      exposure of adults aged 25 years and older.
21
22      8.4.3.2 PM10, PM2 5 (Fine), and PM10 2 5 (Coarse) Particulate Matter Effects on Morbidity
23           At the time of the 1996 PM AQCD, fine particle morbidity studies were mostly limited to
24      Schwartz et al. (1994) , Neas et al. (1994, 1995); Koenig et al.  (1993); Dockery et al. (1996); and
25      Raizenne et al. (1996); and discussion of coarse  particles morbidity effects was also limited to
26      only  a few studies (Gordian et al., 1996; Hefflin et al., 1994) which implicated PM10_25 as a
27      possible important fraction of PM10.  Since the 1996 PM AQCD, several new studies have been
28      published in which newly available size-fractionated PM data allowed investigation of the effects
29      of both fine (PM25) and coarse fraction (PM10_25) particles.  Fine (FP) and coarse fraction (CP)
30      particle results are noted below for studies by morbidity outcome areas, as  follows:
        April 2002                                8-231        DRAFT-DO NOT QUOTE OR CITE

-------
 1      cardiovascular disease (CVD) hospital admissions (HA's); respiratory medical visits and hospital
 2      admissions; and respiratory symptoms and pulmonary function changes.
 3           As discussed in Section 8.3.1 (on cardiovascular effects associated with acute ambient PM
 4      exposure), an extensive new body of evidence has emerged since the 1996 PM AQCD that
 5      evaluates PM10 effects on cardiovascular-related hospital admissions and visits. Especially
 6      notable new evidence has been provided by several new multi-city studies (Schwartz, 1999;
 7      Samet et al., 2000a,b) that yield pooled estimates of PM-CVD effects across numerous U.S.
 8      cities and regions. These studies found not only significant PM associations, but also
 9      associations with other gaseous pollutants as well, thus  hinting at likely independent effects of
10      certain gases (O3, CO, NO2, SO2) and/or interactive effects with PM. These and other individual-
11      city studies generally appear to confirm likely excess risk of CVD-related hospital admission for
12      U.S. cities in the range of 3-10% per 50 //g/m3 PM10, especially among the elderly (> 65 yr).
13           In addition to the PM10 studies, several new U.S. and Canadian studies evaluated fine-mode
14      PM effects on cardiovascular outcomes. Moolgavkar (2000a) reported PM2 5 to be significantly
15      associated with CVD HA for lag 0 and 1 in Los Angeles.  Burnett et al. (1997a) reported that fine
16      particles were significantly associated with  CVD HA in a single pollutant model, but not when
17      gases were included in multipollutant models for the 8 largest Canadian city data. Stieb et al.
18      (2000) reported both PM10 and PM25 to be associated with CVD emergency department (ED)
19      visits in single pollutant, but not multipollutant models.  Similarly, Morgan et al. (1998) reported
20      that PM2 5 measured by nepholonetry was associated with CVD HA for all ages and 65+ yr, but
21      not in the multipollutant model.  Tolbert et al. (2000a) reported that coarse particles were
22      significantly associated with dysrhythmias, whereas PM2 5 was not.  Other studies (e.g., Liao
23      et al., 1999; Creason et al., 2001; Pope et al., 1999b,c) reported associations between increases in
24      PM25 and several measures of decreased heart rate variability, but Gold et al. (2000) reported a
25      negative association of PM25 with heart rate and decreased variability in r-MSSD (one heart rate
26      variability measure). A recent study by Peters and colleagues (2001) reported significant
27      temporal  associations between acute (2-h or 24-h) measures of PM25 and myocardial infarction.
28      Overall, these new studies collectively appear to implicate fine particles, as well as possibly some
29      gaseous co-pollutants, in cardiovascular morbidity, but the relative contributions of fine particles
30      acting alone or in combination with gases such as O3, CO, NO2 or SO2 remain to be more clearly
31      delineated and quantified. The most difficult issue relates to interpretation of reduced PM effect

        April 2002                                8-232       DRAFT-DO NOT QUOTE OR CITE

-------
 1      size and /or statistical significance when co-pollutants derived from the same source(s) as PM are
 2      included in multipollutant models.
 3           Section 8.3.1 also discussed U.S. and Canadian studies that present analyses of coarse
 4      fraction particles (CP) relationships to CVD outcomes. Lippmann et al. (2000) found significant
 5      positive associations of PM10_2 5 with ischemic heart disease hospital admissions in Detroit
 6      (RR =1.10, CI 1.026,  1.18).  Tolbert et al. (2000a) reported significant positive associations of
 7      heart dysrhythmias with CP (p = 0.04) as well as for elemental carbon  (p = 0.004), but these
 8      preliminary results must be interpreted with caution until more complete analyses are carried out
 9      and reported. Burnett et al. (1997b) noted that CP was the most robust of the particle metrics
10      examined to inclusion of gaseous covariates for cardiovascular hospitalization, but concluded
11      that particle mass and chemistry could not be identified as an independent risk factor for
12      exacerbation of cardiorespiratory disease in this study. Based on another Canadian study,
13      Burnett et al. (1999), reported statistically significant associations for CP in univariate models
14      but not in multipollutant models; but the use of estimated rather than measured PM exposures
15      indices limits the interpretation of the PM results reported.
16           The collective evidence reviewed  above, in general, appears to suggest excess risks for
17      CVD-related hospital admissions of approximately 4.0 to 10% per 25 //g/m3 PM2 5 or PM10_2 5
18      increment.
19           Section 8.3.2 also discussed new studies of effects of short-term PM exposure on the
20      incidence of respiratory hospital admissions and medical visits.  Several new U.S. and Canadian
21      studies have yielded particularly interesting results suggestive of roles  of both fine and coarse
22      particles in respiratory-related hospital admissions.  In an analysis of Detroit data, Lippmann
23      et al. (2000) found comparable effect size estimates for PM2 5 and PM10_2 5.  That is, the excess
24      risk for pneumonia hospital admissions (in no co-pollutant model) was 13% (CI 3.7, 22) per
25      25 Mg/m3 PM25 and 12% (CI 0.8, 24) per 25 //g/m3 PM10_25. Because PM25 and PM10_25 were not
26      highly correlated, the observed association between coarse particles and health outcomes were
27      possibly not confounded by smaller particles. Despite the  greater measurement error associated
28      with PM10_25 than with either PM2 5 and  PM10, this indicator of the coarse particles within the
29      thoracic fraction was associated with some of the outcome measures. The interesting result is
30      that PM10_2 5 appeared to be a separate factor from other PM metrics, especially given the  effect
31      estimates of PM10_25 with  pneumonia hospital admissions (lag 1; RR= 1.11, 95% CI:  1.006,

        April 2002                                 8-233        DRAFT-DO NOT QUOTE OR CITE

-------
 1      1.233). Burnett et al. (1997b) also reported PM (PM10, PM2 5, and PM10.2 5) associations with
 2      respiratory hospital admissions, even with O3 in the model. Notably, the PM10_25 association was
 3      significant (RR =1.13 for 25 //g/m3; CI = 1.05 - 1.20); and inclusion of ozone still yielded a
 4      significant coarse mass RR =1.11  (CI = 1.04 - 1.19). Moolgavkar et al. (2000) showed the most
 5      consistent association for PM10 across lags (0-4d), while PM2 5 yielded the strongest positive PM
 6      metric association at lag 3 days.  Also, Moolgavkar (2000a) reported that, in Los Angeles, both
 7      PM10 and PM2 5 yielded both positive and negative associations at different lags for single
 8      pollutant models but not in two pollutant models.  Delfmo et al. (1997) reported that both PM2 5
 9      and PM10 are positively associated with ED visits for respiratory disease. Morgan et al. (1998)
10      reported that PM25 estimated from nephelometry yielded a PM25 association with COPD hospital
11      admissions for 1-hr max PM that was more positive than 24-h average PM2 5.
12           Some new studies appear to substantiate PM associations with asthma-related hospital
13      admissions. For example, Norris et al (1999) reported associations of emergency department
14      visits for asthma in children with both PM25 and PM10_25.  Two other studies presented uniquely
15      different analyses of hospital admissions in the Seattle, Washington area. Sheppard et al.  (1999)
16      studied relationships between PM metrics that included PM10_2 5 and non-elderly adult hospital
17      admissions for asthma in the greater Seattle area and reported significant relative rates for PM10,
18      PM25 and PM10.25 (lagged 1 day). For PM10.25, the relative risk was 1.04 (95% CI 1.01, 1.07).
19      In a different analysis, Lumley and Heagerty (1999) examined PMX and PM104 in the King
20      County, WA (Seattle) area during the same time period but for hospital admissions for overall
21      respiratory disease. Since only a significant hospital admission association was found with PMl 0
22      and not PM1(M, a dominant role by sub-micron particles in PM2 5 - asthma HA association was
23      suggested, but this may not be an appropriate conclusion based on  several differences between
24      the study analysis methods and differences between asthma versus respiratory outcome measures
25      used in the two Seattle studies. For a 16% decrease in PM10 levels, Friedman et al.  (2001)
26      reported decreased hospital admissions for asthmatics during the Olympics in Atlanta.
27           Several other studies (Chen et al. 2000; Choudhury et al., 1997; Moolgavlar 2000a;
28      Lippsett et al.,  1997) report results for areas (e.g., Reno-Sparks, NV; Anchorage, AK; Phoenix,
29      AZ; Santa Clara, CA) where coarse fraction particles tend to constitute a large fraction of PM10
30      but no measures of PM10_2 5 were available.  These studies showing significant PM10 effects on
31      respiratory hospital admissions provide additional data suggestive of likely coarse fraction

        April 2002                                8-234        DRAFT-DO NOT QUOTE OR CITE

-------
 1      particle effects on respiratory morbidity.  It is possible that vegetative burning (e.g., wood) in
 2      these western cities may produce coarse particles whose toxicity may differ from that of coarse
 3      crustal fraction particles.
 4           Thus, although PM10 mass has most often been implicated as the PM pollution index
 5      affecting respiratory hospital admissions, the overall collection of new studies reviewed in
 6      Section 8.3.2 appear to suggest relative roles for both fine and coarse PM mass fractions, such as
 7      PM2 5  and PM10.2 5.
 8           Section 8.3.3 assessed relationships between PM exposure on lung function and respiratory
 9      symptoms.  While most data examine PM10 effects, several studies also examined fine and coarse
10      fraction particle effects. Schwartz and Neas (2000) report that cough was the only  response in
11      which coarse fraction particles appeared to provide an independent contribution to  explaining the
12      increased incidence. The  correlation between CM and PM25 was moderate (0.41).  Coarse
13      fraction particles had little association with evening  peak flow. Tiittanen et al. (1999) also
14      reported a significant effect of PM10_25 for cough.  Thus, cough may be an appropriate outcome
15      related to coarse fraction particle effects. However,  the limited data base suggests that further
16      study is appropriate. The  report by Zhang, et al. (2000) of an association between coarse fraction
17      particles and the indicator "runny nose" is noted also.
18           Published epidemiological studies have  collectively indicated that exposure to PM air
19      pollution can be associated with adverse human health effects, and that asthmatics  represent a
20      population that can be especially affected by acute exposures to air pollution (e.g.,  see Koren and
21      Utell,  1997). In particular, prospective epidemiologic studies of panels of individuals confirm
22      the air pollution-asthma exacerbation association.
23           For respiratory symptoms and PFT changes, several new asthma studies report relationships
24      with ambient PM measures. The peak flow analyses results for asthmatics tend to  show small
25      decrements for both PM10 and PM2 5.  Several studies included PM2 5 and PM10 independently in
26      their analyses of peak flow.  Of these, Naeher et al. (1999), Tiittanen et al. (1999), Pekkanen et
27      al. (1997), and Romieu et al. (1996) all found comparable results for PM2 5 and PM10. The study
28      of Peters et al. (1997c) found slightly larger effects for PM25. The study of Schwartz and Neas
29      (2000) found larger effects for PM2 5 than for coarse fraction particles. Three studies included
30      both PM10 and PM25 in their analyses of respiratory  symptoms. The studies of Peters et al.
31      (1997c) and Tiittanen et al. (1999) found similar effects for the two PM measures.  Only the

        April 2002                                 8-23 5        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Romieu et al. (1996) study found slightly larger effects for PM25. While the PM associations
 2      with adverse health effects among asthmatics and others are well documented, the type/source(s)
 3      of those particles most associated with adverse health effects among asthmatics are not known at
 4      this time. Indeed, the makeup of PM varies greatly from place to place and over time, depending
 5      upon factors such as the sources that contribute to the pollution and the prevailing atmospheric
 6      conditions, affecting particle formation, coagulation, transformation, and transport. One
 7      suspected causal PM agent is the fine particle component of diesel combustion exhaust.
 8           Two studies (Delfino et al., 1998; Ostro et al., 2001) examined PM effects on asthmatics
 9      using one hour maximum exposure measures by TEOM, and both studies indicate a relationship
10      with measures of respiratory symptoms. Further research is needed at these shorter exposure
11      times for different PM size fractions.
12           For non-asthmatics, several studies evaluated PM2 5 effects. Naeher et al. (1999) reported
13      similar AM PEF decrements for both PM2 5 and PM10. Neas et al. (1996) reported a
14      nonsignificant negative association for PEF and PM2 b and Neas et al. (1999) also reported
15      negative but nonsignificant PEF results. Schwartz and Neas (2000) reported a significantly PM
16      PEF association with PM2 5, and Tiittanen et al. (1999) also reported negative but nonsignificant
17      association for PEF andPM25. Gold et al. (1999) reported significantly PEF results.  Schwartz
18      and Neas (2000) reported significant PM2 5 effects relative to lower respiratory symptoms.
19      Tiittanen et al. (1999) showed significant effects for cough and PM25 for a 4-day average.
20           Another study conducted by Peters et al. (1997c) in Erfurt, Germany in 1992 is unique for
21      two reasons: (1) they studied the size distribution in the range 0.01 to 2.5 //m and (2) examined
22      the number of particles.  They report that the health effects of 5 day means of the number count
23      (NC) for ultrafme particles were larger than those related to the mass of the fine particles. For
24      NC 0.01 -0.1, cough was significant for the same day and the five day mean.
25           In a chronic respiratory disease study of 22-24 North American communities evaluated in
26      the  1996 PM AQCD, Raizenne et al. (1996) found PM2A to be related to a statistically significant
27      FVC deficit of -3.21% (-4.98, -1.41).  Dockery et al. (1996) also reported PM2A associations
28      with increased bronchitis; odds ratio =1.50 (95% CI = 0.91, 2.47).
29           The above new studies offer much more information than was available in 1996. Effects
30      were noted for several morbidity endpoints: cardiovascular hospital admissions, respiratory
31      hospital admissions and cough.  Still insufficient data exists from these relatively limited studies

        April 2002                               8-236       DRAFT-DO NOT QUOTE OR CITE

-------
 1      to allow strong conclusions at this time as to which size-related ambient PM components may be
 2      most strongly related to one or another morbidity endpoints.  Very preliminarily, however, fine
 3      particles appear to be more strongly implicated in cardiovascular outcomes than are coarse
 4      fraction particles, whereas both seem to impact respiratory endpoints.
 5
 6      8.4.4  The Question of Lags
 7           The effect of selecting lags on the  resulting model for PM health effects is one of the main
 8      issues in model  selection.  Using simulated data with parameters similar to a Seattle PM10_25 data
 9      series, Lumley and Sheppard (2000) showed that the bias resulting from the selection is shown to
10      be similar in size to the relative risk estimates from the measured data.  More precisely, the log
11      relative risk from the measured Seattle data is about twice the mean bias in the simulated control
12      data, and the published estimate of relative risk is only at the 90th percentile of the bias
13      distribution in these control analysis. The selection rule used was to choose the lag (between
14      0 and 6 day) with the largest estimated relative risk. In comparisons to real data from Seattle for
15      other years  and from Portland, OR (with similar weather patterns to Seattle), similar bias issues
16      became evident.
17           In most of the past air pollution health effects time-series studies, after the basic model (the
18      best model with weather and seasonal cycles as covariates) was developed, several pollution lags
19      (usually 0 to 3 or 4 days) were individually introduced and the most significant lag(s) chosen for
20      the RR calculation. While this practice may bias the chance  of finding a significant association,
21      without a firm biological reason to establish a fixed pre-determined lag, it appears reasonable.
22      Due to likely individual variability in response to air pollution, the apparent lags of effects
23      observed for aggregated population counts are expected to be "distributed" (i.e., symmetric or
24      skewed bell-shape).  The "most significant lag" in such distributed lags is also expected to
25      fluctuate statistically.  The "vote-counting" of the most significant lags reported in the past
26      PM-mortality studies shows that 0 and 1 day lags are,  in that order, the most frequently reported
27      "optimal" lags, but such estimates may be biased because these lags are also likely the most
28      frequently examined ones. Thus, a more systematic approach across different data sets was
29      needed to investigate this issue.
30           The Samet et al.  (2000b) analysis of the 90 largest U.S. cities provides particularly useful
31      information on this matter. Figure 8-28 depicts the Samet et al. (2000b) overall pooled results,

        April 2002                                 8-237       DRAFT-DO NOT QUOTE OR CITE

-------
                            I
                           -0.2
 I
0.0
 I
0.2
 I
0.4
                                                            0.6
 I
0.8
                               % Change in Mortality per 10 jjg/m3 Increase in PM
 i
1.0
                                                                        10
       Figure 8-28.  Marginal posterior distribution for effects of PM10 on all cause mortality at
                     lag 0,1, and 2 for the 90 cities. From Samet et al. (2000a,b).  The numbers in
                     the upper right legend are posterior probabilities that overall effects are
                     greater than 0.
       Source:  Samet et al. (2000b).
 1      showing the posterior distribution of PM10 effects for the 90 cities for lag 0, 1, and 2 days. It can
 2      be seen that the effect size estimate for lag Iday is about twice that for lag 0 or lag 2 days,
 3      although their distributions overlap. However, a careful examination of Figures 6 and 7 in the
 4      NMMAPS I Report suggests that the maximum PM10 effect may occur in different cities with
 5      somewhat different lag relationships. In terms of the magnitude of the estimated PM10 effects,
 6      Table 8-40, based on NMMAPS I Figure 7 (posterior bivariate distribution for each county; PM10
 7      effect adjusted for O3), suggests that somewhat different patterns may apply in different
 8      locations. These data suggest that while lag 1 effects are typically the largest, there may be some
 9      situations in which lag 0 or lag 2 effects are larger.
10           The NMMAPS mortality and morbidity analyses and another HEI-sponsored study on PM
11      components (Lippmann et al., 2000) illustrate three different ways to deal with temporal
12      structure: (1) assume all sites have the same lag, e.g., 1 day, for a given effect; (2) use the lag or
       April 2002
              8-238
                  DRAFT-DO NOT QUOTE OR CITE

-------
         TABLE 8-40. COMPARISON
           ANALYSES FOR 0,1, AND
OF PM10 EFFECT SIZES ESTIMATED BY NMMAPS
2 DAY LAGS FOR THE 20 LARGEST U.S. CITIES
        County
     Ordered PM,n effect sizes
                                                      10
       Los Angeles
       New York
       Chicago
       Dallas/Fort Worth
       Houston
       San Diego
       Santa Ana /Anaheim
       Phoenix
       Detroit
       Miami
       Philadelphia
       Seattle
       San Jose
       Cleveland
       San Bernardino
       Pittsburgh
       Oakland
       San Antonio
       Riverside
     Lag 0 < lag 1 « lag 2
     Lag 0 = lag 1 » lag 2
     Unreadable
     Lag 0> lag 1, lag Klag2
     Lag 0< lag 1, lag 1 > lag 2
     Lag 0 = lag 1 > lag 2
     Lag 0 > lag 1 > lag 2
     Lag 0 = lag 1< lag 2
     Lag 0 < lag 1, lag 1 > lag 2
     Lag 0 < lag 1 = lag 2
     Lag 0< lag 1, lag 1 > lag 2
     Lag 0< lag 1, lag 1 > lag 2
     Lag 0 > lag 1 = lag 2
     Lag 0> lag 1, lag Klag2
     Lag 0 > lag 1 = lag 2
     Lag 0< lag 1, lag 1 > lag 2
     Lag 0 < lag 1 = lag 2
     Lag 0 = lag 1< lag 2
     Lag 0< lag 1, lag 1 > lag 2
1     moving average giving the largest or most significant effect and for each pollutant and endpoint;
2     and (3) use a flexible distributed lag model, with parameters adjusted to each site
3          The NMMAPS mortality analyses used the first approach.  This approach introduces a
4     consistent response model across all locations. However, since the cardiovascular, respiratory, or
5     other causes of acute mortality usually associated with PM are not at all specific, there is little
6     a priori reason to believe that they must have the same relation to current or previous PM
7     exposures at different sites.  The imposed consistency in lag that maximizes the aggregate effect
      April 2002
         8-239
DRAFT-DO NOT QUOTE OR CITE

-------
 1      of lag 1 across all cities, in Figure 15-18 and 24 of NMMAPS II, may obscure important regional
 2      or local differences for lags other than 1 day.  Moolgavkar (2000a,b) illustrates this point for
 3      three large U.S. cities where strong PM effects on cardiovascular mortality occur at lags 4-5 and
 4      1-2 days in Maricopa County, lag 3 in Cook County, and lag 0 in Los Angeles County.  These
 5      may correspond to the onset or exacerbation of different illnesses leading to cardiovascular
 6      mortality.
 7           The NMMAPS morbidity studies evaluate 0- and 1-day lags, the moving average  of 0 and
 8      1-day lags, polynomial distributed lag models, and unrestricted distributed lag models.  The
 9      first-stage models for each city in the study were fitted for each city, with no restriction as to a
10      consistent model across all cities, and combined across all 14 cities in the  second stage  as shown
11      in Table 14 and Figure 23 of NMMAPS H. A comparison of the data tabulated in the NMMAPS
12      Report Appendices shows large differences across cities in the apparent magnitude of the PM10
13      effect, depending on how the PM concentration data over the preceding few days are used.
14           The approach used in Lippmann et al. (2000) and many other studies is to use the model
15      that maximizes some global model goodness-of-fit criterion. This leads to selection of different
16      models at different sites, as might be expected. However, the best-fitting model (for lags, for
17      example) is often the model with the largest or most significant PM10 coefficient.  All models for
18      the pollutant(s) of interest are usually compared among themselves only after a preliminary
19      baseline model has been fitted. The baseline model takes into account most of the other
20      variables with which PM10 could be plausibly associated,  so that the remaining variation in
21      morbidity or mortality that can be explained by including PM10 indicators with different temporal
22      structures is nearly "orthogonal" or independent of the baseline model.  The restriction  to the
23      same lag day at all sites certainly increases the precision of that estimate, but possibly at the cost
24      of obscuring different relationships between time of exposure and health effect at other sites.
25           An additional complication in assessing the shape of a distributed lag is that the apparent
26      spread of the distributed lag may depend on the pattern of persistence of air pollution (i.e.,
27      episodes may persist for a few days), which may vary from city to city and from pollutant to
28      pollutant. If this is the case, fixing the lag across cities or across pollutants may not be  ideal, and
29      may tend to obscure important nuances of lag structures that may provide important clues to
30      possible different lags between PM exposures and different cause-specific effects.
        April 2002                                8-240       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Thus, it is possible that the extent of lag and its spread may vary depending on the cause of
 2      death. For example, Rossi et al. (1999) report that, in their analysis of TSP-cause specific
 3      mortality in Milan, Italy, the lags varied for different cause of death (i.e., same day for respiratory
 4      infections and heart failure; 3-4 days for myocardial infarction and COPD). Thus, the lag for
 5      total mortality may exhibit mixed lags (weighted by the frequency of deaths in each cause).
 6      Another example was reported for a recent Mexico City study (Borja-Aburto et al., 1998), in
 7      which they found significant PM2 5-total mortality associations for same day and 4-day lag, but
 8      not for the intervening 2 to 3 days (percent increases per 25 //g/m3 were 3.38, -4.00, 1.03,
 9      1.08,3.43, 2.49, for 0 through 5 day lags, respectively).  The authors state:  "This phenomenon is
10      consistent with both a harvesting of highly susceptible persons on the day of exposure to high
11      pollution levels and a lagged increase in mortality due to delayed effects of reduction of
12      pulmonary defenses, cardiovascular complications, or other homeostatic changes among
13      less-compromised individuals".  It is interesting to note that Wichmann et al. (2000) also
14      reported that the most predictive single day effects on mortality for mass concentrations of
15      0.01-2.5 (j, particles were either immediate (0-1 d lag) or delayed (4-5 d lag) for their data from
16      Erfurt, Germany.
17           It should also be noted that if one chooses the most significant single lag day only, and if
18      more than one lag day shows positive (significant or otherwise) associations with mortality, then
19      reporting a RR for only one lag would also underestimate the pollution effects. Schwartz
20      (2000b) investigated this issue, using the 10 U.S. cities data where daily PM10 values were
21      available for 1986-1993. Daily total  (non-accidental) deaths of persons 65 years  of age and older
22      were analyzed.  For each city, a GAM Poisson model adjusting for temperature, dewpoint,
23      barometric pressure, day-of-week, season, and time was fitted.  Effects of distributed lag were
24      examined using four models:  1-day mean at lag 0 day; 2-day mean at lag 0 and 1  day; second-
25      degree distributed lag model using lags 0 through 5 days; unconstrained distributed lag model
26      using lags 0 through 5 days. The inverse variance weighted averages of the ten cities'  estimates
27      were used to combine results.  The results indicated that the effect size estimates for the
28      quadratic distributed model and unconstrained distributed lag model were similar. Both
29      distributed lag models resulted in substantially larger effect size estimates (7.25% and 6.62%,
30      respectively, as  percent excess total death per 50 //g/m3 increase in PM10) than the single day lag
31      (3.29%) and moderately larger effect size estimates than the two-day average models (5.36%).

        April 2002                                8-241        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Samet et al. (2000a,b) also applied 7- and 14-day unconstrained distributed lag models to
 2      Chicago, Minneapolis/St. Paul, and Pittsburgh data, and reported that the sum of the 7-day
 3      distributed lag coefficients was greater than the estimates based on a single day's value, but the
 4      14-day estimate was substantially lower than the 7-day estimate in Chicago and Minneapolis/
 5      St. Paul. Thus, it is possible that the usual RR estimate using one lag day may underestimate PM
 6      effects.
 7           Mis-specification of the lag structure may cause important modeling biases. Most of the
 8      published literature for the U.S. evaluates only single-day models, a choice dictated by the every-
 9      sixth-day sampling schedule used for PM10 in many U.S. cities. When this occurs, it is not
10      possible to evaluate multi-day models with greater biological plausibility, such as moving
11      average models and distributed lag models. Only three of the 20 largest U.S. cities used in the
12      NMMAPS mortality study (Chicago, Minneapolis-St. Paul, Pittsburgh) had daily data (Samet
13      et al., 2000a,b,c).  The 14 cities used in the NMMAPS hospital admissions study had daily PM10
14      data, but some of these cities were too small to be included among the 90 largest cities in the
15      mortality study (Canton and Youngstown,  OH, Boulder and Colorado Springs, CO). An every-
16      other-day sampling schedule was used in the Harvard Six City Study, for which the PM data on a
17      given day has been used as though it were  a two-day moving, alternately concurrent with
18      mortality on half the days and lagging mortality by one day on the other days.  While the most
19      commonly used lags in PM time series models are zero or one day, some studies have found PM
20      effects with longer lags (Loomis et al., 1999, in Mexico City; Ponka et al., 1999, for Helsinki),
21      and other studies have found effects at both short and long time lags in  some cities (Moolgavkar,
22      2000a,b). It is therefore plausible that mortality or hospital admissions from PM may arise from
23      different responses or PM-associated diseases with different characteristic lags, for example, that
24      cardiovascular responses may arise almost immediately after exposure,  within zero or one days
25      or even within two hours (Peter et al., 2002, for myocardial infarction).
26           One would then expect to see different best-fitting lags for different cause-specific
27      mortality or hospital admissions.  This idea was fully demonstrated in Lippmann et al. (2000)
28      where different single-day lag models for different health endpoints, PM metrics, and gaseous
29      pollutants were included in the model.  The best-fitting PM models had lag 0 to 3 days,
30      depending on the endpoint.  This problem is not solved by use of distributed lag models.
31      Schwartz and Zanobetti (2002) found it necessary to use different distributed lag models in each

        April 2002                                 8-242        DRAFT-DO NOT QUOTE OR CITE

-------
 1      of the 10 cities whose concentration-response functions were combined by meta-smoothing. One
 2      wonders if the meta-smoothing results by region (Dominici et al., 2002) might have been
 3      changed if the concentration-response function and optimal lag for each city had been used.
 4      In this case, model mis-specification may involve a combination of two potential biases.
 5
 6      8.4.5  New Assessments of Mortality Displacement
 7           There have been a few studies that investigated the question of "harvesting", a phenomenon
 8      in which a deficit in mortality occurs following days with (pollution-caused) elevated mortality,
 9      due to depletion of the susceptible population pool.  This issue is very important in interpreting
10      the public health implication of the reported short-term PM mortality effects.  The 1996 PM
11      AQCD discussed suggestive evidence observed by Spix et al. (1993) during a period when air
12      pollution levels were relatively high.  Recent studies, however, generally typically used data from
13      areas with lower, non-episodic pollution levels.
14           Schwartz (2000c) separated time-series air pollution, weather, and mortality data from
15      Boston, MA, into three components:  (1) seasonal and longer fluctuations; (2) "intermediate"
16      fluctuations; (3) "short-term" fluctuations.  By varying the cut-off between the intermediate and
17      short term, evidence of harvesting was sought. The idea is, for example, if the extent of
18      harvesting were a matter of a few days, associations between weekly average values of mortality
19      and air pollution (controlling for seasonal cycles) would not be seen. For COPD, Schwartz
20      (2000c) reported evidence indicating that most of the mortality was only displaced by a few
21      weeks; for pneumonia, heart attacks, and all cause mortality, the effect size  increased as longer
22      time scales were included. The percent increase in deaths associated with a 25 //g/m3 increase in
23      PM25 increased from 5.3% (95%CI: 6.8, 9.0) to 9.64% (95%CI: 8.2, 11.1).
24           Schwartz and Zanobetti (2000) used the same approach described above to analyze a larger
25      data set from Chicago, IL for 1988-1993. Total (non-accidental), in-hospital, out-of-hospital
26      deaths, as well as heart disease, COPD, and pneumonia elderly hospital admissions were
27      analyzed to investigate possible PM10 "harvesting" effects. GAM Poisson models adjusting for
28      temperature, relative humidity, day-of-week, and season were applied in baseline models using
29      the average of the same day and previous day's PM10.  Seasonal and trend decomposition
30      techniques called STL were applied to the health outcome and exposure data to decompose them
31      into different time-scales (i.e., short-term to long-term), excluding long seasonal cycles (120 day

        April 2002                                8-243        DRAFT-DO NOT QUOTE OR CITE

-------
 1      window).  The associations were examined with smoothing windows of 15, 30, 45, and 60 days.
 2      The effect size estimate for deaths outside hospital was larger than for deaths inside hospital.
 3      All cause mortality showed an increase in effect size at longer time scales. The effect size for
 4      deaths outside hospital increases more steeply with increasing time scale than that for inside
 5      hospital deaths.
 6           Zanobetti et al. (2000b) used GAM distributed lag models to help quantify mortality
 7      displacement in Milan, Italy, 1980-1989. Non-accidental total deaths were regressed on smooth
 8      functions of TSP distributed over the same day and the previous 45 days using penalized splines
 9      for the smooth terms and seasonal cycles, temperature, humidity, day-of-week, holidays, and
10      influenza epidemics.  The mortality displacement was modeled as the initial positive  increase,
11      negative rebound (due to depletion), followed by another positive coefficients period, and the
12      sum of the three phases were considered as the total cumulative effect.  TSP was positively
13      associated with mortality up to 13 days, followed by nearly zero coefficients between 14 and
14      20 days, and then followed by smaller but positive coefficients up to the 45th day (maximum
15      examined). The sum of these coefficients was over three times larger than that for the single-day
16      estimate.
17           Zeger et al. (1999) first illustrated, through simulation, the implication of harvesting for PM
18      regression coefficients (i.e., mortality relative risk) as observed in frequency domain.  Three
19      levels of harvesting, 3 days, 30 days, and 300 days were simulated.  As expected, the  shorter the
20      harvesting, the larger the PM coefficient in the higher frequency range. However, in  the real data
21      from Philadelphia, the regression coefficients increased toward the lower frequency range,
22      suggesting that the extent of harvesting, if it exists, is not in the short-term range.  Zeger
23      suggested that "harvesting-resistant" regression coefficients could be obtained by excluding the
24      coefficients in the very high frequency range (to eliminate short-term harvesting) and in the very
25      low frequency range (to eliminate seasonal confounding).  Since the  observed frequency domain
26      coefficients in the very high frequency range were smaller than those in the mid frequency range,
27      eliminating the "short-term harvesting" effects would only increase the average of those
28      coefficients in the rest of the frequency range.
29           Frequency domain analyses are rarely performed in air pollution health effects studies,
30      except perhaps the spectral analysis (variance decomposition by frequency) to identify seasonal
31      cycles.  Examinations of the correlation by frequency (coherence) and the regression  coefficients

        April 2002                                8-244       DRAFT-DO NOT QUOTE OR CITE

-------
 1      by frequency (gain) may be useful in evaluating the potentially frequency-dependent
 2      relationships among multiple time series. A few past examples in air pollution health effects
 3      studies include: (1) Shumway et al.'s (1983) analysis of London mortality analysis, in which
 4      they observed that significant coherence occurred beyond two week periodicity (they interpreted
 5      this as "pollution has to persist to affect mortality"); (2) Shumway et al.'s (1988) analysis of
 6      Los Angeles mortality data, in which they also found larger coherence in the lower frequency;
 7      (3) Ito's (1990) analysis of London mortality data in which he observed relatively constant gain
 8      (regression coefficient) for pollutants across the frequency range, except the annual cycle. These
 9      results also suggest that associations and effect size, at least, are not concentrated in the very high
10      frequency range.
11           Schwartz (2000c), Zanobetti et al. (2000b), and Zeger et al.'s (1999) results all suggest that
12      the extent of harvesting, if any, is not a matter of only a few days. Other past studies that used
13      frequency domain analyses are also  at least qualitatively in agreement with the evidence against
14      the short-term only harvesting.  Since very long wave cycles (> 6 months) need to be controlled
15      in time-series analyses to avoid seasonal confounding, the extent of harvesting beyond 6 months
16      periodicity is not possible in time-series study design.  While these studies suggest that observed
17      short-term associations are not simply due to short-term harvesting, more data are needed to
18      obtain quantitative estimates of the extent of prematurity of deaths.
19
20      8.4.6  Concentration-Response Relationships for Ambient PM
21           In the 1996 PM AQCD, the limitations of identifying 'threshold' in the concentration-
22      response relationships in observational studies were discussed including the low data density in
23      the lower PM concentration range, the small number of quantile indicators often used, and the
24      possible influence of measurement error.  Also,  a threshold for a population, as opposed to a
25      threshold for an individual, has some conceptual issues that need to be noted.  For example,
26      Schwartz (1999) discussed that, since individual thresholds would vary from person to person
27      due to individual differences in genetic level susceptibility and pre-existing disease conditions, it
28      would be almost mathematically impossible for a threshold to exist in the population. This
29      argument holds only if the most sensitive members of a population are sensitive to very low
30      concentrations, which may not be the case.  The person-to-person difference in the relationship
31      between personal exposure and the concentration observed at a monitor would also add to the

        April 2002                                8-245        DRAFT-DO NOT QUOTE OR CITE

-------
 1      variability.  Because one cannot directly measure but can only compute or estimate a population
 2      threshold, it would be difficult to interpret an observed threshold, if any, biologically.  Despite
 3      these issues, several studies have attempted to address the question of threshold by analyzing
 4      large databases, or by conducting simulations.
 5           Daniels et al. (2000) examined the presence of threshold using the largest 20 U.S. cities for
 6      1987-1994. The authors compared three log-linear GAM regression models:  (1) using a linear
 7      PM10 term;  (2) using a cubic spline of PM10 with knots at 30 and 60 //g/m3 (corresponding
 8      approximately to 25 and 75 percentile of the distribution); and, (3) using a threshold model with
 9      a grid search in the range between 5 and 200 //g/m3 with 5 //g/m3 increment. The covariates
10      included in these models are similar to those used by the same research group previously (Kelsall
11      et al.,  1997; Samet et al., 2000a,b), including the smoothing function of time, temperature and
12      dewpoint, and  day-of-week indicators.  Total, cardiorespiratory, and other mortality series were
13      analyzed. These models were fit for each city separately, and for model (1) and (2), the
14      combined estimates across cities were obtained by using inverse variance weighting if there was
15      no heterogeneity across cities, or by using a two-level hierarchical model if there was
16      heterogeneity.  The best fit among the models, within each city and over all cities, were also
17      determined using the Akaike's Information Criterion (AIC). The results using the spline model
18      showed that, for total and cardiorespiratory mortality, the spline curves were roughly linear,
19      consistent with the lack of a threshold.  For mortality from other causes, however, the curve did
20      not increase until PM10 concentrations exceeded 50 //g/m3. While the test of heterogeneity
21      indicated that there was considerable heterogeneity in these curves across cities (see Figure
22      8-29), the shapes of the curves were similar across cities, with no indication of one city unduly
23      influencing the overall estimate of the curves. The hypothesis of linearity was examined by
24      comparing the  AIC values across models.  The results suggested that the linear model was
25      preferred over  the spline and the threshold models. Thus, these results suggest that linear models
26      without a threshold may well be appropriate for estimating the effects of PM10 on the types of
27      mortality of main interest.
28           Thus, while these studies do not refute the usual assumption of a linear no-threshold
29      concentration-response function, neither do they provide unqualified support for that assumption.
30      Sensitivity analyses for individual city studies' concentration-response function would be
31      helpful. Schwartz and Zanobetti (2000) investigated the presence of threshold by simulation and

        April 2002                                8-246       DRAFT-DO NOT QUOTE OR CITE

-------
                                                      Total
                          o.os
                     -c
                     o
                     
-------
 1      span (0.7) was used in each of the cities.  The predicted values of the log relative risks were
 2      computed for 2 //g/m3 increments between 5.5 //g/m3 and 69.5 //g/m3 of PM10 levels.  Then, the
 3      predicted values were combined across cities using inverse-variance weighting. The simulation
 4      results indicated that the "meta-smoothing" approach did not bias the underlying relationships for
 5      the linear and threshold models, but did result in a slight downward bias for the logarithmic
 6      model. Measurement error (additive or multiplicative) in the simulations did not cause upward
 7      bias in the relationship below threshold. The threshold detection in the simulation was not very
 8      sensitive to the choice of span in smoothing. In the analysis of real data from 10 cities, the
 9      combined curve did not show evidence of a threshold in the PM10-mortality associations.
10           Cakmak et al. (1999) investigated methods to detect and estimate threshold levels in time
11      series studies. Based on the realistic range of error observed from  actual Toronto pollution data
12      (average site-to-site correlation: 0.90 for O3; 0.76 for COH; 0.69 for TSP; 0.59 for SO2; 0.58 for
13      NO2; and 0.44 for CO), pollution levels were generated with multiplicative error for six levels of
14      exposure error (1.0, 0.9, 0.8, 0.72, 0.6, 0.4, site-to-site correlation). Mortality series were
15      generated with three PM10 threshold levels (12.8 Mg/m3, 24.6 Mg/m3, and 34.4 Mg/m3).  LOESS
16      with a 60% span was used to observe the exposure-response curves for these 18 combinations of
17      exposure-response relationships with error.  A parameter threshold model was also fit using non-
18      linear least  squares.  Graphical presentations indicate that LOESS adequately detects threshold
19      under no error, but the thresholds were "smoothed out" under the extreme error scenario.  Use of
20      a parametric threshold model was adequate to give "nearly unbiased" estimates of threshold
21      concentrations even under the conditions of extreme measurement error, but the uncertainty in
22      the threshold estimates increased with the degree of error.  They concluded, "if threshold exists,
23      it is highly likely that standard statistical analysis can detect it".
24           The Smith et al. (2000) study of associations between daily total mortality and PM25 and
25      PM10_2 5 in Phoenix, AZ (during 1995-1997) also investigated the possibility of a threshold.
26      In the linear model, the authors found that mortality was significantly associated with PM10_2 5,
27      but not with PM2 5. In modeling possible thresholds, they applied:  (1) a piecewise linear model
28      in which several possible thresholds were specified; and (2) a B-spline (spline  with cubic
29      polynomials) model with 4 knots. Using the piecewise model, there was no indication that there
30      was a threshold for PM10_2 5. However, for PM2 5, the piecewise model resulted in suggestive
31      evidence for a threshold, around 20 to 25 //g/m3. The B-spline results also showed no evidence

        April 2002                                 8-248       DRAFT-DO NOT QUOTE OR CITE

-------
 1      of threshold for PM10_2 5, but for PM2 5, a non-linear curve showed a change in the slope around
 2      20 //g/m3. A further Bayesian analysis for threshold selection suggested a clear peak in the
 3      posterior density of PM2 5 effects around 22 //g/m3. These results, if they in fact reflect reality,
 4      make it difficult to evaluate the relative roles of different PM components (in this case, PM2 5
 5      versus PM10_2 5).  However, the concentration-response curve for PM2 5 presented in this
 6      publication suggests more of a U- or V-shaped relationship than the usual "hockey stick"
 7      relationship. Such a relationship is, unlike the temperature-mortality relationship, difficult to
 8      interpret biologically.  Because the sample size of this data (~3 years) is relatively small, further
 9      investigation of this issue using similar methods but a larger data set is warranted. Other studies
10      evaluate non-linear relationships using a multi-city meta-smoothing approach based on non- or
11      semi-parametric smoothers rather than on linear parametric models.
12           Many ad hoc decisions go into model selection in air pollution health effects studies. The
13      effect of some of these decisions on relative risk estimates for Birmingham, AL, PM10 data,
14      previously analyzed by Schwartz (1993) and others, is illustrated by Smith et al. (2000).  The
15      response variable is non-accidental mortality.  Specifically, the selection of meteorological
16      variables, the selection of an exposure variable (as a weighted average of lagged PM values), and
17      the possibility  of nonlinear effects,  such as threshold effects, are investigated. The results are
18      sensitive to the inclusion of humidity in addition to temperature. This inclusion decreases the
19      resulting PM10 coefficient. The model is highly sensitive to the definition of an exposure
20      measure.  For example, when lags 0-4 were averaged, there was no significant effect. In an
21      attempt to account for a nonlinear PM-mortality effect, there appeared to be little effect of
22      exposure below 80 //g/m3, and a threshold analysis (as well as a generalized additive models
23      approach) supported the conclusion that the main effect is at higher values of PM. Although this
24      paper was based on an intensive analysis of a single data  set (in contrast to other studies, such as
25      NMMAPS analysis, which combined data form many cities), it demonstrated the very wide range
26      of interpretations that are possible using alternative, but statistically valid, analyses of the same
27      data.
28
29
30
        April 2002                                8-249        DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.4.7  New Assessments of Consequences of Measurement Error
 2      8.4.7.1 Theoretical Framework for Assessment of Measurement Error
 3           Since the 1996 PM AQCD, there have been some advances in conceptual framework
 4      development to investigate the effects of measurement error on PM health effects estimated in
 5      time-series studies. Several new studies evaluated the extent of bias caused by measurement
 6      errors under a number of scenarios with varying extent of error variance and covariance structure
 7      between co-pollutants.
 8           Zidek et al. (1996) investigated, through simulation, the joint effects of multi-collinearity
 9      and measurement error in Poisson regression model, with two covariates with varying extent of
10      relative errors and correlation. Their error model was of classical error form (W=X+U, where W
11      and X are surrogate and true measurements, respectively, and the error U is normally distributed).
12      The results illustrated the transfer of effects from the "causal" variable to the confounder.
13      However, for the confounder to have larger coefficients than the true predictor, the correlation
14      between the two covariates had to be large (r = 0.9), with moderate error (a > 0.5) for the true
15      predictor, and no error for the confounder in their scenarios. The transfer-of-causality effect was
16      mitigated when the confounder also became subject to error. Another interesting finding that
17      Zidek et al. reported is the behavior of the standard errors of these coefficients: when the
18      correlation between the covariates was high (r = 0.9) and both covariates had no error, the
19      standard errors for both coefficients were inflated by factor of 2; however, this phenomenon
20      disappeared when the confounder had error.  Thus, multi-collinearity influences the significance
21      of the coefficient of the causal variable only when the confounder is accurately measured.
22           Zeger et al. (2000) also conducted a mathematical analysis of PM mortality effects in
23      ordinary least square model (OLS) with the classical error model, under varying extent of error
24      variance and correlation between two predictor variables. The error described here was
25      analytical error (e.g., discrepancy between the co-located monitors). In general, they found that
26      positive regression coefficients are only attenuated, but null predictors (zero coefficient) or weak
27      predictors are only able to appear stronger than true positive predictors under unusual conditions:
28      (1) true predictors must have very large positive or negative correlation (i.e., |r| > 0.9);
29      (2) measurement error must be substantial (i.e., error variance ~ signal variance); and
30      (3) measurement errors must have a large negative correlation. They concluded that estimated


        April 2002                               8-250        DRAFT-DO NOT QUOTE OR CITE

-------
 1      FP health effects are likely underestimated, although the magnitude of bias due to the analytical
 2      measurement error is not very large.
 3           Zeger et al. (1999) illustrated the implication of the classical error model and the Berkson
 4      error model (i.e., X = W + U) in the context of time-series study design. Their simulation of the
 5      classical error model with two predictors, with various combinations of error variance and
 6      correlation between the predictors/error terms, showed results similar to those reported by Zidek
 7      et al. (1996). Most notably, for the transfer of the effects of one variable to the other (i.e., error-
 8      induced confounding) to be  large, the two predictors or their errors need to be substantially
 9      correlated.  Also, for the spurious association of a null predictor to be more significant than the
10      true predictor, their measurement errors have to be extremely negatively correlated—a condition
11      not yet demonstrated as occurring in actual air pollution data sets.
12           Zeger et al. also laid out a comprehensive framework for evaluating the effects of exposure
13      measurement error on estimates of air pollution mortality relative risks in time-series  studies.
14      The error, the difference between personal exposure and the central  station's measurement of
15      ambient concentration was decomposed into three components:  (1) the error due to having
16      aggregate rather than individual exposure; (2) the difference between the average personal
17      exposure and the true ambient concentration level; and, (3) the difference between the true and
18      measured ambient concentration level. By aggregating individual risks to  obtain expected
19      number of deaths, they showed that the first component of error (the aggregate rather  than
20      individual) is a Berkson error, and, therefore is not a significant contributor to bias in the
21      estimated risk. The second error component is a  classical error and  can introduce bias if there are
22      short-term associations between indoor source contributions and ambient concentration levels.
23      Recent analysis, however, both using experimental data (Mage et al., 1999; Wilson et al., 2000)
24      and theoretical interpretations and models (Ott et al., 2000) indicate that there is no relationship
25      between the ambient concentration and the nonambient components of personal exposure to PM.
26      However, a bias can arise due to the difference between the personal exposure to ambient PM
27      (indoors plus outdoors) and  the ambient concentration.  The third error component is the
28      difference between the true and the measured ambient concentration. According to Zeger et al.
29      the final term is largely of the Berkson type if the average of the available  monitors is an
30      unbiased estimate of the true spatially averaged ambient level.
        April 2002                                 8-251        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Using this framework, Zeger et al. (2000) then used PTEAM Riverside, CA data to
 2      estimate the second error component and its influence on estimated risks.  The correlation
 3      coefficient between the error (the average population PM10 total exposure minus the ambient
 4      PM10 concentration) and the ambient PM10 concentration was estimated to be -0.63. Since this
                                 ^,
 5      correlation is negative, the fiz (the  estimated value of the pollution-mortality relative risk in the
 6      regression of mortality onzp the daily ambient concentration) will tend to underestimate the
                   yv
 7      coefficient fix that would be obtained in the regression of mortality on xt, the daily average total
 8      personal exposure, in a single-pollutant analysis. Zeger et al. (2000) then proceed to assess the
 9      size of the bias that will result from this exposure misclassification, using daily ambient
10      concentration, zt.  As shown in Equation 9, the daily average total personal exposure, xt, can be
11      separated into a variable component, Ql zp dependent on the daily ambient concentration, zp and
12      a constant component, 60, independent of the ambient concentration.
13
14                                      xt =  OQ + Qlzt + £t                                 (8-5)
15      where £t is  an error term.
16           If the nonambient component of the total personal exposure is independent of the ambient
17      concentration, as appears to be the case, Equation 9 from Zeger et al. (2000) becomes the
18      regression analysis equation familiar to exposure analysts (Dockery and Spengler, 1981; Ott
19      et al., 2000; Wilson et al., 2000).  In this case, 60 gives the average nonambient component of the
20      total personal exposure and Ql gives the ratio of the ambient component of personal exposure to
21      the ambient concentration.  (The ambient  component of personal exposure includes exposure to
22      ambient  PM while outdoors and, while indoors, exposure to ambient PM that has infiltrated
23      indoors.) In this well-known approach to  adjust for exposure measurement error, called
                                                                                   ~    ~  yv
24      regression calibration (Carroll et al., 1995), the estimate of f)x has the simple form ftx = ftz/9l.
25      Thus, for the regression calibration, the value of (3X (based on the total personal exposure) does
26      not depend on the total personal exposure but is given by [)z, based on the ambient concentration,
27      times 0l3 the ratio of the ambient component of personal exposure to the ambient concentration.
28      A regression analysis of the PTEAM data gave an estimate Ql = 0.60.
29           Zeger et al. (2000) use Equation 9, with 9o = 59.95 and 6j = 0.60, estimated from the
30      PTEAM data, to simulate values of daily average personal exposure, x*t, from the ambient

        April 2002                                8-252        DRAFT-DO NOT QUOTE OR CITE

-------
 1      concentrations, zt, for PM10 in Riverside, CA, 1987-1994. They then compare the mean of the
                  /\
 2      simulated jGf. s, obtained by the series of log-linear regressions of mortality on the simulated x*t,
                                                                               /^
 3      with the normal approximation of the likelihood function for the coefficient Pz  from the
                                                                   yv.   yv.
 4      log-linear regression of mortality directly on zt.  The resulting /3Z / fix = 0.59, is very close to
                                                                                   yv
 5      0j = 0.60. Dominici et al. (2000) provide a more complete analysis of the bias in f5z as an

 6      estimate of fix using the PTEAM Study and four other data sets and a more complete statistical
                                                            ^,             ^,
 7      model. Their findings were qualitatively similar in that fix was close to j3z /Oj. Thus, it appears

 8      that the bias is very close to Ql which depends not on the total personal exposure but only on the

 9      ratio of the ambient component of personal exposure to the ambient concentration.

10           Zeger et al. (2000), in the analyses described above, also suggested that the error due to the

11      difference between the average personal exposure and the ambient level (the second error type

12      described above) is likely the largest source of bias in estimated relative risk. This suggestion at

13      least partly comes from the comparison of PTEAM data and site-to-site correlation (the third type

14      of error described above) for PM10 and O3 in 8 US cities. While PM10 and O3 both showed

15      relatively high site-to-site correlation («0.6-0.9), a similar extent of site-to-site correlation for

16      other pollutants is not necessarily expected. Ito et al. (1998) estimated site-to-site correlations

17      (after adjusted for seasonal cycles) for PM10, O3, SO2, NO2, CO, temperature, dewpoint

18      temperature, and relative humidity, using multiple stations' data from seven central and eastern

19      states (IL, IN, MI, OH, PA, WV, WI), and found that, in a geographic scale of less 100 miles,

20      these variables could be categorized into three groups in terms of the extent of correlation:

21      weather variables (r > 0.9); O3, PM10, NO2 (r: 0.6 - 0.8); CO  and SO2 (r < 0.5).  These results

22      suggest that the contribution from the third component of error, as described in Zeger et al.

23      (2000), would vary among pollution  and weather variables.  Furthermore, the contribution from

24      the second component of error would also vary among pollutants; i.e., the ratio of ambient

25      exposure to ambient concentration, called the attenuation coefficient, is expected to be different

26      for each pollutant. Some of the ongoing studies are expected to shed some light on this issue.

27      However, more information is needed on attenuation coefficients for a variety of pollutants.

28           With regard to the PM exposure,  longitudinal studies (Wallace, 2000; Mage et al., 1999),

29      show reasonably good correlation (r = 0.6 to 0.9) between ambient PM concentrations and

30      average population PM exposure, lending support for the use of ambient  data as a surrogate for

31      personal exposure to ambient PM in time-series mortality or morbidity studies.  Furthermore,

        April 2002                                8-253       DRAFT-DO  NOT QUOTE OR CITE

-------
 1      fine particles are expected to show even better site-to-site correlation than PM10. Wilson and Suh
 2      (1997) examined site-to-site correlation of PM10, PM25, and PM10_25 in Philadelphia and
 3      St. Louis, and found that site-to-site correlations were high (r ~ 0.9) for PM2 5 but low for PM10_2 5
 4      (r ~ 0.4), indicating that fine particles have smaller errors in representing community-wide
 5      exposures.  This finding supports Lipfert and Wyzga's (1997) speculation that the stronger
 6      mortality associations for fine particles than coarse particles found in the Schwartz et al. (1996a)
 7      study may be due in part to larger measurement error for coarse particles.
 8           However, as Lipfert and Wyzga (1997) suggested, the issue is not whether the fine particle
 9      association  with mortality is a "false positive", but rather, whether the weaker mortality
10      association  with coarse particles is a "false negative".  Carrothers and Evans (2000) also
11      investigated the joint effects of correlation and relative error, but they specifically addressed the
12      issue of fine (FP) vs. coarse particle (CP) effect, by assuming three levels of relative toxicity of
13      fine versus  coarse particles (Ppp / PCP = 1, 3, and 10) and, then, evaluating the bias, (B = |E[PF]/
14      E[PC,]} / (PF / PC), as a function of FP-CP correlation and relative error associated with FP and
15      CP. Their results indicate: (1) if the FP and CP have the same toxicity, there is no bias (i.e.,
16      B=l) as long as FP and CP are measured with equal precision, but, if, for example, FP is
17      measured more precisely than CP, then FP will appear to be more toxic than CP (i.e., B > 1);
18      (2) when FP is more toxic than CP (i.e., Ppp / PCP = 3 and  10), however, the equal precision of FP
19      and CP results in downward bias of FP (B < 1), implying a relative overestimation of the less
20      toxic CP. That is, to achieve non-bias, FP must be measured more precisely than CP, even more
21      so as the correlation between FP and CP increases. They also applied this model to real data
22      from the Harvard Six Cities Study, in particular, the data from Boston and Knoxville.  Estimation
23      of spatial variability for Boston was based on external data and a range of spatial variability for
24      Knoxville (since there was no spatial data available for this city).  For Boston, where the
25      estimated FP-CP correlation was low (r = 0.28), estimated error was smaller for FP than for CP
26      (0.85 vs. 0.65,  as correlation between true vs. error-added series), and the observed FP to CP
27      coefficient ratio was high (11), the calculated FP to CP coefficient ratio was even larger (26)-thus
28      providing evidence against the hypothesis that  FP is absorbing some of the  coefficient of CP.
29      For Knoxville, where FP-CP correlation was moderate (0.54), the error for FP was  smaller than
30      for CP (0.9  vs. 0.75), and the observed FP to CP coefficient ratio was 1.4, the calculated true FP
31      to CP coefficient ratio was smaller (0.9) than the observed value, indicating that the coefficient

        April 2002                                8-254       DRAFT-DO NOT QUOTE OR CITE

-------
 1      was overestimated for the better-measured FP, while the coefficient was underestimated for the
 2      worse-measured CP. Since the amount (and the direction) of bias depended on several variables
 3      (i.e., correlation between FP and CP; the relative error for FP and CP; and, the underlying true
 4      ratio of the FP toxicity to CP toxicity), the authors concluded "...for instance, it is inadequate to
 5      state that differences in measurement error among fine and coarse particles will lead to false
 6      negative findings for coarse particles".
 7           Fung and Krewski (1999) conducted a simulation study of measurement error adjustment
 8      methods for Poisson models, using scenarios similar to those used in the simulation studies that
 9      investigated implication of joint effects of correlated covariates with measurement error. The
10      measurement error adjustment methods employed were the Regression Calibration (RCAL)
11      method (Carroll et al.,  1995) and the Simulation Extrapolation (SEVIEX) method (Cook and
12      Stefanski, 1994).  Briefly, RCAL algorithm  consists of:  (1) estimation of the regression of X on
13      W (observed version of X, with error) and Z (covariate without error); (2) replacement of X by
14      its estimate from (1), and conducting the standard analysis (i.e., regression);  and (3) adjustment
15      of the resulting standard error of coefficient to account for the calibration modeling. SIMEX
16      algorithm consists of:  (1) addition of successively larger amount of error to  the original  data;
17      (2) obtaining naive regression coefficients for each of the error added data sets; and, (3) back
18      extrapolation of the obtained coefficients to the error-free case using a quadratic or other
19      function. Fung and Krewski examined the cases for: (1) Px = 0.25; Pz = 0.25; (2) Px = 0.0;
20      pz = 0.25; (3) Px = 0.25; Pz = O.O., all with varying level of correlation (-0.8  to 0.8) with and
21      without classical additive error, and also considering Berkson type error. The behaviors of naive
22      estimates were essentially similar to other simulation studies. In most cases with the classical
23      error, RCAL performed better than SIMEX  (which performed comparably when X-Z correlation
24      was small), recovering underlying coefficients.  In the presence of Berkson type error,  however,
25      even RCAL did not recover the underlying coefficients when X-Z correlation was large ( > 0.5).
26      This is the first study to examine the performance of available error adjustment methods that can
27      be applied to time-series Poisson regression. The authors recommend RCAL over SEVIEX.
28      Possible reasons why RCAL performed better than SEVIEX in these scenarios were not discussed,
29      nor are they clear from the information given in the publication.  There has not been a study to
30      apply these error adjustment methods in real time-series health effects studies. These
31      methodologies require  either replicate measurements or some knowledge on the nature of error

        April 2002                                8-255       DRAFT-DO NOT QUOTE OR CITE

-------
 1      (i.e., distributional properties, correlation, etc.).  Since the information regarding the nature of
 2      error is still being collected at this time, it may take some time before applications of these
 3      methods become practical.
 4           Another issue that measurement error may affect is the detection of threshold in time-series
 5      studies.  Lipfert and Wyzga (1996) suggested that measurement error may obscure the true shape
 6      of the exposure-response curve, and that such error could make the exposure-response curve to
 7      appear linear even when a threshold may exist. However, based on a simulation with realistic
 8      range  of exposure error (due to site-to-site correlation), Cakmak  et al. (1999) illustrated that the
 9      modern smoothing approach, LOESS, can adequately detect threshold levels (12.8 //g/m3,
10      24.6 //g/m3, and 34.4 //g/m3) even with the presence of exposure error (see also Section 8.4.6
11      above).
12           Other issues related to exposure error that have not been investigated include potential
13      differential error among subpopulations. If the exposure errors are different between susceptible
14      population groups (e.g., people with COPD) and the rest of the population, the estimation of bias
15      may need to take such differences into account. Also, the exposure errors may vary from season
16      to season,  due to seasonal differences in the use of indoor emission sources and air exchange
17      rates due to air conditioning and heating. This may possibly explain reported season-specific
18      effects of PM and other pollutants. Such season-specific contributions of errors from indoor and
19      outdoor sources are also expected to be different from pollutant to pollutant.
20           In summary, the studies that examined joint effects of correlation and error suggest that PM
21      effects are likely underestimated, and that spurious PM effects (i.e., qualitative bias such as
22      change in the sign of coefficient) due to transferring of effects from other covariates require
23      extreme conditions and are, therefore, unlikely. Also, one simulation study suggests that, under
24      the likely range of error for PM, it is unlikely that a threshold is ignored by common smoothing
25      methods. More data are needed to examine the exposure errors for other pollutants,  since their
26      relative error contributions will influence their relative significance in relative risk estimates.
27
28      8.4.7.2  Spatial Measurement Error Issues That May Affect the Interpretation of
29             Multi-Pollutant Models with Gaseous Co-Pollutants
30           The measurement error framework put forth in Dominici et al. (2000) and  Zeger et al.
31      (2000) explicitly assumes that  one of the error components has a Berkson error structure.
32      As summarized in (Zeger et al., 2000, p. 421): "This Berkson model is appropriate when z
        April 2002                               8-256       DRAFT-DO NOT QUOTE OR CITE

-------
 1      represents a measurable factor [e.g. measured PM or another pollutant] that is shared by a group
 2      of participants whose individual [true] exposures x might vary because of time-activity patterns.
 3      For example, z might be the spatially averaged ambient level of a pollutant without major indoor
 4      sources and x might be the personal exposures that, when averaged across people, match the
 5      ambient level."  This assumption is likely accurate for sulfates, less so for fine particles and for
 6      PM10, and almost certainly incorrect for gases such as CO and NO2 that may vary substantially on
 7      an intra-urban spatial scale with widely distributed local sources.
 8           The usual  characterization of longitudinal or temporal pollutant correlation may not
 9      adequately characterize the spatial variation that is the more important aspect of association in
10      evaluating possible Berkson errors. Temporal correlation coefficients, even across large
11      distances (e.g. Ito et al., 2000) may be a consequence of large-scale weather patterns affecting the
12      concentrations of many pollutants. Local concentrations for some pollutants with strong local
13      sources and low regional dispersion (especially for CO and NO2, and PM10_2 5 to a lesser extent)
14      may have somewhat smaller temporal correlations and much greater  relative spatial variations
15      than PM. Thus, individuals in a large metropolitan area may have roughly similar levels of PM
16      exposure x on any given day for which the ambient average PM concentration z is an adequate
17      surrogate, whatever their space-time activity patterns, residence, or non-residential micro-
18      environments, while the same individuals may be exposed to systematically higher or lower
19      concentrations of a co-pollutant than the spatial average of the co-pollutant.  This violates the
20      basic assumption of the Berkson error model that within each stratum of the measured (spatially
21      averaged) level  z, the average value of the true concentration x is equal to z, i.e.,
22
23                                           E{ x  z } = z,                                 (8-6)
24
25      where E{.} is the average or expected value over the population.
26           There are  empirical reasons to believe that if the strata are chosen to be locations within a
27      metropolitan area, some individuals far from local sources have consistently less exposure than
28      the average ambient concentration (denoted p) for co-pollutants with local sources such as CO
29      and NO2, and PM25, whose true exposure (denoted q) depends on the location of the person's
30      residence or other micro-environment where most exposure occurs.  For this group,
31

        April 2002                                8-257       DRAFT-DO NOT QUOTE OR CITE

-------
 1                                           E{q  p}p.                                 (8-8)
 7
 8           There is a substantial and growing body of evidence that adverse health effects are
 9      associated with proximity to a major road or highway (Wijst et al., 1994; Monn et al., 2000;
10      Roemer and Hoek, 2001). As shown below, there is good reason to believe that the intra-city
11      variation even in PM25 is substantial within some U.S. cities. If we assume for the sake of
12      argument that concentrations of PM10  or PM2 5 are relatively uniformly distributed, then
13      associations of adverse health effects with proximity to a source cannot be attributed to a
14      pollutant such as PM with a uniform spatial distribution. NO2 is a pollutant often used to
15      illustrate the spatial non-uniformity of the gaseous co-pollutants. Figure 8- from Monn et al.
16      (1997) compares the concentrations of NO2 and PM10 as a function of curbside distance in a
17      moderately busy urban street in Zurich.  The PM10 concentrations decrease only slightly with
18      increasing distance, with the decrease more likely due to decreasing coarse particle levels than to
19      decreasing fine particle concentrations.  The NO2 concentrations show a much  stronger seasonal
20      dependence, decreasing rapidly with increasing distance in the summer and showing little
21      decrease with distance in the winter. However, the belief that PM2 5 is spatially uniform should
22      also not be accepted uncritically, as recent analyses for 27 U.S. cities shown in Chapter 3 and
23      Appendix  3 A of this document demonstrate.
24           The 90th Percentile differences (P90) between a pair of sites may provide a useful guide to
25      the differences between monitor pairs (and by implication, personal exposure to fine particles)
26      that might be reasonably expected within a metropolitan area. Shown below in Table 8-41 are the
27      maximum, median, and minimum differences between monitor pairs, the monitor pairs at which
28      the largest 90th percentile difference occurs (by reference to the tables in Appendix 3 A). Based
29      on these differences, we have shown in Table 8-42 a characterization of cities as "relatively
30      homogeneous" with P90 < 10 //g/m3 and "relatively heterogeneous" if P90 > 10 //g/m3. The
31      results in Appendix 3 A and Table 8-42 show a variety of spatial patterns of association of PM2 5
        April 2002                               8-258       DRAFT-DO NOT  QUOTE OR CITE

-------
    TABLE 8-41. MAXIMUM, MEDIAN, AND MINIMUM 90th PERCENTILE OF
       ABSOLUTE VALUES OF DIFFERENCES BETWEEN FINE PARTICLE
 CONCENTRATIONS AT PAIRS OF MONITORING SITES IN 27 METROPOLITAN
         AREAS IN ORDER OF DECREASING MAXIMUM DIFFERENCE
                     (based on Chapter 3 and Appendix 3A).
City
Los Angeles

Pittsburgh
Riverside-
San Bernardino
Birmingham
Seattle

Gary
Cleveland
Atlanta

Detroit
Salt Lake C.
St. Louis
San Diego
Louisville
Chicago
Washington DC

Steubenville
Boise
Kansas City
Philadelphia
Portland OR
Grand Rapids
Dallas
Milwaukee
Columbia
Tampa
Norfolk
Baton Rouge
N Sites
5
4 (w/o E) *
4
5
5
5
4 (w/o A) *
4
7
7
6 (w/o G) *
5
6
4
4
4
11
6
5 (w/o F)
5
4
6
5
4
4
7
8
4
4
5
3
Maximum
31.0
20.2
21.3
20.2
15.4
15.3
8.4
14.9
14.9
14.0
10.6
13.3
12.6
12.5
11.9
11.2
10.5
10.1
7.7
9.9
8.9
7.4
6.9
6.5
6.1
5.6
5.5
5.3
5.0
4.7
3.2
Pair
CE
AD
BD
BC
AE
AE
CE
BD
BG
EG
CF
CD
AC
AD
CD
AD
EK
AF
AD
AE
BD
CF
BC
AB
BC
AE
FH
AB
BD
AC
AC
Median
13.8
13.7
10.8
12.6
10.1
8.2
7.6
8.2
7.1
9.4
8.3
8.6
7.6
9.5
10.6
8.7
6.2
7.4
6.25
8.45
5.2
4.1
5.2
4.45
4.8
3.3
3.65
3.95
4.45
3.55
2.9
Minimum
11.8
11.8
4.1
7.0
7.5
3.8
3.8
5.9
3.8
6.5
6.5
4.9
3.9
6.0
7.4
6.3
3.5
4.2
4.2
2.5**
3.8
1.9
3.3
4.0
2.8
2.0
2.9
2.7
3.6
2.6
2.5
 * Without one site > 100 km from the others.
 ** Collocated monitors at sites D and E.
April 2002
8-259
DRAFT-DO NOT QUOTE OR CITE

-------
            TABLE 8-42. SUMMARY OF WITHIN-CITY HETEROGENEITY BY REGION
                                Relative Heterogeneity Among Pairs of Monitors
                    Relatively Heterogenous
                                                        Relatively Homogeneous
                               West
                               Los Angeles, CA
                               Riverside, CA
                               Salt Lake City, UT
                               San Diego, CA
East
Atlanta, GA
Birmingham, AL
Chicago, IL
Cleveland, OH
Detroit, MI
Gary, IN
Louisville, KY
Pittsburgh, PA
St. Louis, MO
         Washington, DC (with F)   Seattle, WA (with A)
East                   West
Baton Rouge, LA        Boise, ID
Columbia, SC           Portland, OR
Dallas, TX
Grand Rapids, MI
Kansas City, KS-MO
Milwaukee, WI
Norfolk, VA
Philadelphia, PA
Steubenville, OH
Tampa, FL
Washington, DC (w/o F)   Seattle, WA (w/o A)
 1     within a Metroplitan Statistical Area (MSA). There may be some discernable regional
 2     differences; but, because many major population centers are not represented in Appendix 3A,
 3     further investigation is likely warranted.
 4           The results shown here provide clear evidence that fine particle concentrations may be less
 5     homogenous in at least some MSAs than has been previously assumed.  This provides support
 6     for earlier studies using TSP and PM10 cited below.  As noted in Chapter 3, these differences may
 7     not be strictly related to the distance between monitors, especially where topography plays a role.
 8     In many eastern sites, however, particle distribution  may be more substantially governed by
 9     regional particle concentrations than by local concentrations.
10           A number of recent studies have examined the role of spatial siting of monitors on the
11     estimation of PM effects. Ito et al. (1995) examined the ability of single-site vs. multi-site
12     averages to best estimate total mortality vs.  PM10 in Cook County (Chicago), IL and Los Angeles
13     County, CA. In order to have  a sufficiently large sample size to detect effects, Ito et al. used six
14     PM10 sites in Cook County (Chicago), IL and four sites in Los Angeles County, CA.
        April 2002
                                         8-260
         DRAFT-DO NOT QUOTE OR CITE

-------
 1      A sinusoidal model was used to account for temporal components, although spline or LOESS
 2      methods would now be used. Only one Cook County site had every-day PM samples, and the
 3      others as well as the Los Angeles sites had a one-in-six-day sampling schedule. The monitor
 4      sites were located in urban and suburban settings, according to the State's objectives.  Three of
 5      the Los Angeles sites were residential and one was commercial use.  One of the Cook County
 6      sites was residential, two were commercial, and three were industrial.  One of the Chicago sites
 7      was intended to monitor population exposure, three to monitor maximum concentrations, and
 8      two to monitor both maximum concentrations and personal exposure.  There was considerable
 9      variation among the distribution of PM10 in Cook County (Chicago), IL sites, and among
10      Los Angeles County, CA sites, especially at the upper end of the distribution. The sites were
11      temporally correlated, 0.83 to 0.63 in Cook County, 0.9 to 0.7 in Los Angeles (except for one site
12      pair), across distances of 4 to 26 miles.  The Cook County mortality  estimates were better
13      estimated by some single-site estimates (Site 2 with everyday data, N = 1251) than by an average
14      using all available data with missing values estimated from non-missing data (N = 1357). The
15      every-six-day subsamples from Site 1 (N = 281) and Site 2 (lag 0, N = 246) were better
16      predictors, and from Site 4 (N = 243) and Site 6 (N = 292) about as good predictors of mortality
17      as the corresponding every-six-day averages (N = 351). In Los Angeles, only Site 4 (N = 349)
18      was about as predictive as the spatial  averages (N = 405).
19           Lipfert et al. (2000) examined the relationship between the area in which mortality occurred
20      among residents and the locations of monitoring sites or averages over monitoring sites for
21      several particle size components and particle metrics.  The mortality data were located for
22      Philadelphia, PA, for three aditional suburban Philadelphia counties, for Camden, NJ and other
23      New Jersey counties in the Philadelphia - Camden MSA. A single site was used for fine and
24      coarse particles from the Harvard School of Public Health monitors. Additional PA and NJ
25      thoracic particle data were available for 2 to 4 stations and results averaged for at least two
26      stations reporting data. The authors conclude that mortality in  any part of the region may be
27      associated with air pollution concentrations or average concentrations in any other part of the
28      region, whether particles or gases. The authors suggest two interpretations:  (a) the associations
29      of mortality with pollution were random (from  carrying out multiple significance tests) and not
30      causal, or (b) both particles and gaseous pollutants have a broad regional distribution. The
31      authors note that interpretation (b) may lead to large uncertainties in identifying which pollutant

        April 2002                                8-261        DRAFT-DO NOT QUOTE OR CITE

-------
 1      exposures for the population are primarily responsible for the observed effects. These data could
 2      be studied further to evaluate smaller-scale spatial relationships among health effects and gases.
 3           Lippmann et al. (2000) evaluated the effects of monitor siting choice using 14 TSP
 4      monitoring stations in Detroit, MI, and nearby Windsor, ON, Canada. The stations operated
 5      from 1981-1987  with almost complete data. When a standard log-linear link Poisson regression
 6      model for mortality was fitted to TSP data for each of the 14 sites, the relative risk estimates
 7      were similar for within-site increments of 5th to 95th percentiles, generally highest and positive at
 8      lag day 1, but not statistically significant except for site "w" (site 12,  south of the urban center of
 9      Wayne County) and nearly significant at sites "f" (west of the city of Detroit), "g" (south of the
10      city) and "v" (suburban site in northwestern Wayne County, MI, generally "upwind" of the
11      urban center).  However, as the authors note, all of the reported relative risks are for site-specific
12      increments, which vary by a factor of about  2.5 over the Wayne County - Windsor area. When
13      converted to a common increment of 100 //g/m3 TSP, the largest excess risks are found when the
14      monitor used in the model is "f' (4.5%),  "v" (4.2%), or "w" (3.8%), which also show the most
15      significant effects among the 14 monitors. As the authors note, "... the distributional increments
16      [used] to calculate relative risk tend to standardize the scale of relative risks. This actually makes
17      sense in that if there is a concentration gradient of TSP within a city,  and if the various TSP
18      concentrations fluctuate together, then using a site with a low mean TSP for time-series analysis
19      would result in a larger coefficient. This result does warn against extrapolating the effects from
20      one city to an other using a raw regression coefficient [excess relative risk]"
21           Other recent studies also point out other aspects of intra-urban spatial variation in PM
22      concentrations. Kinney et al. (2000) note that in a personal and ambient PM2 5  and diesel exhaust
23      particle (DEP) exposure study in a dense urban area of New York City, PM25 concentrations
24      showed only a moderate site-to-site variation (37 to 47 //g/m3), probably due to broader regional
25      sources of PM2 5, whereas elemental carbon  concentrations  (EC) showed a four-fold range of site-
26      to-site variations, reflecting the greater local variation in EC from DEP.
27           Several PM health studies for the city  of Seattle (King County), WA (e.g. Levy et al., 2001,
28      for out-of-hospital primary cardiac arrests) have found few statistically significant relationships,
29      attributed by the  authors in part to the fact that Seattle has a topographically diverse terrain with
30      local "hot spots" of residential wood burning,  especially in  winter.  Sheppard et al. (2001) have
31      explored reasons for these findings, particularly focusing on adjustments for location by use of a

        April 2002                                8-262        DRAFT-DO NOT  QUOTE OR CITE

-------
 1      "topographic index" that includes the "downstream" normal flow of wood smoke from higher
 2      elevations, and the trapping of wood smoke in topographic bowls or basins even at higher
 3      elevations.  They also adjusted for weather using a "stagnation index" (the average number of
 4      hours per day with wind speed less than the 25th percentile of wind speeds), and temperature, as
 5      well as interaction terms for stagnation on hilltop sites and temperature at suburban wood-
 6      smoke-exposed valley sites. The adjustments for exposure measurement error based on methods
 7      developed in (Sheppard and Damian, 2000; Sheppard et al., 2001) had little effect on effect size
 8      estimates for the case-crossover study (Levy et al., 2001), but may be useful in other  studies
 9      where localized effects are believed to be important, particularly for the gaseous co-pollutants.
10           Daniels et al. (2001) evaluated the relative sources of variability or heterogeneity in
11      monitoring PM10 in Pittsburgh, PA in  1996. The site is data-rich, having 25 monitors in a
12      rectangle approximately 40 by 80 km. The authors found no isotropic spatial dependence after
13      accounting for other sources of variability, but an indication of heterogeneity in the variability of
14      the small-scale processes over time and space, and heterogeneity in the mean values and
15      covariate effects across sites. Important covariates included temperature, precipitation, wind
16      speed and direction.  The authors concluded that significant unmeasured processes might be in
17      operation. These methods  should also be useful in evaluating the spatial and temporal variations
18      in gaseous co-pollutants,  where small-scale processes are clearly important.
19
20      8.4.7.3 Measurement Error and the Assessment of Confounding by Co-Pollutants in
21             Multi-Pollutant Models.
22           The discussion in Zeger et al. (2000) may be interpreted as addressing the question of
23      whether the apparent lack of a PM10_2 5 effect in models with both fine and coarse particles
24      demonstrates a "false negative" due to the larger measurement error of coarse particle
25      concentrations. However, the more important question may involve the relative attenuation of
26      estimated effects of PM25 and gaseous co-pollutants, especially those such as CO that are known
27      to be highly correlated with PM2 5. Tables 1 and 2 in (Zeger et al., 2000) may be particularly
28      relevant here. The evidence discussed in this chapter supports the hypothesis that PM has
29      adverse health effects, but leaves open the question as to whether the co-pollutants have effects
30      as well when their exposure is measured much less accurately than that of the PM metric.  If both
31      the PM metric and the co-pollutant have effects, Table 1 shows that the co-pollutant effect size
32      estimate may be greatly attenuated and the PM effect size estimate much less so, depending on
        April 2002                                8-263        DRAFT-DO NOT QUOTE OR CITE

-------
 1      the magnitude of the correlation between the true PM and gaseous pollutant exposures, and the
 2      correlation between their measurement errors.  One would expect that PM2 5, CO, and NO2 would
 3      often have a high positive correlation, and their "exposure measurement errors" would also be
 4      positively correlated if PM and the gaseous pollutants were positively correlated due to common
 5      activity patterns, weather, and source emissions.  Thus, the line with corr(x1,x2) = 0.5, var^) =
 6      0.5, var(62) = 2, corr^, 52) = 0.7 seems appropriate.  This implies that the estimated effect of
 7      the more accurately measured pollutant is 64% of the true value, and that of the less accurately
 8      measured pollutant is 14% of the true value. In view of the substantially greater spatial
 9      heterogeneity of traffic-generated ambient pollutants such as CO and NO2, and the relative
10      (though  not absolute) regional spatial uniformity of ambient PM2 5 in some cities, but not in
11      others, it is likely that effect size estimates in multi-pollutant models are attenuated downward to
12      a much greater extent for the gaseous co-pollutants than for the PM metric in some cities, but not
13      in others. This may explain part of the heterogeneity of findings for multi-pollutant models in
14      different cities discussed in Section 8.4.2.2.3. Low effect size estimates for the gaseous co-
15      pollutants in a multi-pollutant model should be interpreted cautiously, as noted in Section
16      8.4.2.2.3. The representativeness of the monitoring  sites for population exposure of both the
17      particle metrics and gaseous pollutants should be evaluated as part of the interpretation of the
18      analysis. Indices such as the maximum 90th percentile of the absolute  difference in
19      concentrations between pairs of sites as well as the median cross-correlation across sites may be
20      useful for characterizing for spatially heterogeneity of gaseous co-pollutants  as well as for fine
21      particles.
22
23      8.4.7.5 Air Pollution Exposure  Proxies in  Long-Term Mortality Studies
24           The AHSMOG Study of mortality (Abbey et al.,  1999; McDonnell et al., 2000), the
25      Harvard 6-Cities Study of mortality (Dockery et al, 1993), the ACS Study (Pope et al., 1995), and
26      the VA/Washington Univ.  Study (Lipfert et al., 2000b) together provided a major step forward in
27      the assessment of the long-term effects of air pollution. These cohort  studies responded to many
28      of the major criticisms of the prior cross-sectional mortality studies, while largely confirming the
29      results of those prior studies. In particular, unlike the ecological cross-sectional studies, these
30      new cohort studies had individual-level information about the members of the study cohort,


        April 2002                                8-264        DRAFT-DO NOT QUOTE OR CITE

-------
 1      allowing the analysis to more properly control for other major factors in mortality, such as
 2      smoking and socio-economic factors.
 3           While several of these studies made use of newly available fine particle mass (PM25) data
 4      to derive useful estimates of health effects of PM25 well before it was routinely measured, these
 5      studies utilized air pollution exposure information in a manner similar to that used in the past
 6      studies. These studies used central site metropolitan area (MA) spatial and time averages of air
 7      pollution exposures, rather than exposure information at the individual level. For this reason, the
 8      AHSMOG, Harvard Six-Cities, ACS, and VA/Washington Univ. studies have been term
 9      "semi-individual" cohort studies of air pollution.
10
11      The AHSMOG Study
12           Although this study covers a large number of years (1977-1992 in Abbey et al., 1999), it is
13      considerably more limited in the availability of particle metrics that were actually observed rather
14      than estimated. Prior to 1987, PM10 could only be estimated from TSP, not observed. Likewise,
15      for the more recent years, McDonnell et al. (2000) used participants who lived near an airport so
16      that PM2 5, and PM10_2 5 as the difference of PM10 and PM25, could be estimated from airport
17      visibility data using the method described in an earlier publication (Abbey et al., 1995b). All of
18      these issues add potential measurement error to the exposure estimates.
19
20      The Veterans' Administration/Washington University Study
21           The air pollution concentrations for the participants' counties of residence at the time of
22      enrollment were used in the analyses, rather concentrations at the 32 VA hospitals in the final
23      study.  County-wide pollution variables for five particle metrics and three gaseous pollutants
24      were used in the study, although TSP was most often the particle metric observed for the earlier
25      years of the study (before 1975 up to!988), which are important in assessing pollution effects for
26      many years of exposure.  However, IPMN data for fine particles and sulfates were available for
27      ca. 1979-1983, as in the ACS study. Effects on average mortality for the intervals 1976-1981,
28      1982-1988, and 1989-1996 were related to multi-year particle exposures for four long intervals:
29      < 1975, 1975-1981,  1982-1988, and 1989-1996.  TSP was used in the first three exposure
30      intervals, PM10 in the most recent. This study examined "concurrent"  exposures (same interval
31      as average mortality), "causal" prior exposures (exposure interval precedes mortality interval),

        April 2002                                8-265        DRAFT-DO NOT QUOTE OR CITE

-------
 1      and "non-causal" PM vs. mortality associations. The mortality associations were also examined
 2      for PM2 5, PM15, and PM15.25 for 1979-1981 and 1982-1984.  This study has a considerable
 3      amount of air pollution data and should be as adequate as other studies for characterizing fixed-
 4      site air pollution concentrations in the place of residence at the time of enrollment.  However, if
 5      any participants moved away from the county where air pollution is measured, but were retained
 6      in the study because they continue to participate in follow-ups at the same clinic, then the use of
 7      the initial residence location may not be an adequate proxy for actual exposure after initial
 8      enrollment.
 9
10      Harvard Six-Cities Air Pollution Exposure Data
11           In the  case of the Harvard Six Cities Study, ambient concentrations of fine particles (PM2 5),
12      total suspended particles (TSP), sulfur dioxide (SO2), ozone (O3), nitrogen dioxide (NO2), and
13      sulfate (SO4=) were measured at a centrally located air monitoring station established within each
14      of the six communities.  Long-term mean concentrations for each pollutant were calculated for
15      periods that were consistent among the six cities, but not across pollutants.  The original
16      epidemiologic analysis characterized ambient air quality as long-term mean concentrations of
17      total particles (TSP) (1977-1985), inhalable and fine particles (1979-1985), sulfate particles
18      (1979-1984), aerosol acidity (H+ ) (1985-1988), sulfur dioxide (1977-1985), nitrogen dioxide
19      (1977-1985), and ozone (1977-1985), as follows:
20      Gases: The gases (SO2, NO2, and O3) were monitored hourly by conventional continuous
21      instrumentation in parts per billion.
22
23      Particles: Mean PM concentrations were reported for four classifications of particles in each of
24      the six cities: TSP (particles with aerodynamic diameters up to 50 //m), inhalable particles, fine
25      particles, and sulfate particles. Values of mass for TSP and sulfate particles were determined
26      from 24-h high-volume  samplers.  Inhalable particle mass was calculated from coarse and fine
27      particle mass, which had been determined from 24-h sample pairs collected by dichotomous
28      samplers. In these, the fine particle channel collected particles smaller than about 2.5 //m and the
29      measurement was recorded directly as fine particle (FP) mass. The coarse particle channel
30      collected particles between 2.5 //m and 10 or 15 //m in aerodynamic diameter (the upper bound
31      measurement depended on the inlet size used at the time).

        April 2002                                8-266       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Acidity: Aerosol acidity (H+) was measured for about one year in each city. However,
 2      measurements were conducted in only two cities at a time. Thus, it was not possible to compare
 3      acidity for a common time period. Furthermore, the acidity data were not linked with particle
 4      data in the same city.  Thus, intercity and inter-pollutant comparisons of FT in this study were
 5      confounded by inter-annual variability.
 6
 7      ACS Study Air Pollution Exposure Data
 8           In the ACS Study (Pope et al.,  1995), two measures of particulate air pollution, were
 9      considered: fine particles and sulfate. No gaseous pollutants were considered. The mean
10      concentration of sulfate air pollution by metropolitan area (MA) during  1980 was estimated using
11      data from the EPA Aerometric Information Retrieval System (AIRS) database.  These means
12      were calculated as the averages of annual  arithmetic mean 24-h sulfate values for all monitoring
13      sites in the 151 MA's considered.  The median concentration of fine particles between 1979 and
14      1983 was estimated from the EPA's dichotomous sampler network. These estimates of fine
15      particle levels had been used previously in a population-based cross-sectional mortality study of
16      50 MA's. Gaseous co-pollutants were not considered in Pope et al's original ACS analysis.
17
18      Six-City Study and ACS Exposure Data Strengths and Weaknesses
19           In each of these studies, there was a single mean pollution concentration assigned for each
20      city for each pollutant for the entire follow-up period considered. Concentrations were not
21      broken into each year or sub-groups  of years (e.g., 5 year averages), largely because data were not
22      available in this form.  This may represent a significant weakness, as a single number could not
23      accurately account for the different exposures in different years of follow-up. However, it is
24      possible that the simultaneous or immediately preceding years alone might not as well represent
25      the effects of long-term pollution exposure.
26           The ACS analysis also uses metropolitan area (MA) pollutant concentrations for air
27      pollution exposure estimates, rather  than individual level measurements. Thus, spatial variability
28      in air pollution levels and potential effects of different housing infiltration rates were not
29      addressed as potential factors in exposure variability. However, individual exposure data would
30      be economically impractical for such large cohorts, and the use of more  localized measurements
31      (e.g., by county) might well lead to more error, due to day-to day mobility between counties by

        April 2002                                8-267       DRAFT-DO NOT QUOTE OR CITE

-------
 1      individuals (e.g., to work and back) and changes of specific residence within an MA over time.
 2      Thus, the MA average may yet be the best metric that can be developed in the absence of
 3      individual  level  exposure data.
 4           Another notable weakness of the original ACS  Study was that only two PM air pollution
 5      metrics were considered. Thus, this study did not consider the potentially confounding
 6      influences  of gaseous air pollutants or other particle indicators.
 7           These two studies' analyses assign the subjects' residence MA on the basis of where they
 8      were enrolled, which can lead to exposure errors if the subjects moved to another MA during the
 9      follow-up period. However, a recent reanalysis of the Six Cities Study cohort (Krewski et al.,
10      2000) indicates that mobility in these older populations is limited, with only 18.5% leaving the
11      original city of enrollment over subsequent decades.
12
13      The HEI Reanalysis of the ACS Study
14           The HEI Reanalysis of these two cohort studies (Krewski et al, 2000) confirmed the
15      databases used in these two studies, but also developed new exposure data for the ACS Study
16      cohort.  In  particular, data for the gaseous pollutants (for the year 1980) were added to the
17      analysis. Table  8-43 below displays summary data for the most recent data available for the
18      analysis of the ACS cohort (Pope et al., 2002). The variables noted with the data source "HEI"
19      were added to the analysis during the HEI reanalysis. These HEI results largely confirmed the
20      original ACS analysis results for PM, but also indicated that SO2 was also correlated with U.S.
21      mortality.
22
23      The 16-Year Follow-Up of the ACS Cohort
24           Also included in Table 8-43 are summaries of the pollutant data developed to provide
25      exposure estimates  for the latest  16-year follow-up analysis of the ACS cohort (Pope et al, 2002).
26      These new data  are  similarly city-wide averages of all monitoring stations in the MA's
27      considered, but for the entire period of follow-up (1982-1998), when possible.  In addition, this
28      new analysis has incorporated the new PM25  air monitoring data collected routinely from 1999
29      onward. As a result, this new analysis has increased  the analysis power both by extending the
30      length of follow-up, and by adding significant new multiple and multi-year air pollution exposure
31      data to the analysis.

        April 2002                               8-268       DRAFT-DO NOT QUOTE OR CITE

-------
          TABLE 8-43.  SUMMARY OF ACS POLLUTION INDICES:  UNITS, PRIMARY
         SOURCES, NUMBER OF CITIES AND SUBJECTS AVAILABLE FOR ANALYSIS,
                         AND THE MEAN LEVELS (standard deviations)
Pollutant
(years of data)
PM25 (79-83)
PM2 5 (99-00)
PM2 5 (ave)
PM10 (82-98)
PM15 (79-83)
PM15.25 (79-83)
TSP (80-81)
TSP (79-83)
TSP (82-98)
SO4 (80-81)
SO4 (90)
SO2 (80)
SO2 (82-98)
NO2 (80)
NO2 (82-98)
CO (80)
CO (82-98)
03 (80)
O3 (82-98)
O3 (82-98 3rd Q.)
Units
Mg/rn3
Mg/nf
Mg/ni3
^tg/rn3
^tg/rn3
Mg/m3
Mg/m3
^g/m3
Mg/m3
Mg/m3
Mg/m3
ppb
ppb
ppb
ppb
ppm
ppm
ppb
ppb
ppb
Sources of Data*
IPMN (HEI)
AIRS (NYU)
Average of two above
AIRS (NYU)
IPMN (HEI)
IPMN (HEI)
NAD (HEI.)
IPMN (HEI)
AIRS (NYU)
IPMN and NAD,
artifact adjusted (HEI)
NYU compilation and
analysis of PM10 filters
AIRS (HEI)
AIRS (NYU)
AIRS (HEI)
AIRS (NYU)
AIRS (HEI)
AIRS (NYU)
AIRS (HEI)
AIRS (NYU)
AIRS (NYU)
No. of Metro
Areas
61
116
51
102
63
63
156
58
150
149
53
118
126
78
101
113
122
134
119
134
No. of Sub.
(1000s)
359
500
319
415
359
359
590
351
573
572
269
520
539
409
493
519
536
569
525
557
Mean (SD)
21.1 (4.6)
14.0 (3.0)
17.7(3.7)
28.8 (5.9)
40.3 (7.7)
19.2(6.1)
68.0(16.7)
73.7(14.3)
56.7(13.1)
6.5 (2.8)
6.2 (2.0)
9.7 (4.9)
6.7 (3.0)
27.9 (9.2)
21.4(7.1)
1.7 (0.7)
1.1(0.4)
47.9(11.0)
45.5 (7.3)
59.7(12.8)
       Source: Pope et al. (2002)
1     Conclusions
2          The pollution exposure data used in these studies, while state-of-the-art when they were
3     conducted, have weaknesses, most notably that these studies, of necessity, have employed city-
4     wide estimates of air pollution exposure, rather than individual-level exposure data. In the case
5     of the mortality control variables (e.g., race and education), the use of individual-level data did
      April 2002
8-269
DRAFT-DO NOT QUOTE OR CITE

-------
 1      not significantly change the air pollution effect estimates from those given by prior "ecological"
 2      cross-sectional mortality analyses using MA aggregate data (e.g., Ozkaynak and Thurston, 1987).
 3      Future research into the human health effects of long-term air pollution exposures needs to
 4      similarly assess whether the use of individual level exposure data would or would not
 5      substantially change the pollution effect estimates.
 6
 7      8.4.9  Heterogeneity of Particulate Matter Effects Estimates
 8           Approximately 35 then-available acute PM exposure community epidemiologic studies
 9      were assessed in the 1996 PM AQCD as collectively demonstrating increased risks of mortality
10      being associated with short-term (24-h) PM exposures indexed by various ambient PM
11      measurement indices (e.g., PM10, PM2 5, BS, COH, sulfates, etc.) in many different cities in the
12      United  States and internationally. Much homogeneity appeared to exist across various
13      geographic locations, with many studies suggesting, for example, increased relative risk (RR)
14      estimates for total nonaccidental mortality on the order of 1.025 to 1.05 (or 2.5 to 5.0% excess
15      deaths) per 50 //g/m3 increase in 24-h PM10, with statistically significant results extending more
16      broadly in  the range of 1.5 to 8.0%.  The elderly > 65 yrs. old and those with preexisting
17      cardiopulmonary conditions had somewhat higher excess risks.  One study, the Harvard Six City
18      Study, also provided estimates of increased RR for total mortality falling in the range of 1.02 to
19      1.056 (2.0  to 5.6% excess deaths) per 25 //g/m3 24-h PM2 5 increment.
20           Now, more than 80 new time-series PM-mortality studies assessed earlier in this chapter
21      provide extensive additional evidence which, qualitatively, largely substantiates significant
22      ambient PM-mortality relationships, again based on 24-h exposures indexed by a wide variety of
23      PM metrics in many different cities of the United States, in Canada, in Mexico, and elsewhere (in
24      South America, Europe, Asia, etc.).  The newly available effect size estimates from such  studies
25      are reasonably consistent with the ranges  derived from the earlier studies reviewed in the 1996
26      PM AQCD.  For example, newly estimated PM10 effects generally fall  in the range of 1.0 to 8.0%
27      excess deaths per 50 //g/m3 PM10 increment in 24-h concentration; whereas new PM25 excess
28      estimates for short-term exposures generally fall in the range of 2 to 8% per 25 //g/m3 increment
29      in 24-h PM2 5 concentration.
30           However, somewhat greater spatial  heterogeneity appears to exist across newly reported
31      study results, both with regard to PM-mortality and morbidity effects.  The newly apparent

        April 2002                               8-270        DRAFT-DO NOT QUOTE OR CITE

-------
 1      heterogeneity of findings across locations is perhaps most notable in relation to reports based on
 2      multiple-city studies in which investigators used the same analytical strategies and models
 3      adjusted for the same or similar co-pollutants and meteorological conditions, raising the
 4      possibility of different findings reflecting real location-specific differences in exposure-response
 5      relationships rather than potential differences in models used, pollutants measured and included
 6      in the models, etc.  Some examples of newly reported and well-conducted multiple-city studies
 7      include: the NMMAPS analyses of mortality and morbidity in 20 and 90 U.S. cities (Samet et al.,
 8      2000a,b; Dominici et al., 2000a); the Schwartz (2000b,c) analyses of 10 U.S. cities; the study of
 9      eight largest Canadian cities (Burnett et al., 2000); the study of hospital admissions in eight U.S.
10      counties (Schwartz, 1999); and the APHEA studies of mortality and morbidity in several
11      European cities (Katsouyanni et al., 1997; Zmirou et al., 1998).  The recently completed large
12      NMMAPS studies of morbidity and mortality in U.S. cities add especially useful and important
13      information about potential U.S. within- and between-region heterogeneity.
14
15      8.4.9.1 Evaluation of Heterogeneity of Particulate Matter Mortality Effect Estimates
16           In all of the U.S. multi-city analyses, the heterogeneity in the PM estimates across cities
17      was not explained by city-specific characteristics in the 2nd stage model. The heterogeneity of
18      effects estimates across cities in the multi-city analyses may be due to chance alone, to mis-
19      specification of covariate effects in small cities, or to real differences from location to location in
20      effects of different location-specific ambient PM mixes, for which no mechanistic explanations
21      are yet known.  Or, the apparent heterogeneity may simply reflect imprecise PM  effect estimates
22      derived from smaller-sized analyses of less extensive available air pollution data or numbers of
23      deaths in some cities tending to obscure more precise effects estimates from larger-size analyses
24      for other locations, which tend to be consistently more positive and statistically significant.
25           Some of these possibilities can be evaluated by using data from the NMMAPS study
26      (Samet et al., 2000b).  Data in Figure 8-3 were optically scanned and digitized, producing
27      reasonably accurate estimates by comparison with the 20 largest U.S. cities in their Table A-2.
28      The cities were divided among  7 regions, and excess risk with 95% confidence intervals plotted
29      against the total number of effective observations, measured by the  number of days of PM10 data
30      times the mean number of daily deaths in the community.  This provides a useful measure of the
31      weight that might be assigned to the results, since the uncertainty of the RR estimate based on a

        April 2002                                8-271        DRAFT-DO NOT QUOTE OR CITE

-------
 1      Poisson mean is roughly inversely proportional to this product. That is, the expected pattern
 2      typically shows less spread of estimated excess risk with increasing death-days of data. A more
 3      refined weight index would also include the spread in the distribution of PM concentrations. The
 4      results are plotted in Figure 8-30 for all cities and Figure 8-31 for each of the 7 regions.
 5           Figure 8-30 for all cities suggests some relationship between precision of the effects
 6      estimates and study weight, overall. That is, the more the mortality-days observations, the
 7      narrower the 95% confidence intervals and the more precise the effects estimates (with nearly all
 8      these for cities with > log 9 mortality-days being positive and many statistically significant at
 9      p < 0.05).
10           The Figure 8-31 depiction for each of the 7 regions is  also informative.  In the Northeast,
11      there is considerable homogeneity (not heterogeneity) of effect size for larger study-size cities,
12      even with moderately wide confidence intervals for those with log mortality-days = 8 to 9, and all
13      clearly exceed the overall nationwide grand mean indicated by the dashed line. On the other
14      hand, the smaller study-size Northeast cities (with much wider confidence intervals at log
15      < 8) show much greater heterogeneity of effects estimates and less precision.  Also, most of the
16      estimates for larger study-size (log > 9) cities in the industrial midwest are positive and several
17      statistically significant, so that an overall significant regional risk is plausible there as well. There
18      may even be some tendency for relatively large risks for some cities with small study sizes and
19      wide confidence intervals in the industrial midwest, and further investigation of that would be of
20      interest. The plot for Southern California in Figure 8-31 clearly shows a rather consistent
21      estimate of effect size and width of the confidence intervals across cities of varying study-size.
22      All risk estimates are positive and most are significant at p < 0.05 or nearly so for the Southern
23      California cities. For Northwestern cities plotted in Figure 8-31, the value for Oakland, CA (at
24      ca. log 9.5) is notable (it being very positive and significant), whereas many but not all of the
25      other cities have positive effect estimates not too far off the nationwide grand mean, but with
26      sufficiently wide confidence intervals so as not to be statistically significant at p <  0.05. The
27      Southwestern cities (except for 2 cities), too, mostly appear to have effect sizes near the
28      nationwide mean, but with confidence intervals too wide to be significant at p < 0.05. The
29      "Other" (non-industrial or "Upper", as per NMMAPS) Midwest cities  and the Southeastern cities
30      in Figure 8-31 show more heterogeneity, although most of the larger study size cities (log > 9.0)
31      tend to be positive and not far off the nationwide mean (even though not significant at p < 0.05).

        April 2002                                8-272         DRAFT-DO NOT QUOTE OR CITE

-------
      5.4
CO
       1.8-
      0.0
 C/)
 CO
 0
 O
 X
 LU
     -5.4
            I         I         I
              All  Cities
     -1.8-
                   Natural Log of Mortality (Days)

Figure 8-30.  The EPA-derived plot showing relationship of PM10 total mortality effects
            estimates and 95% confidence intervals for all cities in the Samet et al.
            (2000a,b) NMMAPS 90-cities analyses in relation to study size (i.e., the
            natural logarithm or numbers of deaths times days of PM observations).
            Note generally narrower confidence intervals for more homogeneously
            positive effects estimates as study size increases beyond about In (mortality-
            days) (i.e., beyond about 8,000 deaths-days of observation). The dashed line
            depicts the overall nationwide effect estimate (grand mean) of approximately
            0.5% per 10 fJ-g/m3 PM10 for models with no co-pollutants.
April 2002
8-273
DRAFT-DO NOT QUOTE OR CITE

-------
5
CL
CO
| 1.8
8.0.0
« -1.8
'(T
Ul
Ul
01
o
X
-
-




Northwest
[fiV-<

s
-

-





• J










Industrial Midwest
JU T T -T-
iff Ti I ^
1 -"-11







-

-






i


_









-

i i i
Northeast
tiff I"
hp-

i i i
     7      8     9     10     11  7   8    9   10   11   12  13  6   7    8    9   10   11    12
       Natural Log of Mortality (Days)      Natural Log of Mortality (Days)      Natural Log of Mortality (Days)
   5.4
CL
CO
a o.o
tr
Ul
Ul
01
o
X
1 1 1
Southern California
"TI 1
11
-
ill
i *
-




•
-

- .
Other Midwest


i





i
• i

T
	 -f--'





j
• '

•
v
•



j

L_L




i
Southeast





h ^R

-L
8     9     10     11     12
  Natural Log of Mortality (Days)
  7    8    9   10   11   12
Natural Log of Mortality (Days)
                                                                     8      9     10     11
                                                                 Natural Log of Mortality (Days)
CL
CO
   5.4
   1.8
8.0.0
«  -1'8

ui
ui

x
UJ  r- ,
                      Southwest
    '8.0    8.5     9.0    9.5    10.0
        Natural Log of Mortality (Days)

Figure 8-31.  The EPA-derived plots showing relationships of PM10-mortality (total,
              nonaccidental) effects estimates and 95% confidence intervals to study size
              (defined as Figure 8-10) for cities broken out by regions as per the NMMAPS
              regional analyses  of Samet et al. (2000a,b).  Dashed line on each plate depicts
              overall nationwide effect estimate (grand mean) of approximately 0.5% per
              10 Aig/m3 PM10 for models with no co-pollutants.
April 2002
      8-274
                                                  DRAFT-DO NOT QUOTE OR CITE

-------
 1      Given the wide range of effects estimates and confidence intervals seen for Southeastern cities,
 2      further splitting of the region might be informative.
 3           In fact, closer reexamination of results for each of the regions may reveal interesting new
 4      insights into what factors may account for any apparent disparities among the cities within a
 5      given region or across regions. Several possibilities readily come to mind. First, cursory
 6      inspection of the mean PM10 levels shown for each city in (Samet et al., 2000b; Appendix A)
 7      suggests that many of the cities showing low effects estimates and wide confidence intervals tend
 8      to be among those having the lowest mean PM10 levels and, therefore, likely the smallest range of
 9      PM10 values across which to distinguish any PM-related effect, if present.  It may also be possible
10      that those areas with higher PM25 proportions of PM10 mass (i.e., larger percentages of fine
11      particles) may show higher effects estimates (e.g., in Northeastern cities) than those with higher
12      coarse-mode fractions (e.g., as would be more typical of Southwestern cities). Also, more
13      industrialized cities with greater fine-particle emissions from coal combustion (e.g., in the
14      industrial Midwest) and/or those with high fine-particle emissions from heavy motor vehicle
15      emissions (e.g., typical of Southern California cities) may show larger PM10 effects estimates
16      than other cities.  Lastly, the extent of air-conditioning use may also account for some of the
17      differences, with greater use in many Southeastern and Southwestern cities perhaps decreasing
18      actual human exposure to ambient particles present versus higher personal exposure to ambient
19      PM (including indoors) in those areas where less air-conditioning is used (e.g., the Northeast and
20      industrial Midwest).  See, for example, Janssen  et al. (2002) results reproduced as Figure 8-11.
21
22      8.4.9.2 Comparison of Spatial Relationships in the NMMAPS and Cohort Reanalyses
23             Studies
24           Both the NMMAPS and HEI Cohort Reanalyses studies had a sufficiently large number of
25      U.S. cities to allow considerable resolution of regional PM effects within the "lower 48" states,
26      but an attempt was made to take this approach to a much more detailed level in the Cohort
27      Reanalysis studies than in NMMAPS. There were: 88 cities with PM10 effect size estimates in
28      NMMAPS; 50 cities with PM25 and 151 cities with sulfates in the original Pope et al. (1995)
29      ACS analyses and in the HEI reanalyses using the original data; and 63 cities with PM2 5 data and
30      144 cities with sulfate data in the additional analyses done by the HEI Cohort Reanalysis team.
31      The relatively large number of data points utilized in the HEL reanalyses effort and additional

        April 2002                                 8-275       DRAFT-DO NOT QUOTE OR CITE

-------
 1      analyses allowed estimation of surfaces for elevated long-term concentrations of PM2 5, sulfates,
 2      and SO2 with resolution on a scale of a few tens to hundreds of kilometers.
 3           The patterns for PM2 5 and sulfates are similar, but not identical. In particular, the modeled
 4      PM25 surface (Krewski et al., 2000; Figure 18) has peak levels around Chicago - Gary, in the
 5      eastern Kentucky - Cleveland region, and around Birmingham AL, with elevated but lower PM2 5
 6      almost  everywhere east of the Mississippi, as well as southern California. This is similar to the
 7      modeled sulfate surface (Krewski et al., 2000; Figure 16), with the absence of a peak in
 8      Birmingham and an emerging  sulfate peak in Atlanta.  The only area with markedly elevated SO2
 9      concentrations is the Cleveland - Pittsburgh region. A preliminary evaluation is that secondary
10      sulfates in particles derived from local SO2 are more likely to be important in the industrial
11      midwest, south from the Chicago - Gary region into Ohio, northeastern Kentucky, West Virginia,
12      and southwest Pennsylvania, possibly related to combustion of high-sulfur fuels.
13           The overlay of mortality with  air pollution patterns is also of much interest. The spatial
14      overlay of long-term PM2 5 and mortality (Krewski et al., 2000; Figure 21) is highest from
15      southern Ohio to northeastern Kentucky /West Virginia, but also includes a significant association
16      over most of the industrial midwest from Illinois to the eastern non-coastal parts of North
17      Carolina, Virginia, Pennsylvania, and New York. This is reflected, in diminished form, by the
18      sulfates and SO2 maps (Krewski et al., 2000; Figures 19 and 20), where there appears to be a
19      somewhat tighter focus of elevated risk in the upper Ohio River Valley area. This suggests that,
20      while S02 may be an important precursor of sulfates in this region, there may also be some other
21      (non-sulfur) contributors to associations between PM2 5 and long-term mortality, embracing a
22      wide area of the North Central Midwest and non-coastal Mid-Atlantic region.
23           It should be noticed that, while a variety of spatial modeling approaches were discussed in
24      the NMMAPS methodology report (NMMAPS Part I, pp. 66-71 [Samet et al., 2000a]), the
25      primary spatial analyses in the 90-city study (NMMAPS, Part II [Samet et al., 2000b]) were
26      based on a simpler seven-region breakdown of the contiguous 48 states. The 20-city results
27      reported for the  spatial model in NMMAPS I show a much smaller posterior probability of a
28      PM10 excess risk of short-term mortality, with a spatial posterior probability vs.  a non-spatial
29      probability of a PM10 effect of 0.89 vs. 0.98 at lag 0, of 0.92 vs. 0.99 at lag 1, and of 0.85 vs. 0.97
30      at lag 2. The evidence that PM10 is associated with an excess short-term mortality risk is still
31      moderately strong with a spatial model, but less strong than with a non-spatial model.

        April 2002                               8-276       DRAFT-DO NOT QUOTE OR CITE

-------
 1           The apparently substantial differences in PM10 and/or PM2 5 effect sizes across different
 2      regions should not be attributed merely to possible variations in measurement error or other
 3      statistical artifact(s). Some of these differences may reflect: real regional differences in particle
 4      composition or co-pollutant mix; differences in relative human exposures to ambient particles or
 5      other gaseous pollutants; sociodemographic differences (e.g., percent of infants or elderly in
 6      regional population); or other important, as of yet unidentified PM effect modifiers.
 7
 8      8.4.9.3  Epidemiologic Studies of Ambient Air Pollution Interventions
 9           To date, assessment of health risk in epidemiologic studies of ambient air pollutants,
10      including PM, has relied largely on studies that focus on increases in exposure, and that inquire
11      whether health risk changes in relation to such increases.  Such studies are used to support
12      qualitative and quantitative inference as to whether decreases in exposure will bring about
13      reductions in health risk, or improvement in health status.
14           Ambient criteria air pollutants are rarely, if ever, the only etiology of the health disorders
15      with which exposure to these pollutants is associated.  For example, numerous reports have
16      implicated ambient air  pollution exposure with exacerbations of pre-existing asthma.  These
17      reports justify the expectation that further reduction in ambient air pollution exposure would
18      reduce the public health burden of asthma exacerbations.  However, many other factors,
19      including allergens, passive smoking, exercise, cold, and stress are also associated with such
20      exacerbations. Asthmatics would continue to be exposed to these factors even with further
21      reduction in ambient air pollution exposure.
22           Thus, reduction of ambient air pollution exposure, even to zero concentration, would not
23      bring about zero risk of the health disorders with which such exposure is associated.  Also,  it is
24      likely that at least some non-pollution risk factors would behave differently in the absence of
25      ambient air pollution exposure as in its presence. That is, in the real world, risk factors probably
26      do not behave in discrete, additive fashion.
27           Therefore, truly quantitative characterization of effects of reduction in air pollution
28      concentrations and exposures requires study of situations in which such reductions actually
29      occur.  In such studies,  it is important to measure both exposure and health status before and after
30      exposure is reduced. It is also highly desirable to identify risk factors other than ambient air
31      pollution, and to ascertain their effects before and after air pollution exposure reduction.

        April 2002                                  8-277       DRAFT-DO NOT QUOTE OR CITE

-------
 1           In his classic monograph (The Environment and Disease:  Association or Causation?), Hill
 2      (1965) addressed the topic of preventive action and its consequences under Aspect 8, stating:
 3           "Experiment: Occasionally it is possible to appeal to experimental, or semi-experimental,
 4           evidence. For example, because of an observed association some preventive action is taken. Does
 5           it in fact prevent? The dust in the workshop is reduced, lubricating oils are changed,  persons stop
 6           smoking cigarettes. Is the frequency of the associated events affected? Here the strongest support
 7           for the causation hypothesis may be revealed."
 8           The available epidemiologic literature on ambient air pollution offers a limited evidence
 9      related to this aspect. In these studies, air pollution concentrations have been temporarily or
10      permanently reduced through regulatory action, industrial shutdown, or other intervening
11      factor(s).
12           In the U.S., the most thoroughly studied example of such ambient air pollution reduction
13      occurred in the Utah Valley, UT, during the 1980s.  The Valley's largest stationary source of PM,
14      a steel mill, was closed due a labor dispute for 13 months from autumn 1986 until autumn 1987.
15      This offered the opportunity to study health effects not only of the closure-related reduction in
16      ambient PM concentrations, but also of the increases in PM that occurred after the re-opening of
17      the mill. Pope et al. have reported extensively on such health effects. These reports  were
18      addressed in more detail in the 1996 PM AQCD than in the present document.  Briefly, these
19      investigators observed reduction in frequency of a variety of health disorders during  the period in
20      which the mill was closed.  These included daily mortality (Pope et al., 1992), respiratory
21      hospital admissions  (Pope, 1989), bronchitis and asthma admissions for preschool children
22      (Pope,  1991), reductions in lung function (Pope et al., 1991), and elementary school  absences
23      (Ransom and Pope,  1992).  Changes in these endpoints were reflected by differing strength of
24      positive associations between measures of these health endpoints and PM mass measurements
25      from filters collected before, during, and after the steel mill shut down.
26           Five experimental studies investigated effects of aqueous extracts of ambient Utah Valley
27      particulate filters employing filter extracts from January through March 1986 (mill open), 1987
28      (mill closed), and 1988 (mill open) (Frampton et al., 1999; Dye et al., 2001; Soukup  et al., 2000;
29      Wu et al., 2001; and Ohio and Devlin, 2001). In all of these studies, investigators observed less
30      intense in vivo or in vitro effects when treating with the 1987 extracts than when treating with
31      extracts from 1986 and/or 1987 (see Chapter 7 of this document).

        April 2002                                8-278        DRAFT-DO NOT  QUOTE OR CITE

-------
 1           Frampton et al. (1999) state that extracts were taken from filters collecting PM10, and that a
 2      total of 36 filters was used, 12 per year.  Soukup et al. (2000) state that PM10 filters were used,
 3      and that 34 filters per year were used (total 102 filters).  Dye et al. (2001) state that TSP filters
 4      were extracted, and that 12 filters per year were used (total 36 filters). Wu et al. (2001) state that
 5      PM10 filters were used, and a total of 102 filters was used.  Ohio and Devlin (2001) state that
 6      "filters containing PM10" were extracted, and that 34 filters each year were used (total 102
 7      filters).  Taken together, these descriptions raise the question whether the two studies that
 8      employed 12 filters per year (Frampton et al. and Dye et al.) were using TSP filters exclusively,
 9      whereas the other three studies, that employed 34 filters per year, employed a mixture of TSP
10      filters and PM10 filters. In any event, the degree of comparability of source filters among these
11      five studies is not entirely clear. Also, there is some uncertainty as to the within-study
12      comparability of filters from year to year, particularly in the studies that employed 34 filters per
13      year. Furthermore, a substantial proportion of the extracted material was probably derived from
14      filter matrix, not ambient PM, and about 10 years elapsed between collection and extraction of
15      the filter samples.
16           Even so, the combined results of these five experimental studies provide support and
17      corroboration for the epidemiologic observations of reduced frequency and severity of health
18      disorders during the period of steel mill  closure. The experimental studies also provide
19      hypotheses regarding potential biological mechanisms underlying some of the observed effects.
20      Perhaps the strongest of these hypotheses is that PM-associated metals were etiologically related
21      to some of the observed disorders, and that reduction in ambient concentrations of these metals
22      was at least partially responsible for the  health benefits  observed during steel mill closure.  In any
23      event, these experimental studies underscore the importance of particle composition in
24      production or promotion of harmful health effects (Beckett, 2001).
25           Avol and colleagues investigated effects of reductions and increases in ambient air
26      pollution concentrations on longitudinal lung function growth in a subsample of participants in
27      the Children's Health Study conducted by the University of Southern California (Avol et al.,
28      2001).  Follow-up lung function tests were administered to 110 children who had moved away
29      from the study area after the baseline lung function test, which was administered while the
30      children lived within the study area. Lung function growth rates were analyzed against
31      differences between the children's original and new communities in annual average

        April 2002                                8-279       DRAFT-DO NOT QUOTE OR CITE

-------
 1      concentrations of PM10, NO2, and O3. Analytical models were adjusted for anthropometric
 2      variables and other relevant covariates. No multi-pollutant analyses were reported.  Moving to a
 3      community with lower ambient PM10 concentration was associated with increased growth rates of
 4      FVC, FEV1, MMEF and PEFR, and moving to a community with higher PM10 concentrations
 5      was associated with decreased growth of these metrics.  These associations were statistically
 6      significant for MMEF and PEFR, and appear to have been marginally significant for FVC and
 7      FEV1.  Moving to a community with lower ambient NO2 or O3 concentration was generally
 8      associated with increased lung function growth, and vice versa. However, associations of change
 9      in lung function growth with change in community levels of NO2 and O3 were not statistically
10      significant.  This study suggests that reduction in long-term  ambient PM10 levels is indeed
11      associated with improvement of children's lung growth, and that increase in these levels is
12      associated with retardation of lung growth.
13          Friedman et al. (2001) investigated the influence of temporary changes in transportation
14      behaviors (instituted to reduce downtown traffic congestion during the 1996 Summer Olympic
15      Games in Atlanta,  GA) on ambient air quality and acute care visits and hospitalizations for
16      asthma in children residing in Atlanta.  Ambient air  quality and childhood asthma during the
17      17 days of the Games were compared to those during a baseline period consisting of the four
18      weeks before and the four weeks after the Games. During the  Games, concentrations of PM10
19      (24-h average), O3 (daily peak 1-h average), CO (8-h average), and NO2 (daily peak 1-h average)
20      were, respectively, 16.1%, 27.9%, 18.5%,  and 6.8% lower than during the baseline period.
21      Twenty-four hour average concentrations of SO2 were 22.1% higher during  the Games than
22      during the baseline period. Reductions in O3, PM10, and carbon monoxide were statistically
23      significant at alpha = 0.05 (p = 0.01, p < 0.001, and p = 0.02, respectively).  Ambient mold
24      counts during the Games did not differ significantly from those during the baseline period. Four
25      sources of asthma frequency data were examined: (1) the Georgia Medicaid claims file; (2) files
26      of a health maintenance  organization; (3) emergency department records for two of Atlanta's
27      three pediatric hospitals; and (4) the  Georgia Hospital Discharge Database.  For all four sources,
28      asthma-related unadjusted and adjusted relative risks during the Games were less than 1 (as
29      compared to RR = 1  during the baseline period).  Relative risks from the Medicaid database were
30      statistically significant (p < 0.005), and those from the HMO approached significance (p < 0.10).
31      These findings suggest strongly that, in Atlanta in summer 1996, temporary improvement in

        April 2002                               8-280        DRAFT-DO NOT QUOTE OR CITE

-------
 1      ambient air quality contributed to temporary reduction in severity of pre-existing asthma. This
 2      reduction could not be attributed specifically to any individual air pollutant. In the opinion of
 3      Friedman et al., reductions in morning rush-hour traffic played an important role in reduction of
 4      asthma-related visits and hospitalizations.
 5           Heinrich et al. (2000) studied effects of long-term air pollution reduction in the former East
 6      Germany on prevalence of respiratory illnesses and symptoms in 5 to 14 year-old children.
 7      Cross-sectional surveys were conducted in 1992-1993 and 1995-1996 in three areas, all of which
 8      experienced reductions in annual mean ambient SO2 and TSP concentrations in the time interval
 9      between the surveys.  Percentage reductions in SO2 and TSP were substantial, ranging from
10      about 40%-60% and about 20%-35%, respectively, in the three areas. Longitudinal changes were
11      not measured for size-specific PM metrics.  After adjustment for relevant covariates, statistically
12      significant temporal decreases in prevalences of bronchitis,  otitis media, frequent colds, and
13      febrile infections were observed.
14           In Hong Kong, a regulation prohibiting the use of fuel oil containing more than 0.5% sulfur
15      by weight went into effect in July 1990. Investigators from the University of Hong Kong studied
16      respiratory health in children and non-smoking women before and after the regulation was
17      implemented.  In a relatively polluted district (District A), the regulation resulted in rapid and
18      substantial reduction in the ambient concentration of sulfur dioxide, and in appreciable but less
19      marked reduction in the concentration of sulfate ion in "respirable suspended particulates" (RSP,
20      thought to be  equivalent to PM10).  Percentage reductions in these sulfur-containing pollutants
21      were considerably smaller in a less polluted district (District B). The regulation was not
22      accompanied  by appreciable reductions in levels of PM metrics (TSP and RSP) in either district.
23           Tarn et al. (1994) reported that the prevalence of bronchial hyperreactivity (BHR) in
24      children (as defined by a > 20% drop in FEV1  in response to histamine challenge)  was higher in
25      District A than in District B, even  after exclusion of children with wheeze and asthma. Wong
26      et al. (1998) measured BHR prevalence rates in these districts in 1991 and 1992, and compared
27      these to rates  before the regulation was implemented.  In both districts, BHR prevalence was
28      statistically significantly lower in 1991 than before the intervention. In 1992, the pre- to post-
29      intervention decrease in BHR prevalence was significantly larger in District A than in
30      District B. Peters et al. reported that before the intervention, prevalences of children's respiratory
31      symptoms (e.g., cough, sore throat, wheeze) were statistically significantly higher in District A

        April 2002                                8-281        DRAFT-DO NOT QUOTE OR CITE

-------
 1      than in District B.  About one year after the intervention, there were greater pre- to post-
 2      intervention declines in prevalences of cough or sore throat, phlegm, and wheezing in District A
 3      than in District B.  Wong et al. reported that before the intervention, the prevalence of poor
 4      respiratory health in non-smoking women was significantly higher in District A than in District
 5      B. Also, effects of passive smoking on the women's respiratory health were stronger in District
 6      A than in District B, but not significantly so.  About one year after the intervention, declines in
 7      frequency of poor respiratory health were observed, but these declines did not differ significantly
 8      between districts.  Taken together, these Hong Kong studies suggest that reduction of sulfur in
 9      fuel oil brought about appreciable improvement in children's respiratory health, and discernible
10      but lesser improvement in non-smoking women's respiratory health.  These studies also suggest
11      that these benefits were associated with reduction in sulfur-containing ambient air pollutants, but
12      not necessarily with reduction in TSP or RSP per se.
13           Taken together, these epidemiologic intervention studies lend confidence that further
14      reduction of ambient air pollution exposures  in the U. S. would benefit public health. It is likely
15      that such reduction would bring about both respiratory and cardiovascular health benefits.
16      Available studies also give reason to expect that further reductions in both particulate and
17      gaseous air pollutants would benefit health. On balance, these studies suggest that selective
18      reduction in ambient PM concentrations might well bring about greater benefit than would
19      selective reduction in concentrations of other ambient criteria air pollutants.  Furthermore, the
20      experimental studies of Utah Valley filter extracts point to PM-associated metals as a likely
21      cause or promoter of at least some of the health disorders associated with ambient PM. Beyond
22      this, available epidemiologic intervention studies do not yet give direct,  quantitative evidence as
23      to the relative health benefits that would result from selective reduction of specific PM size
24      fractions.  Also, these studies do not yet provide firm grounds for quantitative prediction of the
25      relative health benefits of single-pollutant reduction strategies vs. multi-pollutant reduction
26      strategies. Even in an almost ideal "natural experiment" such as Utah Valley, potentially
27      confounding factors other than ambient PM concentrations also changed during the steel mill
28      closure. These included concentrations of other pollutants and possible  changes in population
29      due to out- and in-migration influenced by the closing and re-opening of the steel mill.  While
30      changes in ambient PM concentrations undoubtedly played a role, other factors may also have
31      modified the size of the changes in health effects.

        April 2002                                8-282       DRAFT-DO NOT QUOTE OR CITE

-------
 1      8.5 KEY FINDINGS AND CONCLUSIONS DERIVED FROM
 2          PARTICULATE MATTER EPIDEMIOLOGY STUDIES
 3           It is not possible to assign any absolute measure of certainty to conclusions based on the
 4      findings of the epidemiology studies discussed in this chapter. However, these observational
 5      study findings would be further enhanced by supportive findings of causal studies from other
 6      scientific disciplines (dosimetry, toxicology, etc.), in which other factors could be eliminated or
 7      controlled, as discussed in Chapters 6 and 7. The epidemiology studies discussed in this chapter
 8      demonstrate biologically-plausible responses in humans exposed at ambient concentrations. The
 9      most salient conclusions derived from the PM epidemiology studies include:
10
11      (1)   A large and reasonably convincing body of epidemiology evidence confirms earlier
12           associations between short- and long-term ambient PM10 exposures (inferred from
13           stationary air monitor measures) and mortality/morbidity effects and suggest that PM10
14           (or one or more PM10 components) is a probable contributing cause of adverse human
15           health effects.
16      (2)   It is likely that there is significant spatial heterogeneity in the city-specific excess risk
17           estimates for the relationships between short-term ambient PM10 concentrations and acute
18           health effects. The reasons for such variation in effects estimates are not well  understood at
19           this time, but do not negate ambient PM's likely causative contribution to observed PM-
20           mortality and/or morbidity associations in many locations.  Possible factors contributing to
21           the heterogeneity include geographic differences in air pollution mixtures, composition of
22           PM components, and personal and sociodemographic factors affecting PM exposure (such
23           as use of air conditioners, education, and so on).
24      (3)   A growing body of epidemiology studies confirm associations between short- and long-
25           term ambient PM2 5 exposures (inferred from stationary air monitor measures) and adverse
26           health effects and suggest that PM2 5 (or one  or more PM2 5 components) is a probable
27           contributing cause of observed PM-associated health effects. Some new epidemiology
28           findings also suggest that health effects are associated with mass or number concentrations
29           of ultrafme (nuclei-mode) particles, but not necessarily more so than for other ambient fine
30           PM components.
        April 2002                                8-283        DRAFT-DO NOT QUOTE OR CITE

-------
 1      (4)   A smaller body of evidence appears to support an association between short-term ambient
 2           thoracic coarse fraction (PM10_2 5) exposures (inferred from stationary air monitor measures)
 3           and short-term health effects in epidemiology studies. This suggests that PM10_2 5, or some
 4           constituent component(s) of PM10_25, may be a contributory cause of adverse health effects
 5           in some locations.  Reasons for differences among findings on coarse-particle health effects
 6           reported for different cities are still poorly understood, but several of the locations where
 7           significant PM10_2 5 effects have been observed (e.g., Phoenix, Mexico City,  Santiago) tend
 8           to be in drier climates and may have contributions to observed effects due to higher levels
 9           of organic particles from biogenic processes (endotoxins, molds, etc.) during warm months.
10           Other studies suggest that coarse thoracic fraction (PM10_2 5) particles of crustal origin are
11           generally unlikely to exert notable health effects under most ambient exposure conditions,
12           (however, see Item 14, below). Also, in some western U.S. cities where PM10_25 is a large
13           part of PM10, the relationship between hospital admissions and PM10 may be an indicator of
14           response to coarse thoracic particles from wood burning.
15      (5)   Long-term PM exposure durations, on the order of months to years, as well  as on the order
16           of a few days, are statistically associated with serious human health effects (indexed by
17           mortality, hospital admissions/medical visits, etc.). More chronic PM exposures, on the
18           order of years or decades, appear to be associated with life shortening well beyond that
19           accounted for by the simple accumulation of the more acute effects of short-term PM
20           exposures (on the order of a few days).  Some uncertainties remain regarding the magnitude
21           of and mechanisms underlying chronic health effects of long-term PM exposures and the
22           relationship between chronic exposure and acute responses to short-term exposure.
23      (6)   Recent investigations of the public health implications of such chronic PM exposure-
24           mortality effect estimates were also reviewed.  Life table calculations by Brunekreef (1997)
25           found that relatively small differences in long-term exposure to airborne PM of ambient
26           origin can have substantial effects on  life expectancy. For example, a calculation for the
27           1969-71 life table for U.S. white males indicated that a chronic exposure increase of
28           10 //g/m3 PM was associated with a reduction of 1.31 years for the entire population's life
29           expectancy at age 25. Also, new evidence of associations of PM exposure with infant
30           mortality (Bobak and Leon, 1992, 1999; Woodruff et al., 1997; Loomis et al., 1999) and/or
31           intrauterine growth retardation (Dejmek et al., 1999) and consequent increase risk for many

        April 2002                                 8-284       DRAFT-DO NOT QUOTE OR CITE

-------
 1           serious health conditions associated with low birth weight, if further substantiated, would
 2           imply that life shortening in the entire population from long-term PM exposure could well
 3           be significantly larger than that estimated by Brunekreef (1997).
 4      (7)   Considerable coherence exists among effect size estimates for ambient PM health effects.
 5           For example, results derived from several multi-city studies, based on pooled analyses of
 6           data combined across multiple cities (thought to yield the most precise estimates of mean
 7           effect size), show the percent excess total (non-accidental) deaths estimated per 50 //g/m3
 8           increase in 24-h PM10 to be: 2.3% in the 90 largest U.S. cities (4.5% in the Northeast U.S.
 9           region); 3.4% in 10 large U.S. cities; 3.5% in the 8 largest Canadian cities; and 2.0% in
10           western European cities (using PM10 = TSP*0.55). These combined estimates are
11           consistent with  the range of PM10 estimates previously reported in the 1996 PM AQCD.
12           These and excess risk estimates from many other individual-city studies, generally falling
13           in the range of ca. 1.5 to 8.0% per 50 //g/m3 24-h PM10 increment, also comport well with
14           numerous new studies confirming increased cause-specific cardiovascular- and respiratory-
15           related mortality. They are also coherent with larger effect sizes reported for cardiovascular
16           (in the range of ca. 3.0 to 10.0% per  50 //g/m3 24-h PM10 increment) and respiratory (in the
17           range of ca. 5 to 25% per 50 //g/m3 24-h PM10) hospital admissions and visits, as would be
18           expected for these morbidity endpoints versus those for PM10-related mortality.
19      (8)   Several independent panel studies (but not all) that evaluated temporal associations
20           between PM exposures and measures of heart beat rhythm in elderly subjects provide
21           generally consistent indications of decreased heart rate variability (HRV) being associated
22           with ambient PM exposure (decreased HRV being an indicator of increased risk for serious
23           cardiovascular outcomes, e.g., heart attacks). Other studies point toward changes in blood
24           characteristics (e.g., C-reactive protein levels) related to increased risk of ischemic heart
25           disease also being associated with ambient PM exposures. However, these heart rhythm
26           and blood characteristics findings should currently be viewed as providing only limited or
27           preliminary support for PM-related cardiovascular effects.
28      (9)   Notable new evidence now exists which substantiates positive associations between
29           ambient PM concentrations and increased respiratory-related hospital admissions,
30           emergency department, and other medical visits, particularly in relation to PM10 levels.
31           Of much interest are new findings tending to implicate not only fine particle components

        April 2002                                8-285       DRAFT-DO NOT QUOTE OR CITE

-------
 1           but also coarse thoracic (e.g., PM10_2 5) particles as likely contributing to exacerbation of
 2           asthma conditions.  Also of much interest are emerging new findings indicative of likely
 3           increased occurrence of chronic bronchitis in association with (especially chronic) PM
 4           exposure. Also of particular interest are reanalyses or extensions of earlier prospective
 5           cohort studies of long-term ambient PM exposure effects which demonstrate substantial
 6           evidence for association of increased lung cancer risk with such PM exposures, especially
 7           exposure to fine PM or its subcomponents.
 8      (10) One major methodological issue affecting epidemiology studies of both short-term and
 9           long-term PM exposure effects is that ambient PM of varying size ranges is typically found
10           in association with other air pollutants, including gaseous criteria pollutants (e.g, O3, NO2,
11           SO2, CO), air toxics, and/or bioaerosols.  Available statistical methods for assessing
12           potential confounding arising from these associations may not yet be fully adequate. The
13           inclusion of multiple pollutants often produces statistically unstable estimates.  Omission of
14           other pollutants may incorrectly attribute their independent effects to PM. Second-stage
15           regression methods may have certain pitfalls that have not yet been fully evaluated. Much
16           progress in sorting out relative contributions of ambient PM components versus other
17           co-pollutants is nevertheless being made and, overall, tends to substantiate that observed
18           PM effects are at least partly due to ambient PM acting alone or in the presence of other
19           covarying gaseous pollutants.  However, the statistical association of health effects with
20           PM acting alone or with other pollutants should not be taken as an indicator of a lack of
21           effect of the other pollutants.  Indeed, the effects of the other pollutants may at times be
22           greater or less than the effects attributed to PM and may vary from place to place or from
23           time to time.
24      (11) It is possible that differences in observed health effects will be found to depend on site-
25           specific differences in chemical and physical  composition characteristics of ambient
26           particles and on factors affecting exposure (such as air conditioning) as well as on
27           differences in PM mass concentration.  For example, the Utah Valley study  (Dockery et al.,
28           1999; Pope et al., 1991, 1999b) showed that PM10 particles, known to be richer in metals
29           during exposure periods while the steel mill was operating, were more highly associated
30           with adverse health effects than was PM10 during the PM exposure reduction while the steel
31           mill was closed. In contrast, PM10 or PM2 5 was relatively higher in crustal particles during

        April 2002                                 8-286       DRAFT-DO NOT QUOTE OR CITE

-------
 1           windblown dust episodes in Spokane and in three central Utah sites than at other times, but
 2           was not associated with higher total mortality. These differences require more research that
 3           may become more feasible as the PM2 5 sampling network produces air quality data related
 4           to speciated samples.
 5      (12)  The above reasons suggest it is inadvisable to pool epidemiology studies at different
 6           locations, different time periods, with different population sub-groups, or different health
 7           endpoints, without assessing potential causes and the consequences of these differences.
 8           Published multi-city analyses using common  data bases, measurement devices, analytical
 9           strategies, and extensive independent external review, as carried out in the APHEA and
10           NMMAPS  studies are likely to be useful. Pooled analyses of more diverse collections of
11           independent studies of different cities, using varying methodology and/or data quality or
12           representativeness, are likely less credible and should not, in general, be used without
13           careful assessment of their underlying scientific comparability.
14      (13)  It may be possible that different PM size components or particles with different
15           composition or sources produce effects by different mechanisms manifested at different
16           lags, or that different preexisting conditions may lead to different delays between exposure
17           and effect.  Thus,  although maximum effect sizes for PM effects have often been reported
18           for 0-1 day lags, evidence is also beginning to suggest that more consideration should be
19           given to lags of several days. Also, if it is considered that all health effects occurring at
20           different lag days  are all  real effects, so that the risks for each lag day should be additive,
21           then higher overall risks  may exist that are higher than implied by maximum estimates for
22           any particular single or two-day lags. In that  case, multi-day averages or distributed lag
23           models should be  used.
24      (14)  Certain classes of ambient particles may be distinctly less toxic than others and may not
25           exert human health effects at typical ambient exposure concentrations or only under special
26           circumstances.  Coarse thoracic particles of crustal origin, for example, may be relatively
27           non-toxic under most circumstances compared to those of combustion origin such as wood
28           burning.  However, crustal particles may be sufficiently toxic to cause human health effects
29           under some conditions; resuspended crustal particles, for example, may carry toxic trace
30           elements and other components from previously  deposited fine PM, e.g., metals from
31           smelters (Phoenix) or steel mills (Steubenville, Utah Valley), PAH's from automobile

        April 2002                                 8-287       DRAFT-DO NOT QUOTE OR CITE

-------
 1           exhaust, or pesticides from administration to agricultural lands. Likewise, fine particles
 2           from different sources have different effect sizes.  More research is needed to identify
 3           conditions under which one or another class of particles may cause little or no adverse
 4           health effects, as well as conditions under which particles may cause notable effects.
 5      (15)  Certain epidemiology evidence suggests that reducing ambient PM10 concentrations may
 6           reduce a variety of health effects on a time scale from a few days to a few months.  This has
 7           been found in epidemiology studies of "natural experiments" such as in the Utah Valley,
 8           and by supporting toxicology studies using the particles from ambient community sampling
 9           filters from the Utah Valley. Recent studies in Germany and in the Czech Republic also
10           tend to support a hypothesis that reductions in air pollution are associated with reductions
11           in the incidence of adverse health effects, but these studies cannot unambiguously attribute
12           improved health to reduced PM alone.
13      (16)  Adverse health effects in children are emerging as a more important area of concern than in
14           the  1996 PM AQCD. Unfortunately, relatively little is known about the relationship of PM
15           to the most serious health endpoints (low birth weight, preterm birth, neonatal and infant
16           mortality, emergency hospital admissions  and mortality in older children).
17      (17)  Little is yet known about involvement of PM  exposure in the progression from less serious
18           childhood conditions, such as asthma and  respiratory symptoms, to more serious disease
19           endpoints later in life.  This is an important health issue because childhood illness or death
20           may cost a very large number of productive life-years. Lastly, new epidemiologic studies
21           of ambient PM associations with increased non-hospital medical visits (physician visits)
22           and asthma effects suggest likely much larger health impacts and costs to society due to
23           ambient PM than just those indexed by mortality and/or hospital admissions/visits.
24
25
26
        April 2002                                 8-288       DRAFT-DO NOT QUOTE OR CITE

-------
REFERENCES

Abbey, D. E.; Mills, P. K.; Petersen, F. F.; Beeson, W. L. (1991) Long-term ambient concentrations of total
      suspended particulates and oxidants as related to incidence of chronic disease in California Seventh-Day
      Adventists. Environ. Health Perspect. 94: 43-50.
Abbey, D. E.; Lebowitz, M. D.; Mills, P. K.; Petersen, F. F.; Beeson, W. L.; Burchette, R. J. (1995a) Long-term
      ambient concentrations of particulates and oxidants and development of chronic disease in a cohort of
      nonsmoking California residents. In: Phalen, R. F.; Bates, D. V., eds. Proceedings of the colloquium on
      paniculate air pollution and human mortality and morbidity; January 1994; Irvine, CA.  Inhalation Toxicol.
      7: 19-34.
Abbey, D. E.; Ostro, B. E.; Fraser, G.; Vancuren, T.; Burchette, R. J. (1995b) Estimating fine  particulates less that
      2.5 microns in aerodynamic diameter (PM2 5) from airport visibility data in California. J. Exposure Anal.
      Environ. Epidemiol. 5: 161-180.
Abbey, D. E.; Burchette, R. J.; Knutsen, S. F.; McDonnell, W. F.; Lebowitz, M. D.; Enright, P. L. (1998) Long-term
      paniculate and other air pollutants and lung function in nonsmokers. Am. J. Respir. Crit. Care Med.
      158: 289-298.
Abbey, D. E.; Nishino, N.; McDonnell, W. F.; Burchette, R. J.; Knutsen, S. F.; Beeson, W. L.; Yang, J. X. (1999)
      Long-term inhalable particles and other air pollutants related to mortality in nonsmokers. Am. J. Respir. Crit.
      Care Med.  159: 373-382.
Ackermann-Liebrich, U.; Leuenberger, P.; Schwartz, J.; Schindler, C.; Monn, C.; Bolognini, B.; Bongard,  J.  P.;
      Brandli, O.; Domenighetti, G.; Elsasser, S.; Grize, L.; Karrer, W.; Keller, R.; Keller-Wossidlo, H.;
      Ktinzli, N.; Martin, B.  W.; Medici, T. C.; Perruchoud, A. P.; Schoni, M. H.; Tschopp, J. M.; Villiger,  B.;
      Wuthrich, B.; Zellweger, J. P.; Zemp, E. (1997) Lung function and long term exposure  to air pollutants in
      Switzerland. Am. J. Respir. Crit. Care Med. 155: 122-129.
Alberdi Odriozola, J. C.; Diaz Jimenez, J.; Montero Rubio, J. C.; Miron Perez, I. J.; Pajares Ortiz, M. S.; Ribera
      Rodrigues,  P. (1998) Air pollution and mortality in Madrid, Spain: a time-series analysis. Int. Arch. Occup.
      Environ. Health 71: 543-549.
Anderson, H. R.; Ponce de Leon, A.; Bland, J. M.; Bower, J. S.;  Strachan, D. P.  (1996) Air pollution and daily
      mortality in London: 1987-92. Br. Med. J. 312: 665-669.
Anderson, H. R.; Spix, C.; Medina, S.; Schouten, J. P.;  Castellsague, J.; Rossi, G.; Zmirou, D.; Touloumi,  G.;
      Wojtyniak, B.; Ponka,  A.; Bacharova, L.; Schwartz, J.; Katsouyanni, K. (1997) Air pollution and daily
      admissions for chronic obstructive pulmonary disease in 6 European cities: results from the APHEA project.
      Eur. Respir. J. 10: 1064-1071.
Anderson, H. R.; Ponce de Leon, A.; Bland, J. M.; Bower, J. S.; Emberlin, J.; Strachen, D. P.  (1998) Air pollution,
      pollens, and daily admissions for asthma in London 1987-92. Thorax 53: 842-848.
Anderson, H. R.; Bremner, S. A.; Atkinson, R. W.; Harrison, R. M.; Walters, S.  (2001) Paniculate matter and daily
      mortality and hospital admissions in the west midlands conurbation of the United Kingdom: associations with
      fine and coarse particles, black smoke and sulphate. Occup. Environ. Med. 58: 504-510.
Archer, V. E. (1990) Air pollution and fatal lung disease in three Utah counties.  Arch. Environ. Health 45: 325-334.
Atkinson, R. W.; Bremner, S. A.; Anderson, H. R.; Strachan, D. P.; Bland, J. M.; Ponce de Leon, A. (1999a)
      Short-term  associations between emergency hospital admissions for respiratory and cardiovascular disease
      and outdoor air pollution in London. Arch. Environ. Health 54: 398-411.
Atkinson, R. W.; Anderson, H. R.; Strachan, D. P.; Bland, J. M.; Bremner, S. A.; Ponce de Leon, A. (1999b)
      Short-term  associations between outdoor air pollution and visits to accident and emergency departments in
      London for respiratory complaints. Eur. Respir. J. 13: 257-265.
Atkinson, R. W.; Anderson, H. R.; Sunyer, J.;  Ayres, J.; Baccini, M.; Vonk, J. M.; Boumghar, A.; Forastiere, F.;
      Forsberg, B.; Touloumi, G.; Schwartz, J.; Katsouyanni, K. (2001) Acute effects of paniculate air pollution on
      respiratory  admissions: results from APHEA 2 project. Am. J. Respir. Crit. Care Med.  164: 1860-1866.
Avol, E. L.; Gauderman, W. J.; Tan, S. M.; London, S.  J.; Peters, J. M. (2001) Respiratory effects of relocating to
      areas of differing air pollution levels. Am. J. Respir. Crit.  Care Med. 164: 2067-2072.
Awasthi, S.; Click, H. A.; Fletcher, R. H.; Ahmed, N. (1996) Ambient air pollution & respiratory symptoms
      complex in preschool children. Indian J. Med. Res. 104: 257-262.
Bailey, D. L. R.; Clayton,  P. (1982) The measurement of suspended particle and total carbon concentrations  in the
      atmosphere using standard smoke shade methods. Atmos. Environ. 16: 2683-2690.
April 2002                                      8-289        DRAFT-DO NOT QUOTE OR CITE

-------
Baldi, I.; Tessier, J. F.; Kauffmann, F.; Jacqmin-Gadda, H.; Nejjari, C.; Salamon, R. (1999) Prevalence of asthma
      and mean levels of air pollution: results from the French PAARC survey. Eur. Respir. J. 14: 132-138.
Barbone, F.; Bovenzi, M; Cavalleri, F.; Stanta, G. (1995) Air pollution and lung cancer in Trieste, Italy. Am. J.
      Epidemiol.  141: 1161-1169.
Basu, R.; Samet, J. M. (1999) A review of the epidemiological evidence on health effects of nitrogen dioxide
      exposure from gas stoves. J. Environ. Med. 1: 173-187.
Bateson, T. F.; Schwartz, J. (1999) Control for seasonal variation and time trend in case-crossover studies of acute
      effects of environmental exposures. Epidemiology 10: 539-544.
Beckett, W. S. (2001) The air pollution detectives. Am. J. Respir. Crit. Care Med.  164: 515-516.
Beeson, W. L.; Abbey, D. E.; Knutsen, S. F. (1998) Long-term concentrations of ambient air pollutants and incident
      lung cancer in California adults: results from the AHSMOG study. Environ. Health Perspect. 106: 813-823.
Berglund, D. J.; Abbey, D. E.; Lebowitz, M. D.; Knutsen, S. F.; McDonnell, W. F. (1999) Respiratory symptoms
      and pulmonary function in an elderly nonsmoking population. Chest 115: 49-59.
Beyer, U.; Franke, K.; Cyrys, J.; Peters, A.; Heinrich, J.; Wichmann, H. E.; Brunekreef, B. (1998) Air pollution and
      respiratory  health of children: the PEACE panel study in Hettstedt and Zerbst, Eastern Germany. Eur. Respir.
      Rev. 8: 61-69.
Blanchard, C. L.; Lurmann, F. W.;  Roth, P. M.; Jeffries, H. E.; Korc, M. (1999) The use of ambient data to
      corroborate analyses of ozone control strategies. Atmos. Environ. 33: 369-381.
Bobak, M.; Leon,  D. A. (1992) Air pollution and infant mortality in the Czech Republic, 1986-1988.
      Lancet (8826): 1010-1014.
Bobak, M.; Leon,  D. (1998)  Air pollution and infant mortality: the effects are specific for respiratory causes in
      postneonatal period. Epidemiology 9: S58.
Bobak, M.; Leon,  D. A. (1999) Pregnancy outcomes and outdoor air pollution: an ecological study in districts of the
      Czech Republic 1986-8. Occup. Environ. Med. 56: 539-543.
Bobak, M.; Roberts, A. (1997) Heterogeneity of air pollution effects is related to average temperature [letter].
      Br. Med. J. 315:  1161.
Boezen, M.; Schouten, J.; Rijcken, B.; Vonk, J.; Gerritsen, J.; VanDerZee, S.; Hoek, G.; Brunekreef, B.;
      Postma, D.  (1998) Peak expiratory flow variability, bronchial responsiveness, and susceptibility to ambient
      air pollution in adults. Am. J. Respir. Crit. Care Med. 158: 1848-1854.
Boezen, H. M.; VanDerZee, S. C.; Postma, D. S.; Vonk, J. M.; Gerritsen, J.; Hoek, G.; Brunekreef, B.;
      Rijcken, B.; Schouten, J. P.  (1999) Effects of ambient air pollution on upper and lower respiratory symptoms
      and peak expiratory flow in  children. Lancet 353: 874-878.
Borja-Aburto, V H.; Loomis, D. P.; Bangdiwala, S. I.; Shy, C. M.; Rascon-Pacheco, R. A. (1997) Ozone,
      suspended particulates, and daily mortality in Mexico City. Am. J. Epidemiol. 145: 258-268.
Borja-Aburto, V H.; Castillejos, M.; Gold, D. R.; Bierzwinski, S.; Loomis, D. (1998) Mortality and ambient fine
      particles in southwest Mexico City, 1993-1995. Environ. Health Perspect. 106: 849-855.
Braga, A. L. F.; Conceicao, G. M. S.; Pereira, L. A. A.; Kishi, H.  S.; Pereira, J. C.  R.; Andrade, M. F.; Goncalves,
      F. L. T.; Saldiva, P. H. N.; Latorre, M. R. D. O. (1999) Air pollution and pediatric respiratory hospital
      admissions in Sao Paulo, Brazil. J. Environ. Med. 1: 95-102.
Braga, A. L. F.; Zanobetti, A.; Schwartz, J. (2000) Do respiratory epidemics confound the association between air
      pollution and daily deaths? Eur. Respir. J. 16: 723-728.
Braga, A. L. F.; Zanobetti, A.; Schwartz, J. (2001) The lag structure between paniculate air pollution and
      respiratory  and cardiovascular deaths in ten U.S. cities. J. Occup. Environ. Med. 43: 927-933.
Brauer, M.; Ebelt, S. T.; Fisher, T.  V; Brumm,  J.; Petkau, A. J.; Vedal, S. (2001) Exposure of chronic obstructive
      pulmonary  disease patients to particles: respiratory and cardiovascular health effects. J. Exposure Anal.
      Environ. Epidemiol. 11:  490-500.
Braun-Fahrlander, C.; Vuille, J. C.; Sennhauser, F. H.; Neu, U.; Ktinzle, T.; Grize, L.; Gassner, M.; Minder, C.;
      Schindler, C.; Varonier, H. S.; Wuthrich, B.; SCARPOL team. (1997) Respiratory health and long-term
      exposure to air pollutants in Swiss schoolchildren. Am. J. Respir. Crit. Care Med.  155: 1042-1049.
Bremner, S. A.; Anderson, H. R.; Atkinson, R. W.; McMichael, A. J.; Strachan, D. P.; Bland, J. M.; Bower, J. S.
      (1999) Short term associations between outdoor air pollution and mortality in London 1992-4. Occup.
      Environ. Med. 56: 237-244.
Brook, J. R.; Wiebe, A. H.; Woodhouse, S. A.; Audette, C. V; Dann, T. F.; Callaghan, S.; Piechowski, M.;
      Dabek-Zlotorzynska, E.; Dloughy, J. F. (1997) Temporal and spatial relationships in fine particle strong
      acidity, sulphate, PM10, andPM25 across multiple Canadian locations. Atmos. Environ. 31: 4223-4236.
Brunekreef, B. (1997) Air pollution and life expectancy: is there a relation? Occup. Environ. Med. 54: 781-784.

April 2002                                      8-290        DRAFT-DO NOT QUOTE OR CITE

-------
Buffler, P. A.; Cooper, S. P.; Stinnett, S.; Contant, C.; Shirts, S.; Hardy, R. J.; Agu, V.; Gehan, B.; Burau, K. (1988)
      Air pollution and lung cancer mortality in Harris County, Texas, 1979-1981. Am. J. Epidemiol.
      128: 683-699.
Burnett, R.  T.; Dales, R. E.; Raizenne, M. E.; Krewski, D.; Summers, P. W.; Roberts, G. R.; Raad-Young, M;
      Dann, T.; Brook, J. (1994) Effects of low ambient levels of ozone and sulfates on the frequency of
      respiratory admissions to Ontario hospitals. Environ. Res. 65: 172-194.
Burnett, R.  T.; Dales, R.; Krewski, D.; Vincent, R.; Dann, T.; Brook, J. R. (1995) Associations between ambient
      paniculate sulfate and admissions to Ontario hospitals for cardiac  and respiratory diseases. Am. J. Epidemiol.
      142: 15-22.
Burnett, R.  T.; Cakmak, S.; Brook, J. R.; Krewski, D. (1997a) The role of paniculate size and chemistry in the
      association between summertime ambient air pollution and hospitalization for cardiorespiratory diseases.
      Environ. Health Perspect. 105: 614-620.
Burnett, R.  T.; Brook, J. R.; Yung, W. T.; Dales, R. E.; Krewski, D.  (1997b) Association between ozone and
      hospitalization for respiratory diseases in 16 Canadian cities. Environ. Res. 72: 24-31.
Burnett, R.  T.; Dales, R. E.; Brook, J. R.; Raizenne, M. E.; Krewski, D. (1997c) Association between ambient
      carbon monoxide levels and hospitalizations for congestive heart failure in the elderly in 10 Canadian cities.
      Epidemiology  8: 162-167.
Burnett, R.  T.; Cakmak, S.; Brook, J. R. (1998a) The effect of the urban ambient air pollution mix on daily mortality
      rates in 11 Canadian cities. Can. J. Public Health 89: 152-156.
Burnett, R.  T.; Cakmak, S.; Raizenne, M. E.; Stieb, D.; Vincent, R.; Krewski, D.; Brook, J. R.;  Philips, O.;
      Ozkaynak, H. (1998b) The association between ambient carbon monoxide levels and daily mortality in
      Toronto, Canada. J. Air Waste Manage. Assoc. 48: 689-700.
Burnett, R.  T.; Smith-Doiron, M.; Stieb, D.; Cakmak, S.;  Brook, J. R. (1999) Effects of paniculate and gaseous air
      pollution on cardiorespiratory hospitalizations. Arch. Environ. Health 54:130-139.
Burnett, R.  T.; Brook, J.; Dann, T.; Delocla, C.; Philips, O.; Cakmak, S.;  Vincent, R.; Goldberg, M. S.; Krewski, D.
      (2000) Association between paniculate- and gas-phase components of urban air pollution and daily mortality
      in eight Canadian cities. In: Grant, L. D., ed. PM2000: paniculate  matter and health. Inhalation Toxicol.
      12(suppl. 4): 15-39.
Burnett, R.; Ma, R.; Jerrett, M.; Goldberg, M. S.; Cakmak, S.; Pope, C. A., Ill; Krewski, D. (2001a) The spatial
      association between community air pollution and mortality: a new method of analyzing correlated geographic
      cohort data. Environ. Health Perspect. 109(suppl. 3): 375-380.
Burnett, R.  T.; Smith-Doiron, M.; Stieb, D.; Raizenne, M. E.; Brook, J. R.; Dales, R. E.; Leech, J. A.; Cakmak, S.;
      Krewski, D.  (200 Ib) Association between ozone and hospitalization for acute respiratory diseases in children
      less than 2 years of age. Am. J. Epidemiol. 153: 444-452.
Burton, R. M.; Suh, H. H.; Koutrakis, P. (1996) Spatial variation in paniculate concentrations within metropolitan
      Philadelphia. Environ. Sci. Technol. 30: 400-407.
Cakmak, S.; Burnett,  R. T.; Krewski, D. (1999) Methods  for detecting and estimating population threshold
      concentrations for air pollution-related mortality with exposure  measurement error. Risk Anal. 19: 487-496.
Calderon-Garciduenas, L.; Mora-Tiscareno, A.; Chung, C. J.; Valencia, G.; Fordham, L. A.; Garcia, R.;  Osnaya, N.;
      Romero, L.;  Acuna, H.; Villarreal-Calderon, A.; Devlin, R. B.; Koren, H.  S. (2000) Exposure to air pollution
      is associated with lung hyperinflation in healthy children and adolescents in southwest Mexico City: a pilot
      study. Inhalation Toxicol. 12: 537-561.
Cassino, C.; Ito, K.; Bader, L; Ciotoli, C.; Thurston, G.; Reibman, J. (1999) Cigarette smoking  and ozone-associated
      emergency department use for asthma by adults in New York City. Am. J. Respir. Crit. Care Med.
      159: 1773-1779.
Castillejos, M.; Borja-Aburto, V. H.; Dockery, D. W.; Gold, D. R.; Loomis, D. (2000) Airborne coarse particles and
      mortality. In: Inhalation Toxicology: proceedings of the third colloquium on paniculate air pollution and
      human health;  June, 1999; Durham, NC. Inhalation Toxicology 12(suppl.  1): 61-72.
Chang, L.-T.; Koutrakis, P.; Catalano, P. J.; Suh, H. H. (2000) Hourly  personal exposures to fine particles and
      gaseous pollutants—results from Baltimore, Maryland. J. Air Waste Manage. Assoc. 50: 1223-1235.
Chay, K. Y.; Greenstone, M. (2001) The impact of air pollution on infant mortality: evidence from geographic
      variation in pollution shocks induced by a recession. Berkeley, CA: University of California, Berkeley,
      Department of Economics. Available:  http://elsa.berkeley.edu/users/kenchay/ [29 April,  2002].
Chen, C.; Chock, D. P.; Winkler, S. L. (1999) A simulation study of confounding in generalized linear models  for
      air pollution epidemiology. Environ. Health Perspect. 107: 217-222.


April 2002                                      8-291         DRAFT-DO NOT QUOTE OR CITE

-------
Chen, L.; Yang, W.; Jennison, B. L.; Omaye, S. T. (2000) Air paniculate pollution and hospital admissions for
      chronic obstructive pulmonary disease in Reno, Nevada. Inhalation Toxicol. 12: 281-298.
Chen, L.; Yang, W.; Jennison, B. L.; Goodrich, A.; Omaye, S. T. (2002) Air pollution and birth weight in northern
      Nevada, 1991-1999. Inhalation Toxicol. 14: 141-157.
Chew, F. T.; Goh, D. Y. T.; Ooi, B. C.; Saharom, R.; Hui, J. K. S.; Lee, B. W. (1999) Association of ambient
      air-pollution levels with acute asthma exacerbation among children in Singapore. Allergy (Copenhagen)
      54: 320-329.
Chock, D. P.; Winkler, S.; Chen, C. (2000) A study of the association between daily mortality and ambient air
      pollutant concentrations in Pittsburgh, Pennsylvania. J. Air Waste Manage. Assoc. 50: 1481-1500.
Choudhury, A. H.; Gordian, M. E.; Morris, S.  S. (1997) Associations between respiratory illness and PM10 air
      pollution. Arch. Environ. Health 52: 113-117.
Cifuentes, L. A.; Vega, J.; Kopfer, K.; Lave, L. B. (2000) Effect of the fine fraction of paniculate matter versus the
      coarse mass and other pollutants on daily mortality in Santiago, Chile. J. Air Waste Manage. Assoc.
      50: 1287-1298.
Clayton, C. A.; Perritt, R. L.; Pellizzari, E. D.; Thomas, K. W.; Whitmore, R. W.; Wallace, L. A.; Ozkaynak, H.;
      Spengler, J. D. (1993) Particle total exposure assessment methodology (PTEAM) study: distributions of
      aerosol and elemental concentrations in personal, indoor, and outdoor air samples in a southern California
      community. J. Exposure Anal. Environ. Epidemiol. 3: 227-250.
Clench-Aas, J.; Bartonova, A.; Skjensberg, O. H.; Leegaard, J.; Hagen, L. O.;  Gisver, P.; Moseng, J.; Roemer, W.
      (1998) Air pollution and respiratory health of children: the PEACE study in Oslo, Norway. Eur. Respir. Rev.
      8: 36-43.
Clyde, M. A.; Guttorp, P.; Sullivan, E. (2000) Effects of ambient fine and coarse particles on mortality in Phoenix,
      Arizona.  J.  Exposure Anal. Environ. Epidemiol.: submitted.
Cohen, A. J. (2000)  Outdoor air pollution and lung cancer. Environ. Health Perspect. 108(suppl. 4): 743-750.
Cohen, A. J.; Pope, C. A., III.  (1995) Lung cancer and air pollution.  Environ. Health Perspect.
      103(suppl.  8): 219-224.
Conceicao, G. M.  S.; Miraglia, S. G. E. K.; Kishi, H. S.; Saldiva, P. H. N.; Singer, J. M. (2001) Air pollution and
      child mortality: a time-series study in Sao Paulo, Brazil. Environ. Health Perspect. 109(suppl. 3): 347-350.
Coull, B. A.; Schwartz, J.; Wand, M. P. (2001) Respiratory health and air pollution: additive mixed model analyses.
      Biostatistics 2: 337-349.
Coutant, R. W. (1977) Effect of environmental variables on collection of atmospheric sulfate. Environ. Sci.
      Technol.  11:873-878.
Creason, J.; Neas, L.; Walsh, D.; Williams, R.; Sheldon, L.; Liao, D.; Shy, C. (2001) Paniculate matter and heart
      rate variability among elderly retirees: the Baltimore  1998 PM study. J.  Exposure Anal. Environ. Epidemiol.
      11: 116-122.
Cropper, M. L.; Simon, N. B.;  Alberini, A.; Arora, S.; Sharma, P. K. (1997) The  health benefits of air pollution
      control in Delhi. Am. J. Agric. Econ. 79: 1625-1629.
Cuijpers, C. E. J.;  Swaen, G. M. H.; Wesseling, G.; Wouters, E. F. M. (1994) Acute respiratory effects of summer
      smog in primary school children. Toxicol. Lett. 72: 227-235.
Dab, W.; Medina, S.; Quenel,  P.; Le Moullec, Y.; Le Tertre, A.; Thelot, B.; Monteil, C.; Lameloise, P.; Pirard, P.;
      Momas, L;  Ferry, R.; Festy, B. (1996) Short term respiratory health effects of ambient air pollution: results of
      the APHEA project in Paris. In: St Leger, S., ed. The APHEA project. Short term effects of air pollution on
      health: a European approach using epidemiological time series data. J. Epidemiol. Community Health
      50 (suppl. 1): S42-S46.
Damia, A. D.; Fabregas, M. L.; Tordera, M. P.; Torrero, L.  C. (1999) Effects of air pollution and weather conditions
      on asthma exacerbation. Respiration 66: 52-58.
Daniels, M.; Dominici, F.; Samet, J. M.; Zeger, S. L. (2000) Estimating paniculate matter-mortality dose-response
      curves and threshold levels: an analysis of daily time-series for the 20 largest US cities. Am. J. Epidemiol.
      152: 397-406.
Dejmek, J.; Selevan, S. G.; Benes, I.; Solansky, I.; Sram, R. J. (1999) Fetal growth and maternal exposure to
      paniculate matter during pregnancy. Environ. Health Perspect. 107: 475-480.
Delfino, R. J.; Coate, B. D.; Zeiger, R. S.; Seltzer, J. M.; Street, D. H.; Koutrakis, P. (1996) Daily asthma severity in
      relation to personal ozone exposure and outdoor fungal spores. Am. J. Respir. Crit. Care Med. 154: 633-641.
Delfino, R. J.; Murphy-Moulton, A. M.; Burnett, R. T.; Brook, J. R.; Becklake, M. R. (1997) Effects of air pollution
      on emergency room visits for respiratory illnesses in Montreal, Quebec. Am. J. Respir. Crit. Care Med.
      155: 568-576.

April 2002                                      8-292         DRAFT-DO NOT QUOTE OR CITE

-------
Delfino, R. J.; Murphy-Moulton, A. M.; Becklake, M. R. (1998) Emergency room visits for respiratory illnesses
      among the elderly in Montreal: association with low level ozone exposure. Environ. Res. 76: 61-11.
Dewanji, A.; Moolgavkar, S. H. (2000) A Poisson process approach for recurrent event data with environmental
      covariates. Environmetrics 11: 665-673.
Diaz, J.; Garcia, R.; Ribera, P.; Alberdi, J. C.; Hernandez, E.; Pajares, M. S.; Otero, A. (1999) Modeling of air
      pollution and its relationship with mortality and morbidity in Madrid, Spain. Int. Arch. Occup. Environ.
      Health 72: 366-376.
Dockery, D. W.; Spengler, J. D.; Neas, L. M.; Speizer, F. E.; Ferris, B. G., Jr.; Ware, J. H.; Brunekreef, B. (1989)
      An epidemiologic study of respiratory health status and indicators of indoor air pollution from combustion
      sources. In: Harper, J. P., ed. Combustion processes and the quality of the indoor environment: transactions
      of an international specialty conference; September 1988; Niagara Falls, NY. Pittsburgh, PA: Air & Waste
      Management Association; pp. 262-271. (A&WMA transactions series: TR-15).
Dockery, D. W.; Schwartz, J.; Spengler, J. D. (1992) Air pollution and daily mortality: associations with particulates
      and acid aerosols. Environ. Res. 59: 362-373.
Dockery, D. W.; Pope, C. A., Ill; Xu, X.; Spengler, J. D.; Ware, J. H.; Fay, M. E.; Ferris, B. G., Jr.; Speizer, F. E.
      (1993) An association between air pollution and mortality in six U.S. cities. N. Engl. J. Med.
      329: 1753-1759.
Dockery, D. W.; Cunningham, J.; Damokosh, A. L; Neas, L. M.; Spengler, J. D.; Koutrakis, P.; Ware, J. H.;
      Raizenne, M.; Speizer, F. E. (1996) Health effects of acid aerosols on North American children: respiratory
      symptoms. Environ. Health Perspect. 104: 500-505.
Dockery, D. W.; Pope, C. A., Ill; Kanner, R. E.; Villegas, G. M.; Schwartz, J. (1999) Daily changes in oxygen
      saturation and pulse rate associated with paniculate air pollution and barometric pressure. Cambridge, MA:
      Health Effects Institute; research report no. 83.
Dominici, F.; Samet, J. M.; Zeger, S. (2000a) Combining evidence on air pollution and daily mortality from the
      largest 20 US cities: a hierarchical modeling strategy. J. R. Stat. Soc. A 163: 263-302.
Dominici, F.; Zeger, S. L.; Samet, J. (2000b) A measurement error model for time-series studies of air pollution and
      mortality. Biostatistics 1: 157-175.
Dominici, F.; Daniels, M.; Zeger, S. L.; Samet, J. M. (2002) Air pollution and mortality: estimating regional and
      national dose-response relationships. J. Am. Stat. Assoc.  97: 100-111.
Dye, J. A.; Lehmann, J. R.; McGee, J. K.; Winsett, D. W.; Ledbetter, A. D.; Everitt, J. L; Ohio, A. J.; Costa, D. L.
      (2001) Acute pulmonary toxicity of paniculate matter filter extracts in rats: coherence with epidemiological
      studies in Utah Valley residents. Environ. Health Perspect. 109(suppl. 3): 395-403.
Fairley, D. (1990) The relationship of daily mortality to  suspended particulates in Santa Clara county, 1980-86.
      Environ. Health Perspect. 89:  159-168.
Fairley, D. (1999) Daily mortality and air pollution in Santa Clara County, California: 1989-1996. Environ. Health
      Perspect. 107: 637-641.
Forsberg, B.;  Segerstedt, B.; Stjernberg, N.; Roemer, W. (1998) Air pollution and respiratory health of children:
      the PEACE panel study in Umea, Sweden. Eur. Respir. Rev. 8: 12-19.
Frampton, M. W.; Ohio, A. J.; Samet, J. M.; Carson, J. L.; Carter, J. D.; Devlin, R. B. (1999) Effects of aqueous
      extracts of PM10 filters from the Utah Valley on human airway epithelial cells. Am. J. Physiol.
      277: L960-L967.
Friedman, M. S.; Powell, K. E.; Hutwagner, L.; Graham, L. M.; Teague, W. G.  (2001) Impact of changes in
      transportation and commuting behaviors during the 1996 summer Olympic games in Atlanta on air quality  and
      childhood asthma. JAMA J. Am. Med. Assoc. 285: 897-905.
Frischer, T.; Studnicka, M.; Gartner, C.; Tauber, E.; Horak, F.; Veiter, A.; Spengler, J.; Ktihr, J.; Urbanek, R.
      (1999) Lung function growth and ambient ozone: a three-year population study in school children. Am. J.
      Respir. Crit. Care Med. 160: 390-396.
Fusco, D.; Forastiere, F.; Michelozzi, P.; Spadea, T.; Ostro, B.;  Area, M.; Perucci, C. A. (2001) Air pollution and
      hospital admissions for respiratory conditions in Rome, Italy. Eur. Respir. J. 17: 1143-1150.
Gamble, J. L. (1998) Effects of ambient air pollution on daily mortality: a time series analysis of Dallas, Texas,
      1990-1994. Presented at: 91st annual meeting and exhibition of the Air & Waste Management Association;
      June; San Diego, CA. Pittsburgh, PA: Air &  Waste Management Association; paper no. 98-MP26.03.
Garcia-Aymerich, J.; Tobias, A.; Anto, J. M.; Sunyer, J. (2000)  Air pollution and mortality in a cohort of patients
      with chronic obstructive pulmonary disease: a time series analysis. J. Epidemiol. Community Health
      54: 73-74.


April 2002                                      8-293         DRAFT-DO NOT QUOTE OR CITE

-------
Garty, B. Z.; Kosman, E.; Ganor, E.; Berger, V.; Garty, L.; Wietzen, T.; Waisman, Y.; Mimouni, M.; Waisel, Y.
      (1998) Emergency room visits of asthmatic children, relation to air pollution, weather, and airborne allergens.
      Ann. Allergy Asthma Immunol. 81: 563-570.
Gauderman, W. J.; Mcconnell, R.; Gilliland, F.; London, S.; Thomas, D.; Avol, E.; Vora, H.; Berhane, K.;
      Rappaport, E. B.; Lurmann, F.; Margolis, H. G.; Peters, J. (2000) Association between air pollution and lung
      function growth in southern California children. Am. J. Respir. Crit. Care Med. 162: 1383-1390.
Gauvin, S.; Zmirou, D.; Pin, L; Quentin, J.; Balducci, F.; Boudet, C.; Poizeau, D.; Brambilla, C. (1999) Short-term
      effect of exposure to suspended paniculate matter (PM10) on the respiratory function of urban asthmatic and
      control adults. J. Environ. Med. 1: 71-79.
Gherghinova, M.; Kostyanve,  S.; Ivanova, M. (1989) Theoretic values of body plethysmography indices in healthy
      children between 7 and 14 years of age. Pediatria 28: 52-58.
Ohio, A. J.; Devlin, R. B. (2001) Inflammatory lung injury after bronchial instillation of air pollution particles.
      Am. J. Respir. Crit. Care Med. 164: 704-708.
Gielen, M. H.; VanDer Zee, S. C.; Van Wijnen, J. H.; Van Steen, C. J.; Brunekreef, B.  (1997) Acute effects of
      summer air pollution on respiratory health of asthmatic children. Am. J.  Respir. Crit. Care Med.
      155:2105-2108.
Gold, D. R.; Damokosh, A. L; Pope, C. A., Ill; Dockery, D. W.; McDonnell, W. F.; Serrano, P.; Retama, A.;
      Castillejos, M. (1999) Paniculate and ozone pollutant effects on the respiratory function of children in
      southwest Mexico City. Epidemiology 10: 8-16.
Gold, D. R.; Litonjua, A.; Schwartz, J.; Lovett, E.; Larson, A.; Nearing, B.; Allen, G.; Verrier, M.; Cherry, R.;
      Verrier, R. (2000) Ambient pollution and heart rate variability. Circulation 101: 1267-1273.
Goldberg, M. S.; Bailar, J. C., Ill; Burnett, R. T.; Brook, J. R.; Tamblyn, R.; Bonvalot, Y.; Ernst, P.; Flegel, K. M.;
      Singh,  R. K.; Valois, M.-F. (2000) Identifying subgroups of the general population that may be susceptible to
      short-term increases in paniculate air pollution: a time-series study in Montreal, Quebec. Cambridge, MA:
      Health Effects Institute; research report 97. Available: http://www.healtheffects.org/pubs-research.htm
      [15 February, 2001].
Goldberg, M. S.; Burnett, R. T.; Bailar, J. C., Ill; Brook, J.; Bonvalot, Y.; Tamblyn, R.;  Singh, R.; Valois, M.-F.
      (200 la) The association between daily mortality and ambient air particle pollution in Montreal, Quebec. 1.
      Nonaccidental mortality. Environ. Res. 86: 12-25.
Goldberg, M. S.; Burnett, R. T.; Bailar, J. C., Ill; Brook, J.; Bonvalot, Y.; Tamblyn, R.;  Singh, R.; Valois, M.-F.;
      Vincent, R. (200 Ib) The association between daily mortality and ambient air particle pollution in Montreal,
      Quebec. 2. Cause-specific mortality. Environ. Res. 86: 26-36.
Goldberg, M. S.; Burnett, R. T.; Bailar, J. C., Ill; Tamblyn, R.; Ernst, P.; Flegel, K.; Brook, J.; Bonvalot, Y.;
      Singh,  R.; Valois, M.-F.; Vincent, R. (200Ic) Identification of persons with cardiorespiratory conditions who
      are at risk of dying from the acute effects of ambient air particles. Environ. Health Perspect.
      109(suppl. 4): 487-494.
Goldberg, M. S.; Burnett, R. T.; Brook, J.; Bailar, J. C., Ill; Valois, M.-F.; Vincent, R. (2001d) Associations
      between daily cause-specific mortality and concentrations of ground-level ozone  in Montreal, Quebec. Am. J.
      Epidemiol. 154: 817-826.
Gordian, M. E.; Ozkaynak, H.; Xue, J.; Morris,  S. S.; Spengler, J. D. (1996) Paniculate  air pollution and respiratory
      disease in Anchorage, Alaska. Environ. Health Perspect. 104: 290-297.
Gouveia, N.; Fletcher, T. (2000) Respiratory diseases in children and outdoor air pollution in Sao Paulo, Brazil:
      a time series analysis. Occup. Environ. Med. 57: 477-483.
Grievink, L.; Van der Zee, S. C.; Hoek, G.; Boezen, H. M.; Van't Veer, P.; Brunekreef, B. (1999) Modulation of the
      acute respiratory effects of winter air pollution by serum and dietary antioxidants: a panel study.  Eur. Respir.
      J. 13: 1439-1446.
Gtintzel, O.; Bollag, U.; Helfenstein, U. (1996) Asthma and exacerbation of chronic bronchitis: sentinel and
      environmental data in a time series analysis. Zentralbl. Hyg. Umweltmed. 198: 383-393.
Guo, Y. L.; Lin, Y.-C; Sung, F.-C; Huang, S.-L.; Ko, Y.-C; Lai, J.-S.; Su, H.-J.; Shaw, C.-K.; Lin, R.-S.; Dockery,
      D. W. (1999) Climate, traffic-related air pollutants, and asthma prevalence in middle-school children in
      Taiwan. Environ. Health Perspect. 107: 1001-1006.
Gwynn, R. C.; Thurston, G. D. (2001) The burden of air pollution: impacts among racial minorities. Environ. Health
      Perspect. 109(suppl. 4): 501-506.
Gwynn, R. C.; Burnett, R. T.; Thurston, G.  D. (2000) A time-series analysis of acidic paniculate matter and daily
      mortality and morbidity in the Buffalo, New York, region. Environ. Health Perspect. 108: 125-133.


April 2002                                       8-294         DRAFT-DO NOT QUOTE OR CITE

-------
Hagen, J. A.; Nafstad, P.; Skrondal, A.; Bjerkly, S.; Magnus, P. (2000) Associations between outdoor air pollutants
      and hospitalization for respiratory diseases. Epidemiology 11: 136-140.
Hajat, S.; Haines, A.; Goubet, S. A.; Atkinson, R. W.; Anderson, H. R. (1999) Association of air pollution with
      daily GP consultations for asthma and other lower respiratory conditions in London. Thorax 54: 597-605.
Hajat, S.; Haines, A.; Atkinson, R. W.; Bremner, S. A.; Anderson, H. R.; Emberlin, J. (2001) Association between
      air pollution and daily consultations with general practitioners for allergic rhinitis in London, United
      Kingdom. Am. J. Epidemiol. 153: 704-714.
Haluszka, J.; Pisiewicz, K.; Miczynski, J.; Roemer, W.; Tomalak, W. (1998) Air pollution and respiratory health in
      children: the PEACE panel study in Krakow, Poland. Eur. Respir. Rev. 8: 94-100.
Harre, E. S. M; Price, P. D.; Ayrey, R. B.; Toop, L. J.; Martin, I. R.; Town, G. I. (1997) Respiratory effects of air
      pollution in chronic obstructive pulmonary disease: a three month prospective  study. Thorax 52: 1040-1044.
Harris, R. J. (1975) A primer of multivariate statistics. New York, NY: Academic Press; pp. 155-224.
Hefflin, B. J.; Jalaludin, B.; McClure, E.; Cobb, N.; Johnson, C. A.; Jecha, L.; Etzel,  R. A. (1994) Surveillance for
      dust storms and respiratory diseases in Washington State,  1991. Arch. Environ. Health 49: 170-174.
Heinrich, J.; Hoelscher, B.; Wjst, M.; Ritz, B.; Cyrys, J.; Wichmann, H.-E. (1999) Respiratory diseases and
      allergies in two polluted areas in East Germany. Environ. Health Perspect. 107: 53-62.
Heinrich, J.; Hoelscher, B.; Wichmann, H. E. (2000) Decline of ambient air pollution and respiratory symptoms in
      children. Am. J. Respir. Crit. Care Med. 161:  1930-1936.
Henderson, B. E.; Gordon, R. J.; Menck, H.; SooHoo, J.; Martin, S. P.; Pike, M. C. (1975) Lung cancer and air
      pollution in southcentral Los Angeles County. Am. J. Epidemiol. 101:  477-488.
Hickey, R. J.; Boyce, D. E.; Harner, E. B.;  Clelland, R. C. (1970) Ecological statistical studies on environmental
      pollution and chronic disease in metropolitan areas of the United States. Philadelphia, PA: Regional Science
      Research Institute. (RSRI discussion paper series no. 35).
Hill, A. B. (1965) The environment and disease: association or causation? Proc. R. Soc. Med. 58: 295-300.
Hiltermann, T. J. N.; de Bruijne, C. R.; Stolk, J.; Zwinderman, A. H.; Spieksma, F. Th. M.; Roemer, W.;
      Steerenberg, P. A.; Fischer, P. H.; van Bree, L.; Hiemstra, P. S. (1997) Effects of photochemical air pollution
      and allergen exposure on upper respiratory tract inflammation in asthmatics. Am. J. Respir. Crit. Care Med.
      156: 1765-1772.
Hiltermann, T. J. N.; Stolk, J.; VanderZee, S. C.; Brunekreef, B.; De Bruijne, C. R.; Fischer, P. H.; Ameling,
      C. B.; Sterk, P. J.; Hiemstra, P.  S.; Van Bree, L.  (1998) Asthma severity and susceptibility to air pollution.
      Eur. Respir. J.  11: 686-693.
Hoek, G.; Schwartz, J. D.; Groot, B.; Eilers, P. (1997) Effects of ambient paniculate matter and ozone on daily
      mortality in Rotterdam, the Netherlands. Arch. Environ. Health 52: 455-463.
Hoek, G.; Dockery, D. W.; Pope, A.; Neas, L.; Roemer, W.; Brunekreef, B. (1998) Association between PM10 and
      decrements in peak expiratory flow rates in children: reanalysis of data from five panel studies. Eur. Respir.
      J. 11: 1307-1311.
Hoek, G.; Brunekreef, B.; Verhoeff, A.; van, Wijnen, J.; Fischer, P. (2000) Daily mortality and air pollution in the
      Netherlands. J. Air Waste Manage. Assoc. 50: 1380-1389.
Hoek, G.; Brunekreef, B.; Fischer, P.;  Van Wijnen, J. (2001) The association between air pollution and heart
      failure, arrhythmia, embolism, thrombosis, and other cardiovascular causes of death in a time series study.
      Epidemiology 12:  355-357.
Hodgkin, J. E.; Abbey, D. E.; Euler, G. L.; Magie, A. R. (1984) COPD prevalence in nonsmokers in high and low
      photochemical air pollution areas. Chest 86: 830-838.
Hong, Y.-C.; Leem, J.-H.; Ha, E.-H.; Christiani, D. C. (1999) PM10 exposure, gaseous pollutants, and daily
      mortality in Inchon, South Korea. Environ. Health Perspect. 107: 873-878.
Howel, D.; Darnell, R.; Pless-Mulloli, T. (2001) Children's respiratory health and daily paniculate levels in
      10 nonurban communities. Environ. Res. 87: 1-9.
Hruba, F.; Fabianova, E.; Koppova, K.; Vandenberg, J. J. (2001) Childhood respiratory symptoms, hospital
      admissions and long-term exposure to airborne paniculate matter. J. Exposure  Anal. Environ. Epidemiol.
      11:33-40.
Huang, Y.-L.; Batterman, S. (2000) Residence location as a measure of environmental exposure: a review of air
      pollution epidemiology studies. J. Exposure Anal. Environ. Epidemiol.  10: 66-85.
Ibald-Mulli, A.; Stieber, J.; Wichmann, H.-E.; Koenig, W.; Peters, A. (2001) Effects of air pollution on blood
      pressure: a population-based approach. Am. J. Public Health 91: 571-577.
April 2002                                     8-295         DRAFT-DO NOT QUOTE OR CITE

-------
Ilabaca, M; Olaeta, I.; Campos, E.; Villaire, J.; Tellez-Rojo, M. M; Romieu, I. (1999) Association between levels
      of fine paniculate and emergency visits for pneumonia and other respiratory illnesses among children in
      Santiago, Chile. J. Air Waste Manage. Assoc. 49: 154-163.
Ito, K. (1990) An examination of the role of aerosol acidity in historical London, England daily mortality
      [dissertation]. Syracuse, NY: New York University. Available from: University Microfilms International,
      Ann Arbor, MI; AAD91-13012.
Jacobs, J.; Kreutzer, R.; Smith, D. (1997) Rice burning and asthma hospitalizations, Butte County, California,
      1983-1992. Environ. HealthPerspect. 105: 980-985.
Jalaludin, B. B.; Chey, T.; O'Toole, B. I.; Smith, W. T.; Capon, A. G.; Leeder, S. R. (2000) Acute effects of low
      levels of ambient ozone on peak expiratory flow rate in a cohort of Australian children. Int. J. Epidemiol.
      29:  549-557.
Jamason, P. F.; Kalkstein, L. S.; Gergen, P. J. (1997) A synoptic evaluation of asthma hospital admissions in
      New York City. Am.  J. Respir. Crit. Care Med. 156: 1781-1788.
Janssen, N. A. H.; Van Vliet, P. H. N.; Aarts, F.; Harssema, H.; Brunekreef, B. (2001) Assessment of exposure to
      traffic related air pollution of children attending schools near motorways. Atmos. Environ. 35: 3875-3884.
Janssen, N. A. H.; Schwartz, J.; Zanobetti, A.; Suh, H. H. (2002) Air conditioning and source-specific particles as
      modifiers of the effect of PM10 on hospital admissions for heart and lung disease. Environ. Health Perspect.
      110:43-49.
Jedrychowski, W.; Becher, H.; Wahrendorf, J.; Basa-Cierpialek, Z. (1990) A case-control study of lung cancer with
      special reference to the effect of air pollution in Poland. J. Epidemiol.  Commun. Health 44: 114-120.
Jedrychowski, W.; Flak, E. (1998) Separate and combined effects of the outdoor and indoor air quality on chronic
      respiratory symptoms adjusted for allergy among preadolescent children. Int. J. Occup. Med. Environ. Health
      11:  19-35.
Jedrychowski, W.; Flak, E.;  Mroz, E. (1999) The adverse effect of low levels of ambient air pollutants on lung
      function growth in preadolescent children. Environ. Health Perspect. 107: 669-674.
Kalandidi, A.; Gratziou, C.;  Katsouyanni, K.; Manalis, N.; Tzala, L.; Pantazopoulou, A.; Efthimiou, M.;
      Roussos, C.; Roemer, W. (1998) Air pollution and respiratory health of children: the PEACE panel study in
      Athens, Greece. Eur.  Respir. Rev. 8: 117-124.
Katsouyanni, K.; Touloumi, G. (1998) Causes of regional differences in air pollution effects are being studied
      further [letter]. Br. Med. J. 316:  1982.
Katsouyanni, K.; Karakatsani, A.; Messari, L; Touloumi, G.; Hatzakis, A.; Kalandidi, A.; Trichopoulos, D. (1990)
      Air pollution and cause specific mortality in Athens. J. Epidemiol. Commun. Health 44: 321-324.
Katsouyanni, K.; Schwartz, J.; Spix, C.; Touloumi, G.; Zmirou, D.; Zanobetti, A.; Wojtyniak, B.; Vonk, J. M.;
      Tobias, A.; Ponka, A.; Medina, S.; Bacharova, L.; Andersen, H. R. (1996) Short term effects of air pollution
      on health:  a European approach using epidemiology time series data: the APHEA protocol. In: St Leger, S.,
      ed. The APHEA project. Short term effects of air pollution on health: a European approach using
      epidemiological time  series data. J. Epidemiol. Community Health 50(suppl.  1): S12-S18.
Katsouyanni, K.; Touloumi, G.; Spix, C.; Schwartz, J.; Balducci, F.; Medina, S.; Rossi, G.; Wojtyniak, B.;
      Sunyer, J.; Bacharova, L.; Schouten, J. P.; Ponka, A.; Anderson, H. R. (1997) Short term effects of ambient
      sulphur dioxide and paniculate matter on mortality in 12 European cities: results from time series data from
      the APHEA project. Br. Med. J. 314: 1658-1663.
Katsouyanni, K.; Touloumi, G.; Samoli, E.; Gryparis, A.; Le Tertre, A.; Monopolis, Y.; Rossi, G.; Zmirou, D.;
      Ballester, F.; Boumghar, A.; Anderson, H. R.; Wojtyniak, B.; Paldy, A.; Braunstein, R.; Pekkanen, J.;
      Schindler, C.; Schwartz, J. (2001) Confounding and effect modification in the short-term effects of ambient
      particles on total mortality: results from 29 European cities within the APHEA2 project. Epidemiology
      12:521-531.
Keatinge, W. R.;  Donaldson, G. C. (2001) Mortality  related to cold and air pollution in London after allowance for
      effects of associated weather patterns. Environ. Res. A 86: 209-216.
Keles, N.;  Ilicali, O. C.; Deger, K. (1999) Impact of air pollution on prevalence of rhinitis in Istanbul. Arch.
      Environ. Health 54: 48-51.
Kelsall, J. E.; Samet, J. M.; Zeger, S. L.; Xu, J. (1997) Air pollution and mortality in Philadelphia, 1974-1988.
      Am. J. Epidemiol. 146: 750-762.
Kingham, S.; Briggs, D.; Elliott, P.; Fischer, P.; Lebret, E. (2000) Spatial variations in the concentrations of
      traffic-related pollutants in indoor and outdoor air in Huddersfield, England. Atmos. Environ. 34: 905-916.
Klemm, R.  J.; Mason, R. M., Jr. (2000) Aerosol research and inhalation epidemiological study (ARIES): air quality
      and daily mortality statistical modeling—interim results. J. Air. Waste Manage. Assoc. 50: 1433-1439.

April 2002                                      8-296         DRAFT-DO NOT QUOTE OR CITE

-------
Klemm, R. I; Mason, R. M, Jr.; Heilig, C. M; Neas, L. M; Dockery, D. W. (2000) Is daily mortality associated
      specifically with fine particles? Data reconstruction and replication of analyses. J. Air Waste Manage. Assoc.
      50: 1215-1222.
Koenig, J. Q.; Larson, T. V.; Hanley, Q. S.; Rebolledo, V.; Dumler, K.; Checkoway, H.; Wang, S.-Z.; Lin, D.;
      Pierson, W. E. (1993) Pulmonary function changes in children associated with fine paniculate matter.
      Environ. Res. 63:26-38.
Kontos, A. S.; Fassois, S. D.; Deli, M. F. (1999) Short-term effects of air pollution on childhood respiratory illness
      in Piraeus, Greece, 1987-1992: nonparametric stochastic dynamic analysis. Environ. Res. 81: 275-296.
Koren, H. S.; Utell, M. J. (1997) Asthma and the environment. Environ. Health Perspect. 105: 534-537.
Korrick, S. A.; Neas, L. M.; Dockery, D. W.; Gold, D. R.; Allen, G. A.; Hill, L. B.; Kimball, K. D.; Rosner, B. A.;
      Speizer, F. E. (1998) Effects of ozone and other pollutants on the pulmonary function of adult hikers.
      Environ. Health Perspect. 106: 93-99.
Kostianev, S.; Gerginova, M.; Ivanova, M. (1994) Reference values of lung function parameters in Bulgarian girls
      aged 7 to 14 years. Pediatria 33: 30-33.
Kotesovec, F.; Vitnerova, N.; Leixner, M.; Benes, I.; Skorkovsky, J.; Roemer, W. (1998) Air pollution and
      respiratory health of children:  the PEACE panel study in Teplice, Czech Republic. Eur. Respir. Rev.
      8: 70-77.
Kotesovec, F.; Skorkovsky, J.; Brynda, J.; Peters, A.; Heinrich, J. (2000) Daily mortality and air pollution in
      northern Bohemia; different effects for men and women. Cent. Eur. J. Public Health 8:  120-127.
Kousa, A.; Monn, C.; Rotko, T.; Aim, S.; Oblesby, L.; Jantunen, M. J. (2001) Personal exposures to NO2 in the
      EXPOLIS-study: relation to residential indoor, outdoor and workplace concentrations in Basel, Helsinki and
      Prague.  Atmos. Environ. 35: 3405-3412.
Kramer, U.; Behrendt, H.; Dolgner, R.; Ranft, U.; Ring, J.; Wilier, H.; Schlipkoter, H.-W. (1999) Airway diseases
      and allergies in East and West German children during the first 5 years after reunification: time trends and
      the impact of sulphur dioxide and total suspended particles. Int. J. Epidemiol. 28: 865-873.
Krewski, D.; Burnett, R. T.; Goldberg, M. S.; Hoover, K.; Siemiatycki, J.; Jerrett, M.; Abrahamowicz, M.; White,
      W. H. (2000) Reanallysis of the Harvard Six Cities study and the American Cancer Society study of
      paniculate air pollution and mortality. A special report of the Institute's Particle Epidemiology Reanalysis
      Project. Cambridge, MA: Health Effects Institute.
Ktinzli, N.; Tager, I. B. (1997) The semi-individual study in air pollution epidemiology: a valid design as compared
      to ecologic studies. Environ. Health Perspect. 105: 1078-1083.
Ktinzli, N.; Ackermann-Liebrich, U.; Brandli, O.; Tschopp, J. M.; Schindler, C.; Leuenberger, P.; SAPALDIA
      Team. (2000) Clinically "small" effects of air pollution on FVC have a large public health impact. Eur.
      Respir. J. 15: 131-136.
Kwon, H.-J.; Cho, S.-H.; Nyberg, F.; Pershagen, G. (2001) Effects of ambient air pollution on daily mortality in a
      cohort of patients with congestive heart failure. Epidemiology 12: 413-419.
Laden, F.; Neas, L. M.; Dockery, D. W.; Schwartz, J. (2000) Association of fine paniculate matter from different
      sources with daily mortality in six U.S. cities. Environ. Health Perspect. 108: 941-947.
Lave, L. B.; Seskin, E. P. (1977) Air  pollution and human health. Baltimore, MD: The Johns Hopkins University
      Press.
Le Tertre, A.; Medina, S.; Samoli, E.; Forsberg, B.; Michelozzi, P.; Boumghar, A.; Vonk, J. M.; Bellini, A.;
      Atkinson, R.; Ayres, J. G.; Sunyer, J.; Schwartz, J.; Katsouyanni, K. (2002) Short term effects of paniculate
      air pollution on cardiovascular diseases in eight European cities. J. Epidemiol. Community Health: in press.
Lebowitz, M. D. (1971) A stimulus response method for the analysis of environmental and health events related in
      time. Environ. Lett. 2: 23-24.
Lebowitz, M. D.; Collins, L.; Holberg, C. J. (1987) Time series analyses of respiratory responses to indoor and
      outdoor environmental phenomena. Environ. Res. 43:  332-341.
Lee, J.-T.; Schwartz, J. (1999) Reanalysis of the effects of air pollution on daily mortality in Seoul, Korea:
      a case-crossover design. Environ. Health Perspect. 107: 633-636.
Lee, J.-T.; Shy, C. M. (1999) Respiratory function as measured by peak expiratory flow rate and PM10:
      six communities study. J. Exposure Anal. Environ. Epidemiol. 9: 293-299.
Lee, R. E., Jr.; Caldwell, J. S.; Morgan, G.  B. (1972) The evaluation of methods for measuring suspended
      particulates in air. Atmos. Environ. 6: 593-622.
Lee, J.-T.; Shin, D.;  Chung, Y. (1999) Air pollution and daily mortality in Seoul and Ulsan, Korea. Environ. Health
      Perspect. 107: 149-154.


April 2002                                      8-297         DRAFT-DO NOT QUOTE OR CITE

-------
Lee, J.-T.; Kim, H.; Hong, Y.-C.; Kwon, H.-J.; Schwartz, J.; Christiani, D. C. (2000) Air pollution and daily
      mortality in seven major cities of Korea, 1991-1997. Environ. Res. 84: 247-254.
Leonardi, G. S.; Houthuijs, D.; Steerenberg, P. A.; Fletcher, T.; Armstrong, B.; Antova, T. (2000) Immune
      biomarkers in relation to exposure to paniculate matter: a cross-sectional survey in 17 cities of central
      Europe. In: Grant, L. D., ed. PM2000: paniculate matter and health. Inhalation Toxicol. 12(suppl. 4): 1-14.
Levy, D. (1998) Fine paniculate  air pollution and out-of-hospital mortality in King County, Washington. In: Vostal,
      J. J., ed. Health effects of paniculate matter in ambient air. Proceedings of an international conference; 1997;
      Prague, Czech Republic. Pittsburgh, PA: Air & Waste Management Association; pp. 262-271. (A&WMA
      publication VIP-80).
Levy, J. I.; Hammitt, J. K.;  Spengler, J. D. (2000) Estimating the mortality impacts of paniculate matter: what can
      be learned frombetween-study variability? Environ. Health Perspect.  108: 109-117.
Lewis, P. R.; Hensley, M. J.; Wlodarczyk, J.; Toneguzzi, R. C.; Westley-Wise, V. J.; Dunn, T.; Calvert, D. (1998)
      Outdoor air pollution and  children's respiratory symptoms in the steel cities of New South Wales. Med. J.
      Aust. 169: 459-463.
Liao, D.; Creason, J.; Shy, C.; Williams, R.; Watts, R.; Zweidinger, R. (1999) Daily variation of paniculate air
      pollution and poor cardiac autonomic control in the elderly. Environ. Health Perspect. 107: 521-525.
Lin, C. A.; Martins, M. A.; Farhat, S. C. L.; Pope, C. A., Ill; Conceicao, G. M. S.; Anastacio, V. M.; Hatanaka, M.;
      Andrade, W. C.; Hamaue, W. R.; Bohm, G. M.; Saldiva, P. H. N. (1999) Air pollution and respiratory illness
      of children in Sao Paulo, Brazil. Paediatr. Perinat. Epidemiol. 13: 475-488.
Linn, W. S.; Shamoo, D. A.; Anderson, K. R.; Peng, R.-C.; Avol, E. L.; Hackney, J.  D.; Gong, H., Jr. (1996)
      Short-term air pollution exposures and responses in Los Angeles area schoolchildren. J. Exposure Anal.
      Environ. Epidemiol. 6: 449-472.
Linn, W. S.; Gong, H., Jr.; Clark, K. W.; Anderson, K. R. (1999) Day-to-day paniculate exposures and health
      changes in Los Angeles area residents with severe lung disease. J. Air Waste Manage. Assoc.
      49:PM108-PM115.
Linn, W. S.; Szlachcic, Y.;  Gong, H., Jr.; Kinney, P. L.; Berhane, K. T. (2000) Air pollution and daily hospital
      admissions in metropolitan Los Angeles. Environ. Health Perspect. 108: 427-434.
Lipfert,  F. W. (2000) The use and misuse of surrogate variables in environmental epidemiology. J. Environ. Med.
      1:267-278.
Lipfert,  F. W.; Wyzga, R. E. (1999) Statistical considerations in determining the health significance of constituents
      of airborne paniculate matter. J. Air Waste Manage. Assoc. 49: PM-182-191.
Lipfert,  F. W.; Morris, S. C.; Wyzga, R. E. (2000a) Daily mortality in the Philadelphia metropolitan area and
      size-classified paniculate  matter. J. Air Waste Manage. Assoc.: 1501-1513.
Lipfert,  F. W.; Perry, H. M., Jr.; Miller, J. P.; Baty, J. D.; Wyzga, R. E.; Carmody, S. E. (2000b) The Washington
      University-EPRI veterans' cohort mortality study: preliminary results. In: Grant, L. D., ed. PM2000:
      paniculate matter and health. Inhalation Toxicol. 12(suppl. 4): 41-73.
Lipfert,  F. W.; Zhang, J.; Wyzga, R. E. (2000c) Infant mortality and air pollution: a comprehensive analysis of U.S.
      data for 1990. J. Air Waste Manage. Assoc. 50: 1350-1366.
Lippmann, M.; Thurston, G. D. (1996) Sulfate concentrations as an indicator of ambient paniculate matter air
      pollution for health risk evaluations. J. Exposure Anal. Environ. Epidemiol. 6: 123-146.
Lippmann, M.; Ito, K.; Nadas, A.; Burnett, R. T. (2000) Association of paniculate matter components  with daily
      mortality and morbidity in urban populations. Cambridge, MA: Health Effects Institute; research report no.
      95.
Lipsett,  M.; Hurley, S.; Ostro, B. (1997) Air pollution and emergency room visits for asthma in Santa Clara County,
      California. Environ.  Health Perspect. 105: 216-222.
Long, W.; Tate, R. B.; Neuman, M.; Manfreda, J.; Becker, A. B.; Anthonisen, N. R.  (1998) Respiratory symptoms
      in a susceptible population due to burning of agricultural residue. Chest 113: 351-357.
Loomis, D.; Castillejos, M.; Gold, D. R.; McDonnell, W.; Borja-Aburto, V. H. (1999) Air pollution and infant
      mortality in Mexico  City.  Epidemiology 10: 118-123.
Lumley, T.; Heagerty, P. (1999) Weighted empirical adaptive variance estimators for correlated data regression.
      J. R. Stat. Soc. B 61(part 2): 459-477.
Lumley, T.; Sheppard, L. (2000)  Assessing seasonal confounding and model selection bias in air pollution
      epidemiology using positive and negative control analyses. Environmetrics 11: 705-717.
Mar, T.  F.; Norris, G. A.; Koenig, J. Q.; Larson, T. V. (2000) Associations between air pollution and mortality in
      Phoenix, 1995-1997. Environ. Health Perspect. 108: 347-353.


April 2002                                      8-298         DRAFT-DO NOT QUOTE OR CITE

-------
Marcus, A. H.; Kegler, S. R. (2001) Confounding in air pollution epidemiology: when does two-stage regression
      identify the problem? Environ. Health Perspect. 109: 1193-1196.
McConnell, R.; Berhane, K.; Gilliland, F.; London, S. J.; Vora, H.; Avol, E.; Gauderman, W. J.; Margolis, H. G.;
      Lurmann, F.; Thomas, D. C.; Peters, J. M. (1999) Air pollution and bronchitic symptoms in southern
      California children with asthma. Environ. Health Perspect. 107: 757-760.
McConnell, R.; Berhane, K.; Gilliland, F.; London, S. J.; Islam, T.; Gauderman, W. J.; Avol, E.; Margolis, H. G.;
      Peters, J. M. (2002) Asthma in exercising children exposed to ozone: a cohort study. Lancet 359: 386-391.
McDonnell, W. F.; Nishino-Ishikawa,  N.; Petersen, F. F.; Chen, L. H.; Abbey, D. E. (2000) Relationships of
      mortality with the fine and coarse fractions of long-term ambient PM10 concentrations in nonsmokers.
      J. Exposure Anal. Environ. Epidemiol. 10: 427-436.
McGregor, G. R.; Walters, S.;  Wordley,  J. (1999) Daily hospital respiratory admissions and winter air mass types,
      Birmingham, UK. Int. J. Biometeorol. 43: 21-30.
Medina, S.; Le Tertre, A.; Quenel, P.; Le Moullec, Y.; Lameloise, P.; Guzzo, J. C.; Festy, B.; Ferry, R.; Dab, W.
      (1997) Air pollution and doctors' house calls: results from the ERPURS system for monitoring the effects of
      air pollution on public health in greater Paris, France, 1991-1995. Environ. Res. 75: 73-84.
Mendelsohn, R.; Orcutt, G. (1979) An empirical analysis of air pollution dose-response curves. J. Environ. Econ.
      Manage. 6: 85-106.
Michelozzi, P.; Forastiere, F.; Fusco, D.; Perucci, C. A.; Ostro, B.; Ancona, C.; Pallotti, G. (1998) Air pollution and
      daily mortality in Rome, Italy. Occup. Environ. Med. 55: 605-610.
Mills, P. K.; Abbey, D.; Beeson, W. L.; Petersen, F. (1991) Ambient air pollution and cancer in California
      Seventh-day Adventists. Arch. Environ. Health 46: 271-280.
Monn, C. (2001) Exposure assessment of air pollutants: a review on spatial heterogeneity and
      indoor/outdoor/personal exposure to suspended paniculate matter, nitrogen dioxide and ozone. Atmos.
      Environ. 35: 1-32.
Monn, Ch.; Carabias, V.; Junker, M.; Waeber, R.; Karrer, M.; Wanner, H. U. (1997) Small-scale spatial variability
      of paniculate matter <10 um(PM10) and nitrogen dioxide. Atmos. Environ. 31: 2243-2247.
Moolgavkar, S. H. (2000a) Air Pollution and Mortality in Three U.S. Counties. Environ. Health Perspect.
      108: 777-784.
Moolgavkar, S. H. (2000b) Air pollution and hospital  admissions for diseases of the circulatory system in three U.S.
      metropolitan areas. J. Air Waste Manage Assoc. 50:  1199-1206.
Moolgavkar, S. H. (2000c) Air pollution and hospital admissions for chronic obstructive pulmonary disease in three
      metropolitan areas in the United States. In: Grant, L. D., ed. PM2000: paniculate matter and health.
      Inhalation Toxicol. 12(suppl. 4): 75-90.
Moolgavkar, S. H.; Luebeck, E. G. (1996) A critical review of the evidence on paniculate air pollution and
      mortality. Epidemiology 7: 420-428.
Moolgavkar, S. H.; Luebeck, E. G.;  Anderson, E. L. (1997) Air pollution and hospital admissions for respiratory
      causes in Minneapolis-St. Paul and Birmingham. Epidemiology 8: 364-370.
Moolgavkar, S. H.; Hazelton, W.; Luebeck, G.; Levy,  D.; Sheppard, L. (2000) Air pollution, pollens, and
      admissions for chronic respiratory disease in King County, Washington. In: Inhalation toxicology:
      proceedings of the third colloquium on paniculate air pollution and human health; June, 1999; Durham, NC.
      Inhalation Toxicology 12(suppl. 1): 157-171.
Morgan, G.; Corbett, S.; Wlodarczyk, J.; Lewis, P. (1998) Air pollution and daily mortality in Sydney, Australia,
      1989 through 1993. Am. J. Public Health 88: 759-764.
Morris, R. D. (2001) Airborne particulates and hospital admissions for cardiovascular disease: a quantitative review
      of the evidence. Environ. Health Perspect.  109(suppl. 4): 495-500.
Morris, R. D.; Naumova, E.  N. (1998) Carbon monoxide and hospital admissions for congestive heart failure:
      evidence of an increased effect  at low temperatures. Environ. Health Perspect. 106: 649-653.
Morris, R. D.; Naumova, E.  N.; Munasinghe, R. L. (1995) Ambient air pollution and hospitalization for congestive
      heart failure among elderly people in seven large US cities. Am. J. Public Health 85: 1361-1365.
Murray, C. J.; Nelson, C. R. (2000)  State-space modeling of the relationship between air quality and mortality.
      J. Air Waste Manage. Assoc. 50:  1075-1080.
Naeher, L. P.; Holford, T. R.; Beckett, W. S.; Belanger, K.; Triche, E. W.; Bracken, M. B.; Leaderer, B. P. (1999)
      Healthy women's PEF variations with ambient summer concentrations of PM10, PM2 5, SO42", H+,  and O3.
      Am. J. Respir. Crit. Care Med. 160: 117-125.
Nauenberg, E.; Basu, K. (1999) Effect of insurance coverage on the relationship between asthma hospitalizations
      and exposure to air pollution. Public Health Rep. 114: 135-148.

April 2002                                      8-299         DRAFT-DO NOT QUOTE OR CITE

-------
Neas, L. M; Dockery, D. W.; Ware, J. H.; Spengler, J. D.; Ferris, B. G., Jr.; Speizer, F. E. (1994) Concentration of
      indoor paniculate matter as a determinant of respiratory health in children. Am. J. Epidemiol.
      139: 1088-1099.
Neas, L. M.; Dockery, D. W.; Koutrakis, P.; Tollerud, D. J.; Speizer, F. E. (1995) The association of ambient air
      pollution with twice daily peak expiratory flow rate measurements in children. Am. J. Epidemiol.
      141: 111-122.
Neas, L. M.; Dockery, D. W.; Burge, H.; Koutrakis, P.; Speizer, F. E. (1996) Fungus spores, air pollutants, and
      other determinants of peak expiratory flow rate in children. Am. J. Epidemiol. 143: 797-807.
Neas, L. M.; Schwartz, J.; Dockery, D. (1999) A case-crossover analysis of air pollution and mortality in
      Philadelphia. Environ. Health Perspect. 107: 629-631.
Neukirch, F.; Segala, C.; Le Moullec, Y.; Korobaeff, M.; Aubier, M. (1998) Short-term effects of low-level winter
      pollution on respiratory health of asthmatic adults. Arch. Environ. Health 53: 320-328.
Norris, G.; Young-Pong, S. N.; Koenig, J. Q.; Larson, T. V.; Sheppard, L.; Stout, J. W. (1999) An association
      between fine particles and asthma emergency department visits for children in Seattle. Environ. Health
      Perspect. 107: 489-493.
Norris, G.; Larson, T.; Koenig, J.; Claiborn, C.; Sheppard, L.; Finn, D. (2000) Asthma aggravation, combustion, and
      stagnant air. Thorax 55: 466-470.
Nyberg, F.; Gustavsson, P.; Jarup, L.; Bellander, T.; Berglind, N.; Jakobsson, R.; Pershagen, G. (2000) Urban air
      pollution and lung cancer in Stockholm. Epidemiology 11: 487-495.
Oglesby, L.; Rotko, T.; Kriitli, P.; Boudet, C.; Kruize, H.; Jantunen, M. J.; Ktinzli, N. (2000) Personal exposure
      assessment studies may suffer from exposure-relevant selection bias. J. Exposure Anal. Environ. Epidemiol.
      10:251-266.
Ostro, B. (1995) Fine paniculate air pollution and mortality in two Southern California counties. Environ. Res.
      70: 98-104.
Ostro, B. D.; Lipsett, M. J.; Wiener, M. B.; Seiner, J. C. (1991) Asthmatic responses to airborne acid aerosols.
      Am. J. Public Health 81: 694-702.
Ostro, B. D.; Lipsett, M. J.; Mann, J.  K.; Braxton-Owens, H.; White, M. C. (1995)  Air pollution and asthma
      exacerbations among African-American children in Los Angeles. In: Phalen, R. F.; Bates, D. V., eds.
      Proceedings of the colloquium on paniculate air pollution and human mortality and morbidity, part II;
      January 1994; Irvine, CA. Inhalation Toxicol. 7: 711-722.
Ostro, B.; Chestnut, L.; Vichit-Vadakan, N.; Laixuthai, A. (1998) The impact of fine paniculate matter in Bangkok,
      Thailand. In: Chow, J.; Koutrakis, P., eds. PM2.5: a fine particle standard. Volume II: proceedings of an
      international specialty conference;  January; Long Beach, CA. Pittsburgh, PA: Air & Waste Management
      Association; pp. 939-949. (A&WMA publication VIP-81).
Ostro, B. D.; Hurley, S.; Lipsett, M. J. (1999a) Air pollution and daily mortality in the Coachella Valley, California:
      a study of PM10 dominated by  coarse particles. Environ. Res. 81: 231-238.
Ostro, B. D.; Eskeland, G.  S.; Sanchez, J. M.; Feyzioglu, T. (1999b) Air pollution and health effects: a study of
      medical visits among children  in Santiago, Chile. Environ. Health Perspect.  107: 69-73.
Ostro, B. D.; Broadwin, R.; Lipsett, M. J. (2000) Coarse and fine particles and daily mortality in the Coachella
      Valley, CA: a follow-up study. J. Exposure Anal. Environ. Epidemiol.: 10: 412-419.
Ostro, B.; Lipsett, M.; Mann, J.; Braxtron-Owens, H.; White,  M. (2001) Air pollution and exacerbation of asthma in
      African-American children in Los Angeles. Epidemiology 12: 200-208.
Osunsanya, T.; Prescott, G.; Seaton, A. (2001) Acute respiratory effects of particles: mass or number? Occup.
      Environ. Med. 58: 154-159.
Ozkaynak, H.; Thurston, G. D. (1987) Associations between 1980 U.S. mortality rates and alternative measures of
      airborne particle concentration. Risk Anal. 7: 449-461.
Ozkaynak, H.; Schatz, A. D.; Thurston, G. D.; Isaacs, R. G.; Husar, R. B. (1985) Relationships between aerosol
      extinction coefficients derived from airport visual range observations and alternative measures of airborne
      particle mass. J. AirPollut. Control Assoc. 35: 1176-1185.
Ozkaynak, H.; Xue, J.; Zhou, H.; Raizenne, M. (1996) Associations between daily mortality and motor vehicle
      pollution in Toronto, Canada. Boston, MA: Harvard University School of Public Health, Department of
      Environmental Health; March  25.
Pantazopoulou, A.; Katsouyanni, K.;  Kourea-Kremastinou, J.; Trichopoulos, D. (1995) Short-term effects of air
      pollution on hospital emergency outpatient visits and admissions in the greater Athens,  Greece area. Environ.
      Res. 69:31-36.


April 2002                                      8-300         DRAFT-DO NOT  QUOTE OR CITE

-------
Pekkanen, J.; Timonen, K. L.; Ruuskanen, J.; Reponen, A.; Mirme, A. (1997) Effects of ultrafine and fine particles
      in urban air on peak expiratory flow among children with asthmatic symptoms. Environ. Res. 74: 24-33.
Pekkanen, J.; Brunner, E. J.; Anderson, H. R.; Tiittanen, P.; Atkinson, R. W. (2000) Daily concentrations of air
      pollution and plasma fibrinogen in London. Occup. Environ. Med. 57: 818-822.
Penttinen, P.; Timonen, K. L.; Tiittanen, P.; Mirme, A.; Ruuskanen, J.; Pekkanen, J. (2001) Number concentration
      and size of particles in urban air: effects on spirometric lung function in adult asthmatic subjects. Environ.
      HealthPerspect. 109: 319-323.
Pereira, L. A. A.; Loomis, D.; Conceicao, G. M. S.; Braga, A. L. F.; Areas, R. M; Kishi, H. S.; Singer, J. M;
      Bohm, G. M.; Saldiva, P. H. N. (1998) Association between air pollution and intrauterine mortality in
      Sao Paulo, Brazil. Environ. Health Perspect. 106: 325-329.
Perry, H. M., Jr.; Schnaper, H. W.; Meyer, G.; Swatzell, R. (1982) Clinical program for screening and treatment of
      hypertension in veterans. J. Natl. Med. Assoc. 74: 433-444.
Peters, A.; Goldstein, I. F.; Beyer, U.; Franke, K.; Heinrich, J.; Dockery, D. W.; Spengler, J. D.; Wichmann, H.-E.
      (1996) Acute health effects of exposure to high levels of air pollution in eastern Europe. Am. J. Epidemiol.
      144: 570-581.
Peters, A.; Doring,  A.; Wichmann, H.-E.; Koenig,  W. (1997a) Increased plasma viscosity during an air pollution
      episode: a link to mortality? Lancet 349: 1582-1587.
Peters, A.; Dockery, D. W.; Heinrich, J.; Wichmann, H. E. (1997b) Short-term effects of paniculate air pollution on
      respiratory morbidity in asthmatic children.  Eur. Respir. J. 10: 872-879.
Peters, A.; Wichmann, H. E.; Tuch, T.; Heinrich, J.; Heyder, J. (1997c) Respiratory effects are associated with the
      number of ultrafine particles. Am. J. Respir. Crit. Care Med. 155: 1376-1383.
Peters, A.; Kotesovec, F.; Skorkovsky, J.; Brynda, J.; Heinrich, J. (1999a) Akute Auswirkung der
      Schwebstaubkonzentrationen in der Aulknluft auf die Mortalitat - Vergleichsstudie Nordost-Bayern /
      Nordbohmen. AbschluBbericht [Acute effects of suspended particle concentrations in the atmosphere on
      mortality - a study comparing northeast Bavaria and north Bohemia. Final report]. Bavaria, Federal Republic
      of Germany: InstitutfurEpidemiologie; report no. GSF-EP S 1/99.
Peters, J. M.; Avol, E.; Navidi,  W.; London, S. J.;  Gauderman, W. J.; Lurmann, F.; Linn, W.  S.; Margolis, H.;
      Rappaport, E.; Gong, H., Jr.; Thomas, D. C. (1999b) A study of twelve southern California communities with
      differing levels and types of air pollution. I. Prevalence of respiratory morbidity. Am. J. Respir. Crit. Care
      Med. 159: 760-767.
Peters, J. M.; Avol, E.; Gauderman, W. J.; Linn, W. S.; Navidi, W.; London, S. J.; Margolis, H.; Rappaport, E.;
      Vora, H.; Gong, H., Jr.; Thomas, D. C.  (1999c) A study of twelve southern California communities with
      differing levels and types of air pollution. II. Effects on pulmonary function. Am. J. Respir. Crit.  Care Med.
      159: 768-775.
Peters, A.; Liu, E.; Verrier, R. L.; Schwartz, J.; Gold, D. R.; Mittleman, M.; Baliff, J.; Oh, J. A.; Allen, G.;
      Monahan, K.; Dockery, D. W. (2000a) Air pollution and incidence of cardiac arrhythmia. Epidemiology
      11: 11-17.
Peters, A.; Frohlich, M.; Doring, A.; Immervoll, T.; Wichmann, H.-E.; Hutchinson, W.  L.; Pepys, M. B.;
      Koenig, W. (2000b) Paniculate air pollution is associated with an acute phase response in men: results from
      the MONICA-Augsburg Study. Eur. Heart J.: in press.
Peters, A.; Dockery, D. W.; Muller, J. E.; Mittleman, M. A. (2001) Increased paniculate air pollution and the
      triggering of myocardial infarction. Circulation 103: 2810-2815.
Petroeschevsky, A.; Simpson, R. W.; Thalib, L.; Rutherford, S. (2001) Associations between outdoor air pollution
      and hospital admissions in Brisbane, Australia. Arch. Environ. Health 56: 37-52.
Pike, M. C.; Jing, J. S.; Rosario, I. P.; Henderson, B. E.; Menck, H. R. (1979) Occupation: "explanation" of an
      apparent air pollution related localized excess of lung cancer in Los Angeles County. In: Breslow, N. E.;
      Whitemore, A. S., eds. Energy and Health: proceedings of a conference sponsored by Siam Institute for
      Mathematics and Society; 1978; Alta, Utah. [SIAM-SIMS conference series 6].
Pless-Mulloli, T.; Howel, D.; King, A.; Stone, I.; Merefield, J.; Bessell, J.; Darnell, R. (2000) Living near opencast
      coal mining sites and children's respiratory health. Occup. Environ. Med. 57: 145-151.
Ponce de Leon, A.; Anderson, H. R.; Bland, J. M.; Strachan, D. P.; Bower, J. (1996) Effects of air pollution on
      daily hospital admissions for respiratory disease in London between 1987-88 and 1991-92. In: St Leger, S.,
      ed. The APHEA project. Short term effects  of air pollution on health: a European approach using
      epidemiological time series data. J. Epidemiol. Community Health 50(suppl. 1): S63-S70.
Ponka, A.; Savela, M.; Virtanen, M. (1998) Mortality and air pollution in Helsinki. Arch.  Environ. Health
      53:281-286.

April 2002                                     8-301         DRAFT-DO NOT QUOTE OR CITE

-------
Pope, C. A., III. (1989) Respiratory disease associated with community air pollution and a steel mill, Utah Valley.
      Am. J. Public Health 79: 623-628.
Pope, C. A., III. (1991) Respiratory hospital admissions associated with PM10 pollution in Utah, Salt Lake, and
      Cache Valleys. Arch. Environ. Health 46: 90-97.
Pope, C. A., Ill; Kalkstein, L S. (1996) Synoptic weather modeling and estimates of the exposure-response
      relationship between daily mortality and paniculate air pollution. Environ. Health Perspect. 104: 414-420.
Pope, C. A., Ill; Dockery, D. W.; Spengler, J. D.; Raizenne, M. E. (1991) Respiratory health and PM10 pollution:
      a daily time series analysis. Am. Rev. Respir. Dis. 144: 668-674.
Pope, C. A., Ill; Schwartz, J.; Ransom, M. R. (1992) Daily mortality and PM10 pollution in Utah valley.
      Arch. Environ. Health 47: 211-217.
Pope, C. A., Ill; Thun, M. J.; Namboodiri, M. M.; Dockery, D. W.; Evans, J. S.; Speizer, F. E.; Heath, C. W., Jr.
      (1995) Paniculate air pollution as a predictor of mortality in a prospective study of U.S. adults. Am. J.
      Respir. Crit. Care Med. 151: 669-674.
Pope, C. A., Ill; Hill, R. W.; Villegas, G. M.  (1999a) Paniculate air pollution and daily mortality on Utah's Wasatch
      Front. Environ. Health Perspect. 107: 567-573.
Pope, C. A., Ill; Dockery, D. W.; Kanner, R. E.; Vollegas, G. M.; Schwartz, J. (1999b) Oxygen saturation, pulse
      rate, and paniculate air pollution: a daily time-series panel study. Am. J. Respir. Crit. Care Med.
      159: 365-372.
Pope, C. A., Ill; Verrier, R. L.; Lovett, E. G.; Larson, A. C.; Raizenne, M. E.; Kanner, R. E.; Schwartz, J.; Villegas,
      G. M.; Gold, D. R.; Dockery, D. W. (1999c) Heart rate variability associated with paniculate air pollution.
      Am. Heart J. 138: 890-899.
Pope, C. A., Ill; Burnett, R. T.; Thun, M. J.; Calle, E. E.; Krewski, D.; Ito, K.; Thurston, G. D. (2002) Lung cancer,
      cardiopulmonary mortality, and long-term exposure to fine paniculate air pollution. JAMA J. Am. Med.
      Assoc. 287: 1132-1141.
Prescott, G. J.; Cohen, G. R.; Elton, R.  A.; Fowkes, F. G. R.; Agius, R. M. (1998) Urban air pollution and
      cardiopulmonary ill health: a 14.5 year time series study. Occup. Environ. Med. 55: 697-704.
Prescott, E.; Lange, P.; Vestbo, J.;  Copenhagen City Heart Study Group. (1999) Socioeconomic status, lung
      function and admission to hospital for COPD: results from the Copenhagen City Heart Study. Eur. Respir. J.
      13: 1109-1114.
Qian, Z. Chapman, R. S.; Tian, Q.; Chen, Y.; Lioy, P. J.; Zhang, J. (2000) Effects of air pollution on children's
      respiratory  health in three Chinese cities. Arch. Environ. Health 55: 126-133.
Rahlenbeck, S. I.;  Kahl, H. (1996)  Air pollution and mortality in East Berlin during the winters of 1981-1989. Int. J.
      Epidemiol.  25: 1220-1226.
Raizenne, M.; Neas, L. M.; Damokosh, A. L; Dockery, D. W.; Spengler, J. D.; Koutrakis, P.; Ware, J. H.; Speizer,
      F. E. (1996) Health effects of acid aerosols on North American children: pulmonary function. Environ.
      Health Perspect. 104: 506-514.
Ransom, M. R.; Pope, C. A., III. (1992) Elementary school absences and PM10 pollution in Utah Valley. Environ.
      Res. 58: 204-219.
Rijnders, E.; Janssen, N. A.; Van Vliet, P. H.; Brunekreef, B. (2001) Personal and outdoor nitrogen dioxide
      concentrations in relation to degree of urbanization and traffic density. Environ. Health Perspect.
      109(suppl.  3): 411-417.
Ritz, B.; Yu, F. (1999) The effect of ambient carbon monoxide on low birth weight among children born in southern
      California between 1989 and 1993. Environ. Health Perspect. 107: 17-25.
Ritz, B.; Yu, F.; Fruin, S.; Chapa, G.; Shaw, G. M.; Harris, J. A. (2002) Ambient air pollution and risk of birth
      defects in southern California. Am. J. Epidemiol. 155:  17-25.
Rodes, C. E.; Holland, D. M. (1981) Variations of NO, NO2 and O3 concentrations downwind of aLos Angeles
      freeway. Atmos. Environ. 15: 243-250.
Roemer, W.; Hoek, G.; Brunekreef, B. (1993) Effect of ambient winter air pollution on respiratory health of
      children with chronic respiratory symptoms. Am. Rev. Respir. Dis. 147: 118-124.
Roemer, W.; Hoek, G.; Brunekreef, B.; Haluszka, J.; Kalandidi, A.; Pekkanen, J. (1998) Daily variations in air
      pollution and respiratory health in a multicentre study: the PEACE project. Eur. Respir. J. 12: 1354-1361.
Roemer, W.; Hoek, G.; Brunekreef, B.; Clench-Aas, J.; Forsberg, B.; Pekkanen, J.; Schutz, A. (2000) PM10
      elemental composition and acute respiratory health effects in European children (PEACE project). Eur.
      Respir. J. 15: 553-559.
Roemer, W. H.; Van Wijnen, J. H . (2001) Daily mortality and air pollution  along busy streets in Amsterdam,
      1987-1998. Epidemiology 12: 649-653.

April 2002                                     8-302         DRAFT-DO NOT QUOTE OR CITE

-------
Romieu, I.; Meneses, F.; Ruiz, S.; Sienra, J. J.; Huerta, J.; White, M. C.; Etzel, R. A. (1996) Effects of air pollution
      on the respiratory health of asthmatic children living in Mexico City. Am. J. Respir. Crit. Care Med.
      154: 300-307.
Romieu, I.; Meneses, F.; Ruiz, S.; Huerta, J.; Sienra, J. J.; White, M.; Etzel, R.; Hernandez, M. (1997) Effects of
      intermittent ozone exposure on peak expiratory flow and respiratory symptoms among asthmatic children in
      Mexico City. Arch. Environ. Health 52: 368-376.
Rooney, C.; McMichael, A. J.; Kovats, R. S.; Coleman, M. P. (1998) Excess mortality in England and Wales, and in
      greater London, during the 1995 heatwave. J. Epidemiol. Community Health 52: 482-486.
Roorda-Knape, M.  C.; Janssen, N.  A. H.; De Hartog, J. J.; Van Vliet, P. H. N.; Harssema, H.; Brunekreef, B. (1998)
      Air pollution from traffic in city districts near major motorways. Atmos. Environ. 32: 1921-1930.
Rosas, I.; McCartney, H. A.; Payne, R. W.; Calderon, C.; Lacey, J.; Chapela, R.; Ruiz-Velazco, S. (1998) Analysis
      of the relationships between environmental factors (aeroallergens, air pollution, and weather) and asthma
      emergency admissions to a hospital in Mexico City. Allergy (Copenhagen) 53: 394-401.
Rossi, G.; Vigotti, M. A.; Zanobetti, A.; Repetto, F.; Gianelle,  V.; and Schwartz, J. (1999) Air pollution and
      cause-specific mortality in Milan, Italy, 1980-1989. Arch. Environ. Health 54: 158-164.
Rothman, K. J.; Greenland, S., eds. (1998) Modern epidemiology. 2nd ed. Philadelphia, PA: Lippincott-Raven
      Publishers.
Rotko, T.; Koistinen, K.; Hanninen, O.; Jantunen, M. (2000) Sociodemographic descriptors of personal exposure to
      fine particles (PM25) inEXPOLISHelsinki. J. Exposure Anal. Environ. Epidemiol. 10: 385-393.
Rutherford, S.; Clark, E.; McTainsh, G.; Simpson, R.; Mitchell, C. (1999) Characteristics of rural dust events shown
      to impact on asthma severity in Brisbane, Australia. Int. J. Biometeorol. 42: 217-225.
Samet, J. M. (2000) What properties of paniculate matter are responsible for health effects? In: Inhalation
      Toxicology:  proceedings of the third colloquium on paniculate air pollution and human health; June, 1999;
      Durham, NC. Inhalation Toxicology 12(suppl. 1): 19-21.
Samet, J. M.; Zeger, S. L.; Kelsall, J. E.; Xu, J.; Kalkstein, L. S. (1996) Weather, air pollution and mortality in
      Philadelphia, 1973-1980, report to the Health Effects Institute on phase IB, Particle Epidemiology Evaluation
      Project. Cambridge, MA: Health Effects Institute; review draft.
Samet, J.; Zeger, S.; Kelsall, J.; Xu, J.; Kalkstein, L. (1998) Does weather confound of modify the association of
      paniculate air pollution with mortality? An analysis of the Philadelphia data, 1973-1980. Environ. Res.
      77: 9-19.
Samet, J. M.; Zeger, S. L.; Dominici, F.; Curriero, F.; Coursac, L;  Dockery, D. W.; Schwartz, J.; Zanobetti, A.
      (2000a) The national morbidity, mortality, and air pollution study. Part II: morbidity, mortality, and air
      pollution in the United States. Cambridge, MA: Health Effects Institute; research report no. 94.
Samet, J. M.; Dominici, F.; Zeger,  S. L.; Schwartz, J.; Dockery, D. W. (2000b) National morbidity, mortality, and
      air pollution study. Part I: methods and methodologic issues. Cambridge, MA: Health Effects Institute;
      research report no. 94.
Samet, J. M.; Dominici, F.; Curriero, F. C.; Coursac, L; Zeger, S. L. (2000c) Fine paniculate  air pollution and
      mortality in 20 U.S. cities, 1987-1994. N. Engl. J. Med.  343: 1742-1749.
Samoli, E.; Schwartz, J.; Wojtyniak, B.; Touloumi, G.; Spix, C.; Balducci, F.; Medina,  S.; Rossi, G.; Sunyer, J.;
      Bacharova, L.; Anderson, H. R.; Katsouyanni, K. (2001) Investigating regional differences in short-term
      effects of air pollution on daily mortality in the APHEA project: a sensitivity analysis for controlling
      long-term trends and seasonality. Environ. Health Perspect. 109: 349-353.
Sarnat, J. A.; Koutrakis, P.; Suh, H. H. (2000) Assessing the relationship between personal paniculate and gaseous
      exposures of senior citizens living in Baltimore, MD. J.  Air Waste Manage. Assoc. 50: 1184-1198.
Sarnat, J. A.; Schwartz, J.;  Catalano, P. J.; Suh, H. H. (2001) Gaseous pollutants in paniculate matter epidemiology:
      confounders or surrogates? Environ. Health Perspect. 109:  1053-1061.
Scarlett, J. F.; Abbott, K. J.; Peacock, J. L.; Strachan, D. P.; Anderson, H. R. (1996) Acute effects of summer air
      pollution on respiratory function in primary school children in southern England. Thorax 51:1109-1114.
Schouten, J. P.; Vonk, J. M.; de Graaf, A. (1996) Short term effects of air pollution on emergency hospital
      admissions for respiratory disease: results of the APHEA project in two major cities in The Netherlands,
      1977-89. In: St Leger, S., ed. The APHEA project. Short term effects of air pollution on health: a European
      approach using epidemiological time series data. J. Epidemiol. Community Health 50(suppl. 1): S22-S29.
Schwartz, J. (1991) Paniculate air pollution and daily mortality in Detroit. Environ. Res. 56: 204-213.
Schwartz, J. (1993) Air pollution and daily mortality in Birmingham, Alabama. Am. J. Epidemiol. 137: 1136-1147.
Schwartz, J. (1994) PM10, ozone, and hospital admissions for the elderly in Minneapolis, MN. Arch. Environ.
      Health 49: 366-374.

April 2002                                     8-303         DRAFT-DO NOT QUOTE OR CITE

-------
Schwartz, J. (1995) Short term fluctuations in air pollution and hospital admissions of the elderly for respiratory
      disease. Thorax 50: 531-538.
Schwartz, J. (1997) Air pollution and hospital admissions for cardiovascular disease in Tucson. Epidemiology
      8: 371-377.
Schwartz, J. (1999) Air pollution and hospital admissions for heart disease in eight U.S. counties. Epidemiology
      10: 17-22.
Schwartz, J. (2000a) Assessing confounding, effect modification, and thresholds in the association between ambient
      particles and daily deaths. Environ. Health Perspect. 108: 563-568.
Schwartz, J. (2000b) The distributed lag between air pollution and daily deaths. Epidemiology 11: 320-326.
Schwartz, J. (2000c) Harvesting and long term exposure effects in the relation between air pollution and mortality.
      Am. J. Epidemiol. 151: 440-448.
Schwartz, J. (2000d) Daily deaths are associated with combustion particles rather than SO2 in Philadelphia.
      Occup. Environ. Med. 57: 692-697.
Schwartz, J. (2001) Air pollution and blood markers of cardiovascular risk. Environ. Health Perspect.
      109(suppl. 3): 405-409.
Schwartz, J.; Dockery, D. W. (1992) Increased mortality in Philadelphia associated with daily air pollution
      concentrations. Am. Rev. Respir. Dis. 145: 600-604.
Schwartz, J.; Marcus, A. (1990) Mortality and air pollution in London: a time series analysis. Am. J. Epidemiol.
      131: 185-194.
Schwartz, J.; Morris, R. (1995) Air pollution and hospital admissions for cardiovascular disease in Detroit,
      Michigan. Am. J. Epidemiol. 142: 23-35.
Schwartz, J.; Neas, L. M. (2000) Fine particles are more strongly associated than coarse particles with acute
      respiratory health effects in schoolchildren. Epidemiology. 11: 6-10.
Schwartz, J.; Zanobetti, A.  (2000) Using meta-smoothing to estimate dose-response trends across multiple studies,
      with application to air pollution and daily death. Epidemiology 11: 666-672.
Schwartz, J.; Dockery, D. W.; Neas, L. M.; Wypij, D.; Ware, J. H.; Spengler, J. D.; Koutrakis, P.; Speizer, F. E.;
      Ferris, E.G., Jr. (1994) Acute effects of summer air pollution on respiratory symptom reporting in children.
      Am. J. Respir. Crit. Care Med. 150: 1234-1242.
Schwartz, J.; Dockery, D. W.; Neas, L. M. (1996a) Is daily mortality associated specifically with fine  particles?
      J. Air Waste Manage. Assoc. 46:  927-939.
Schwartz, J.; Spix, C.; Touloumi, G.; Bacharova, L.; Barumamdzadeh, T.; le Tertre, A.; Piekarksi, T.; Ponce de
      Leon, A.; Ponka, A.; Rossi,  G.; Saez, M.; Schouten, J. P. (1996b) Methodological issues in studies of air
      pollution and daily counts of deaths or hospital admissions. In: St Leger, S., ed. The APHEA project. Short
      term effects of air pollution on health: a European approach using epidemiological time series data.
      J. Epidemiol. Community Health 50(suppl. 1): S3-S11.
Schwartz, J.; Norris, G.; Larson, T.; Sheppard, L.; Claiborne, C.; Koenig, J. (1999) Episodes of high coarse particle
      concentrations are not associated with increased mortality. Environ. Health Perspect. 107: 339-342.
Schwartz, J.; Laden, F.; Zanobetti, A. (2002) The concentration-response relation between PM25 and daily deaths.
      Environ. Health Perspect.: in press.
Seaton,  A.; Soutar, A.; Crawford, V.; Elton, R.; McNerlan, S.; Cherrie, J.; Watt, M.; Agius, R.; Stout, R. (1999)
      Paniculate air pollution and the blood. Thorax 54: 1027-1032.
Segala, C.; Fauroux, B.; Just, J.; Pascual, L.; Grimfeld, A.; Neukirch, F. (1998) Short-term effect of winter air
      pollution on respiratory health of asthmatic children in Paris. Eur. Respir. J. 11: 677-685.
Sheppard, L.; Damian, D. (2000) Estimating short-term PM effects accounting for surrogate exposure
      measurements from ambient monitors. Environmetrics 11: 675-687.
Sheppard, L.; Levy, D.; Norris, G.; Larson, T. V.; Koenig, J. Q. (1999) Effects of ambient  air pollution on
      nonelderly asthma hospital admissions in Seattle, Washington, 1987-1994. Epidemiology 10: 23-30.
Shima, M.; Adachi, M.  (1998) Indoor nitrogen dioxide in homes along trunk roads with heavy traffic. Occup.
      Environ. Med.  55: 428-433.
Shumway, R. H.; Tai, R. Y.; Tai, L. P.; Pawitan, Y. (1983) Statistical analysis of daily London mortality and
      associated weather and pollution effects. Sacramento, CA: California Air Resources Board; contract no.
      Al-154-33.
Shumway, R. H.; Azari, A. S.; Pawitan,  Y. (1988) Modeling mortality fluctuations in Los Angeles as functions of
      pollution and weather effects. Environ. Res. 45: 224-241.
Simpson, R. W.; Williams, G.; Petroeschevsky, A.; Morgan, G.; Rutherford, S. (1997) Associations between
      outdoor air pollution and daily mortality in Brisbane, Australia. Arch. Environ.  Health 52:  442-454.

April 2002                                     8-304         DRAFT-DO NOT QUOTE OR CITE

-------
Smith, M. A.; Jalaludin, B.; Byles, J. E.; Lim, L.; Leader, S. R. (1996) Asthma presentations to emergency
      departments in western Sydney during the January 2194 bushfires. Int. J. Epidemiol. 25: 1227-1236.
Smith, R. L.; Spitzner, D.; Kim, Y.; Fuentes, M. (2000) Threshold dependence of mortality effects for fine and
      coarse particles in Phoenix, Arizona. J. Air Waste Manage. Assoc. 50: 1367-1379.
Soukup, J. M.; Ohio, A. J.; Becker, S. (2000) Soluble components of Utah Valley paniculate pollution alter alveolar
      macrophage function in vivo and in vitro. Inhalation Toxicol. 12: 401-414.
Spix, C.; Heinrich, J.; Dockery, D.; Schwartz, J.; Volksch, G.; Schwinkowski, K.; Collen, C.; Wichmann, H. E.
      (1993) Air pollution and daily mortality in Erfurt, East Germany, 1980-1989. Environ. Health Perspect.
      101: 518-526.
Spix, C.; Anderson, H. R.; Schwartz, J.; Vigotti, M. A.; LeTertre, A.; Vonk, J. M.; Touloumi, G.; Balducci, F.;
      Piekarski, T.; Bacharova, L.; Tobias, A.; Ponka, A.; Katsouyanni, K. (1998)  Short-term effects of air
      pollution on hospital admissions of respiratory diseases in Europe: a quantitative summary of APHEA study
      results. Arch. Environ. Health 53: 54-64.
Stieb, D. M.; Burnett, R. T.; Beveridge, R. C.; Brook, J.  R. (1996) Association between ozone and asthma
      emergency department visits in Saint John, New Brunswick, Canada. Environ. Health Perspect.
      104: 1354-1360.
Stieb, D. M.; Beveridge, R. C.; Rowe, B. H.; Walter, S. D.; Judek, S. (1998a) Assessing diagnostic classification in
      an emergency department: implications for daily time series studies of air pollution. Am. J. Epidemiol.
      148: 666-670.
Stieb, D. M.; Brook, J. R.; Broder, L; Judek, S.; Burnett, R. T.; Beveridge, R. C. (1998b) Personal exposure of
      adults with cardiorespiratory disease to paniculate acid and sulfate in Saint John, New Brunswick, Canada.
      Appl. Occup. Environ. Hyg. 13: 461-468.
Stieb, D. M.; Beveridge, R. C.; Brook, J. R.; Smith-Doiron, M.; Burnett, R. T.; Dales, R. E.; Beaulieu, S.; Judek, S.;
      Mamedov, A. (2000) Air pollution, aeroallergens  and cardiorespiratory emergency department visits in Saint
      John, Canada. J. Exposure Anal. Environ. Epidemiol.: 10: 461-477.
Sunyer, J.; Basagana, X. (2001) Particles, and not gases, are associated with the risk of death in patients with
      chronic obstructive pulmonary disease. Int. J. Epidemiol. 30: 1138-1140.
Sunyer, J.; Spix, C.; Quenel, P.; Ponce-de-Leon, A.; Ponka, A.; Barumandzadeh, T.; Touloumi, G.; Bacharova, L.;
      Wojtyniak, B.; Vonk, J.;  Bisanti, L.; Schwartz, J.; Katsouyanni, K. (1997) Urban air pollution and emergency
      admissions for asthma in four European cities: the APHEA project. Thorax 52: 760-765.
Sunyer, J.; Schwartz,  J.; Tobias, A.; Macfarlane, D.;  Garcia, J.; Anto, J. M. (2000) Patients with  chronic obstructive
      pulmonary disease are at increased risk of death associated with urban particle air pollution: a case-crossover
      analysis. Am. J. Epidemiol. 151: 50-56.
Taggart, S.  C. O.; Custovic, A.; Francis, H. C.; Faragher, E. B.; Yates, C. J.; Higgins, B. G.; Woodcock,  A. (1996)
      Asthmatic bronchial hyperresponsiveness varies with ambient levels of summertime air pollution.  Eur.
      Respir. J. 9: 1146-1154.
Tarn, A. Y. C.; Wong, C. M.; Lam, T. H.; Ong, S. G.; Peters, J.; Hedley, A. J. (1994) Bronchial responsiveness in
      children exposed to atmospheric pollution in Hong Kong. Chest 106: 1056-1060.
Tan, W. C.; Qiu, D.; Liam, B. L.; Ng, T. P.; Lee, S. H.; Van Eeden, S. F.; D'Yachkova, Y.; Hogg, J. C. (2000) The
      human bone marrow response to acute air pollution caused by forest fires. Am. J. Respir. Crit. Care Med.
      161: 1213-1217.
Tanaka, H.; Honma, S.; Nishi, M.; Igarashi, T.; Teramoto, S.; Nishio, F.; Abe, S. (1998) Acid fog and hospital visits
      for asthma: an  epidemiological study. Eur. Respir. J. 11: 1301-1306.
Tellez-Rojo, M. M.; Romieu, I.; Ruiz-Velasco, S.; Lezana, M.-A.; Hernandez-Avila, M. M. (2000) Daily respiratory
      mortality and PM10 pollution in Mexico City: importance of considering place of death. Eur. Respir. J.
      16: 391-396.
Tenias, J. M.; Ballester, F.; Rivera, M. L.  (1998) Association between hospital emergency visits for asthma and air
      pollution in Valencia, Spain. Occup. Environ. Med. 55: 541-547.
Thompson, A. J.; Shields, M. D.; Patterson, C. C. (2001) Acute asthma exacerbations and air pollutants in children
      living in Belfast, Northern Ireland. Arch. Environ. Health 56: 234-241.
Thurston, G. D.; Lippmann, M.; Scott, M. B.; Fine, J. M. (1997) Summertime haze  air pollution and children with
      asthma.  Am. J. Respir. Crit. Care Med. 155: 654-660.
Tiittanen, P.; Timonen, K. L.; Ruuskanen, J.; Mirme, A.; Pekkanen, J. (1999) Fine paniculate air pollution,
      resuspended road dust and respiratory health among symptomatic children. Eur. Respir. J.  13: 266-273.
Timonen, K. L.; Pekkanen, J. (1997) Air pollution and respiratory health among children with asthmatic or cough
      symptoms. Am. J. Respir. Crit. Care Med. 156: 546-552.

April 2002                                      8-305         DRAFT-DO NOT QUOTE OR CITE

-------
Tobias, A.; Campbell, M. J. (1999) Modelling influenza epidemics in the relation between black smoke and total
      mortality. A sensitivity analysis. J. Epidemiol. Community Health 53: 583-584.
Tolbert, P. E.; Klein, M.; Metzger, K. B.; Flanders, W. D.; Todd, K.; Mulholland, J. A.; Ryan, P. B.; Frumkin, H.
      (2000a) Interim results of the study of particulates and health in Atlanta (SOPHIA). J. Exposure Anal.
      Environ. Epidemiol. 10: 446-460.
Tolbert, P. E.; Mulholland, J. A.; Macintosh, D. L.; Xu, F.; Daniels, D.; Devine, O. J.; Carlin, B. P.; Klein, M.;
      Dorley, J.; Butler, A. J.; Nordenberg, D. F.; Frumkin, H.; Ryan,  P. B.; White, M. C. (2000b) Air quality and
      pediatric emergency room visits for asthma in Atlanta, Georgia.  Am. J. Epidemiol. 151: 798-810.
Touloumi, G.; Katsouyanni, K.; Zmirou, D.; Schwartz, J.; Spix, C.; Ponce de Leon, A.; Tobias, A.; Quennel, P.;
      Rabczenko, D.; Bacharova, L.; Bisanti, L.; Vonk, J. M.; Ponka, A. (1997) Short-term effects of ambient
      oxidant exposure on mortality: a combined analysis within the APHEA project. Am. J. Epidemiol.
      146: 177-185.
Tsai, F. C.; Apte, M. G.; Daisey, J. M. (1999) An exploratory analysis of the relationship between mortality and the
      chemical composition of airborne paniculate matter. Berkeley, CA: Lawrence Berkeley National Laboratory,
      Environmental Energy Technologies Division; report no. LBNL-43583.
Tsai, F. C.; Apte, M. G.; Daisey, J. M. (2000) An exploratory analysis of the relationship between mortality and the
      chemical composition of airborne paniculate matter. Inhalation Toxicol. 12(suppl.): 121-135.
Turnovska, T.; Kostianev,  S. (1999) Effects of reduced air pollution on children's pulmonary function. Cent. Eur. J.
      Public Health 7: 77-79.
U.S. Census Bureau. (1995) American housing survey  for the United States in 1993. Washington,  DC: U.S.
      Department of Commerce; current housing reports H150/93. Available:
      http://www.census.gov/hhes/www/housing/ahs/nationaldata.html [29 April, 2002].
U.S. Environmental Protection Agency. (1982) Air quality criteria for paniculate matter and sulfur oxides. Research
      Triangle Park, NC: Office of Health and Environmental Assessment, Environmental Criteria and Assessment
      Office; EPA report no. EPA-600/8-82-029aF-cF. 3v. Available from: NTIS, Springfield,  VA; PB84-156777.
U.S. Environmental Protection Agency. (1986) Second addendum to air quality criteria for paniculate matter and
      sulfur oxides (1982): assessment of newly available health effects information. Research Triangle Park, NC:
      Office of Health and Environmental Assessment, Environmental Criteria and Assessment Office; EPA report
      no. EPA-600/8-86-020F. Available from: NTIS, Springfield, VA; PB87-176574.
U.S. Environmental Protection Agency. (1993) Air quality criteria for oxides of nitrogen. Research Triangle Park,
      NC: Office of Health and Environmental Assessment, Environmental Criteria and Assessment Office; report
      nos. EPA/600/8-9!/049aF-cF. 3v. Available from: NTIS, Springfield, VA; PB95-124533, PB95-124525, and
      PB95-124517.
U.S. Environmental Protection Agency. (1996a) Air quality criteria for paniculate matter. Research Triangle Park,
      NC: National Center for Environmental Assessment-RTF Office; report nos. EPA/600/P-95/00laF-cF. 3v.
U.S. Environmental Protection Agency. (1996b) Air quality criteria for ozone and related photochemical oxidants.
      Research Triangle Park, NC: Office of Research and Development; report nos. EPA/600/AP-93/004aF-cF.
      3v. Available from: NTIS, Springfield, VA; PB96-185582, PB96-185590, andPB96-185608. Available
      online at: www.epa.gov/ncea/ozone.htm.
U.S. Environmental Protection Agency. (2000) Air quality criteria for carbon monoxide. Research Triangle Park,
      NC: National Center for Environmental Assessment; report no. EPA/600/P-99/001F. Available:
      www.epa.gov/ncea/co/ [2000, October 6].
VanDerZee, S. C.; Hoek, G.; Boezen, H. M.; Schouten, J. P.; Van Wijnen, J. H.; Brunekreef, B. (1999) Acute
      effects of urban air pollution on respiratory health of children with and without chronic respiratory
      symptoms. Occup. Environ. Med. 56: 802-813.
VanDerZee, S. C.; Hoek, G.; Boezen, M. H.; Schouten, J. P.; Van Wijnen, J. H.; Brunekreef, B. (2000) Acute
      effects of air pollution on respiratory health of 50-70 yr old adults. Eur. Respir. J. 15: 700-709.
Van Wijnen, J. H.; Van Der Zee, S. C. (1998) Traffic-related air pollutants: exposure of road users and populations
      living near busy roads. Rev. Environ. Health 13: 1-25.
Vasan, R. S.; Larson, M. G.; Leip, E. P.; Kannel, W. B.; Levy, D. (2001) Assessment of frequency of progression to
      hypertension in non-hypertensive participants in the Framingham Heart Study: a cohort study. Lancet
      358: 1682-1686.
Vedal, S.; Petkau, J.; White, R.; Blair, J. (1998) Acute  effects of ambient inhalable particles in asthmatic and
      nonasthmatic children. Am. J. Respir. Crit. Care Med. 157: 1034-1043.
Vena, J. E. (1982) Air pollution as a risk factor in lung cancer. Am. J. Epidemiol. 116: 42-56.


April 2002                                      8-306         DRAFT-DO NOT QUOTE OR CITE

-------
Venn, A. J.; Lewis, S. A.; Cooper, M; Hubbard, R.; Britton, J. (2001) Living near a main road and the risk of
      wheezing illness in children. Am. J. Respir. Crit. Care Med. 164: 2177-2180.
Veterans Administration Cooperative Study Group on Antihypertensive Agents. (1967) Effects of treatment on
      morbidity in hypertension. Results in patients with diastolic blood pressures averaging 115 through 129 mm
      Hg. JAMA J. Am. Med. Assoc. 202: 1028-1034.
Veterans Administration Cooperative Study Group on Antihypertensive Agents. (1970) Effects of treatment on
      morbidity in hypertension. II. Results in patients with diastolic blood pressure averaging 90 through 114 mm
      Hg. JAMA J. Am. Med. Assoc. 213: 1143-1152.
Vigotti, M. A.; Rossi, G.; Bisanti, L.; Zanobetti, A.; Schwartz, J. (1996) Short term effects of urban air pollution on
      respiratory health in Milan, Italy, 1980-89. In: Leger, S., ed. The APHEA project. Short term effects of air
      pollution on health: a European approach using epidemiological time series data. J. Epidemiol. Community
      Health 50(suppl. 1): S71-S75.
Wallace, L. (1996) Indoor particles: a review. J. Air Waste Manage. Assoc. 46: 98-126.
Wang, B.; Peng, Z.; Zhang, X.; Xu, Y.; Wang, H.; Allen, G.; Wang, L.; Xu, X. (1999) Paniculate matter, sulfur
      dioxide, and pulmonary function in never-smoking adults in Chongqing, China. Int. J. Occup. Environ.
      Healths: 14-19.
Ward, D. J.; Miller, M.  R.; Walters, S.; Harrison, R. M.; Ayres, J. G. (2000) Impact of correcting peak flow for
      nonlinear errors on air pollutant effect estimates from a panel study. Eur. Respir. J. 15:  137-140.
Ware, J. H.; Ferris, B. G., Jr.; Dockery, D. W.; Spengler, J. D.; Stram, D. O.; Speizer, F. E. (1986) Effects of
      ambient sulfur oxides and suspended particles on respiratory health of preadolescent children. Am. Rev.
      Respir. Dis. 133:834-842.
Wichmann, H.-E.; Spix, C.; Tuch, T.; Wolke, G.; Peters, A.; Heinrich, J.; Kreyling, W. G.; Heyder, J. (2000) Daily
      mortality and fine and ultrafine particles in Erfurt, Germany. Part I: role of particle number and particle mass.
      Cambridge, MA: Health Effects  Institute; Research Report no. 98.
Wilson, W. E.; Suh, H.  H. (1997) Fine particles and coarse particles: concentration relationships relevant to
      epidemiologic studies. J. Air Waste Manage. Assoc. 47: 1238-1249.
Wjst, M.; Reitmeir, P.; Dold, S.; Wulff, A.; Nicolai, T.; Von Loeffelholz-Colberg, E. F.; Von Mutius, E. (1993)
      Road traffic and  adverse effects on respiratory health in children. Br. Med. J. 307: 596-600.
Wolff, G. T.;  Stroup, C. M.; Stroup, D. P. (1983) The coefficient of haze as a measure of paniculate elemental
      carbon. J. AirPollut. Control Assoc. 33: 746-750.
Wong, C. M.; Lam, T. H.; Peters, J.; Hedley, A. J.; Ong, S. G.; Tarn, A. Y. C.; Liu, J.;  Spiegelhalter, D. J.  (1998)
      Comparison between two districts of the effects of an air pollution intervention on bronchial responsiveness
      in primary school children in Hong Kong. J. Epidemiol. Community Health 52:  571-578.
Wong, C. M.; Hu, Z.  G.; Lam, T. H.; Hedley, A. J.; Peters, J. (1999) Effects of ambient air pollution and
      environmental tobacco smoke on respiratory health of non-smoking women in Hong Kong. Int.  J. Epidemiol.
      28: 859-864.
Woodruff, T.  J.; Grille, J.;  Schoendorf, K. C. (1997) The relationship between selected causes of postneonatal
      infant mortality and paniculate air pollution in the United States. Environ. Health Perspect. 105: 608-612.
Wordley, J.; Walters, S.; Ayres, J. G. (1997) Short term variations in hospital admissions and mortality and
      paniculate air pollution. Occup. Environ. Med. 54: 108-116.
Wu, W.;  Samet, J. M.; Ohio, A. J.; Devlin, R. B. (2001) Activation of the EOF receptor signaling pathway in human
      airway epithelial cells exposed to Utah Valley PM. Am. J. Physiol. 281: L483-L489.
Xu, Z.; Yu, D.; Jing, L.; Xu, X. (2000)  Air pollution and daily mortality in Shenyang, China. Arch. Environ. Health
      55: 115-120.
Yang, W.; Jennison, B.  L.; Omaye, S. T. (1997) Air pollution and asthma emergency room visits in Reno, Nevada.
      Inhalation Toxicol. 9: 15-29.
Ye, F.; Piver,  W. T.; Ando, M.; Portier, C. J. (2001) Effects of temperature and air pollutants on cardiovascular and
      respiratory diseases for males and females  older than 65 years of age in Tokyo,  July and August 1980-1995.
      Environ.  Health Perspect. 109: 355-359.
Yu, O.; Sheppard, L.; Lumley, T.; Koenig, J. Q.; Shapiro, G. G. (2000) Effects of ambient air pollution on
      symptoms of asthma in Seattle-area children enrolled in the CAMP study. Environ. Health Perspect.
      108: 1209-1214.
Zanobetti, A.; Schwartz, J. (2000) Race, gender, and social status as modifiers of the effects of PM10 on mortality.
      J. Occup. Environ. Med. 42: 469-474.
Zanobetti, A.; Schwartz, J. (2001) Are diabetics more susceptible to the health effects  of airborne particles? Am. J.
      Respir. Crit. Care Med. 164: 831-833.

April 2002                                      8-307         DRAFT-DO NOT  QUOTE OR CITE

-------
Zanobetti, A.; Schwartz, J.; Dockery, D. W. (2000a) Airborne particles are a risk factor for hospital admissions for
      heart and lung disease. Environ. Health Perspect. 108: 1071-1077.
Zanobetti, A.; Wand, M. P.; Schwartz, J.; Ryan, L. M. (2000b) Generalized additive distributed lag models:
      quantifying mortality displacement. Biostatistics 1: 279-292.
Zeger, S. L.; Dominici, F.; Samet, J. (1999) Harvesting-resistant estimates of air pollution effects on mortality.
      Epidemiology 10: 171-175.
Zeger, S. L.; Thomas, D.; Dominici, F.; Samet, J. M.; Schwartz, J.; Dockery, D.; Cohen, A. (2000) Exposure
      measurement error in time-series studies of air pollution: concepts and consequences. Environ. Health
      Perspect. 108: 419-426.
Zeghnoun, A.; Beaudeau, P.; Carrat, F.; Delmas, V.; Boudhabhay, O.; Gayon, F.; Guincetre, D.; Czernichow, P.
      (1999) Air pollution and respiratory drug sales in the city of Le Havre, France, 1993-1996. Environ. Res.
      81:224-230.
Zeghnoun, A.; Czernichow, P.; Beaudeau, P.; Hautemaniere, A.; Froment, L.;  Le Tertre, A.; Quenel, P. (2001)
      Short-term effects of air pollution on mortality in the cities of Rouen and Le Havre, France, 1990-1995.
      Arch. Environ. Health 56: 327-335.
Zemp, E.; Elsasser, S.; Schindler, C.; Ktinzli, N.; Perruchoud, A. P.; Domenighetti, G.; Medici, T.;
      Ackermann-Liebrich, U.; Leuenberger, P.; Monn, C.; Bolognini, G.; Bongard, J.-P.; Brandli, O.; Karrer, W.;
      Keller, R.; Schoni, M. H.; Tschopp, J.-M; Villiger, B.; Zellweger, J.-P.; SAPALDIA Team. (1999)
      Long-term ambient air pollution and respiratory symptoms in adults (SAPALDIA study). Am. J. Respir. Crit.
      Care Med. 159: 1257-1266.
Zhang, J.; Qian, Z.; Kong, L.; Zhou, L.; Yan, L.; Chapman, R. S. (1999) Effects of air pollution on respiratory
      health of adults in three Chinese cities. Arch. Environ. Health 54: 373-381.
Zhang, H.; Triche, E.; Leaderer, B. (2000) Model forthe analysis of binary time series of respiratory symptoms.
      Am. J. Epidemiol. 151:  1206-1215.
Zmirou, D.; Schwartz, J.; Saez, M.; Zanobetti, A.; Wojtyniak, B.; Touloumi, G.; Spix, C.; Ponce de Leon, A.;
      Le Moullec, Y.; Bacharova, L.; Schouten, J.; Ponka, A.; Katsouyanni, K. (1998) Time-series analysis of air
      pollution and cause-specific mortality. Epidemiology 9: 495-503.
April 2002                                      8-308        DRAFT-DO NOT QUOTE OR CITE

-------
                  APPENDIX 8A
     SHORT-TERM PM EXPOSURE-MORTALITY
           STUDIES: SUMMARY TABLE
April 2002                  8A-1     DRAFT-DO NOT QUOTE OR CITE

-------
                       TABLE 8A-1.  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS  STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
IQRin//g/m3.
Study Description: Outcomes, Mean outcome rate, and
ages.  Concentration measures or estimates.  Modeling
     methods: lags, smoothing, and covariates.
            Results and Comments.
Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                              PM Index, lag, Excess Risk%
                                                                                                                                                            (95% LCL, UCL), Co-pollutants.
OO
>
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          United States

          Samet et al. (2000a,b).
          90 largest U.S. cities.
          1987-1994.
          PM10 mean ranged from
          15.3 (Honolulu) to
          52.0 (Riverside).
          Dominici et al. (2000).
          20 largest U.S. cities.
          1987-1994. PM10 mean
          ranged from 23.8 /j,g/w?
          (San Antonio) to
          52.0 Mg/m3 (Riverside).
                              Non-accidental total deaths and cause-specific (cardiac,
                              respiratory, and the other remaining) deaths, stratified in
                              three age groups (<65, 65-75, 75+), were examined for their
                              associations with PM10, O3, SO2, NO2, and CO (single, two,
                              and three pollutant models) at lags 0, 1, and 2 days. In the
                              first stage of the hierarchical model, RRs for the pollutants
                              for each city were obtained using GAM Poisson regression
                              models, adjusting for temperature and dewpoint (0-day and
                              average of 1-3 days for both variables), day-of-week,
                              seasonal cycles, intercept and seasonal cycles for three age
                              groups.  In the second stage, between-city variation in RRs
                              were modeled within region. The third stage modeled
                              between-region variation (7 regions).  Two alternative
                              assumptions were made regarding the prior distribution:
                              one with possibly substantial heterogeneity and the other
                              with less or no heterogeneity within region. The weighted
                              second-stage regression included five types of county-
                              specific variables:  (1) mean weather and pollution
                              variables; (2) mortality rate; (3) socio-demographic
                              variables; (4) urbanization; (5) variables related to
                              measurement error.

                              Non-accidental total deaths (stratified in three age
                              groups: <65,  65-75, 75+) were examined for their
                              associations with PM10 and O3 (single, 2, and
                              3 pollutant models) at lags 0, 1, and 2 days. In the
                              first stage of the hierarchical model, RRs for PM10
                              and O3 for each city were obtained using GAM
                              Poisson regression models, adjusting for temperature
                              and dewpoint (0-day and average of 1-3 days for both
                              variables), day-of-week, seasonal cycles,  intercept
                              and seasonal cycles for three age groups.  In the
                              second stage,  between-city variation in RRs were
                              modeled as a function of city-specific covariates
                              including mean PM10 and O3 levels, percent poverty,
                              and percent of population with age 65 and over. The
                              prior distribution assumed heterogeneity across cities.
                              To  approximate the posterior distribution, a Markov
                              Chain Monte  Carlo (MCMC) algorithm with a block
                              Gibbs sampler was implemented.  The second stage
                              also considered spatial model, in which RRs in closer
                              cities were assumed to be more correlated.
                                                    The estimated city-specific coefficients were mostly positive at
                                                    lag 0, 1, and 2 days (estimated overall effect size was largest at
                                                    lag 1, with the estimated percent excess death rate per 10 ,ug/m3
                                                    PM10 being about 0.5%).  The posterior probabilities that the
                                                    overall effects are greater than 0 at these lags were 0.99, 1.00,
                                                    and 0.98, respectively. None of the county-specific variables
                                                    (effect modifiers) in the second-stage regression significantly
                                                    explained the heterogeneity of PM10 effects across cities. In the
                                                    3-stage regression model with the  index for 7 geographical
                                                    regions, the effect of PM10 varied somewhat across the 7 regions,
                                                    with the effect in the Northeast being the greatest.  Adding O3
                                                    and other gaseous pollutants did not markedly change the
                                                    posterior distributions of PM10 effects.  O3 effects, as examined
                                                    by season, were associated with mortality in summer (-0.5
                                                    percent per 10 ppb increase), but not in all season data (negative
                                                    in winter).
                                                    Lag 1 day PM10 concentration positively associated with
                                                    total mortality in most locations (only 2 out of
                                                    20 coefficients negative), though estimates ranged from
                                                    2.1% to -0.4% per 10 Mg/m3 PM10 increase. PM10
                                                    mortality associations changed little with the addition of
                                                    O3 to the model, or with the addition of a third pollutant
                                                    in the model. The pattern of PM10 effects with respiratory
                                                    and cardiovascular were similar to that of total mortality.
                                                    The PM10 effect was smaller (and weaker) with other
                                                    causes of deaths.  The pooled analysis of 20 cities data
                                                    confirmed the overall effect on total and cardiorespiratory
                                                    mortality, with lag 1 day showing largest effect estimates.
                                                    The posterior distributions for PM10 were generally not
                                                    influenced by addition of other pollutants.  In the data for
                                                    which the distributed lags could be examined (i.e., nearly
                                                    daily data), the sum of 7-day distributed lag coefficients
                                                    was greater than each of single day coefficients. City-
                                                    specific covariates did not predict the heterogeneity
                                                    across cities. Regional model results suggested that PM10
                                                    effects in West U.S. were larger than in East and South.
                                                     Posterior mean estimates and 95%
                                                     credible intervals for total mortality
                                                     excess deaths per 50 ,ug/m3 increase in
                                                     PM10 at lag 1 day: 2.3% (0.1, 4.5) for
                                                     "more heterogeneity" across-city
                                                     assumption; 2.2% (0.5, 4.0) for "less or
                                                     no heterogeneity" across cities
                                                     assumption. The largest PM10 effect
                                                     estimated for 7 U.S. regions was for the
                                                     Northeast: 4.6% (2.7, 6.5) excess
                                                     deaths per 50 //g/m3 PM10 increment.
                                                     Total mortality excess deaths per 50
                                                     Mg/m3 increase in PM10:  1.8 (-0.5,
                                                     4.1) for lag 0; 1.9 (-0.4, 4.3) for lag
                                                     1;  1.2 (-1.0, 3.4) for lag 2.

                                                     Cardiovascular disease excess
                                                     deaths per 50 /-ig/m3 PM10:
                                                     3.4(1.0,5.9).

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
Study Description:  Outcomes, Mean outcome rate, and
ages. Concentration measures or estimates. Modeling
      methods:  lags, smoothing, and covariates.
Results and Comments. Design Issues, Uncertainties,
             Quantitative Outcomes.
                                                                                                                                                                   PM Index, lag, Excess Risk%
                                                                                                                                                                  (95% LCL, UCL), Co-pollutants.
OO
>
          United States (cont'd)

          Braga et al. (2000). Five U.S.
          cities: Pittsburgh, PA; Detroit,
          MI; Chicago, IL; Minneapolis-
          St. Paul, MN; Seattle, WA.
          1986-1993. PM10 means were
          35, 37, 37, 28, and 33 //g/m3,
          respectively in these cities.
          Braga etal. (2001).
          Ten U.S. cities.
          Same as Schwartz (2000b).
                               Potential confounding caused by respiratory epidemics on      When respiratory epidemics were adjusted for, small decreases     The overall estimated percent excess
                               PM-total mortality associations was investigated in a subset
                               of the 10 cities evaluated by Schwartz (2000a,b), as
                               summarized below. GAM Poisson models were used to
                               estimate city-specific PM10 effects, adjusting for
                               temperature, dewpoint, barometric pressure, time-trend and
                               day-of-week. A cubic polynomial was used to for each
                               epidemic period, and  a dummy variable was used to control
                               for isolated epidemic  days. Average of 0 and 1 day lags
                               were used.

                               The study examined the lag structure of PM10 effects on
                               respiratory and cardiovascular cause-specific mortality.
                               Using GAM Poisson model adjusting for temporal pattern
                               and weather, three types of lag structures were examined:
                               (1) 7-day unconstrained distributed lags; (2) 2-day average
                               (0- and 1-day lag); and (3) 0-day lag. The results were
                               combined across 10 cities.
                                                      in the PM10 effect were observed (9% in Chicago, 11% in
                                                      Detroit, 3% in Minneapolis, 5% in Pittsburgh, and 15% in
                                                      Seattle).
                                                      The authors reported that respiratory deaths were more affected
                                                      by air pollution levels on the previous days, whereas
                                                      cardiovascular deaths were more affected by same-day pollution.
                                                      Pneumonia, COPD, all cardiovascular disease, and myocardial
                                                      infarction were all associated with PM10 in the three types of lags
                                                      examined.  The 7-day unconstrained lag model did not always
                                                      give larger effect size estimates compared others.
                                                       deaths per 50 ,ug/m3 increase in PM10
                                                       was 4.3% (3.0, 5.6) before controlling
                                                       for epidemics and 4.0% (2.6, 5.3) after.
                                                       Average of 0 and 1 day lags.
                                                       In the 7-day unconstrained distributed
                                                       lag model, the estimated percent excess
                                                       deaths per 10ug/m3 PM10 were 2.7%
                                                       (1.5,2.9), 1.7% (0.1, 3.3), 1.0% (0.6,
                                                       1.4), and 0.6% (0.0,  1.2) for
                                                       pneumonia, COPD, all cardiovascular,
                                                       and myocardial infarction mortality,
                                                       respectively.
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Schwartz (2000a).
          Ten U.S. cities: New Haven,
          CT; Pittsburgh, PA; Detroit,
          MI; Birmingham, AL;
          Canton, OH; Chicago, IL;
          Minneapolis-St. Paul, MN;
          Colorado Springs, CO;
          Spokane, WA; and Seattle,
          WA. 1986-1993. PM10 means
          were 29, 35, 36, 37, 29, 37, 28,
          27, 41, and 33, respectively in
          these cities.
                               Daily total (non-accidental) deaths (20, 19, 63, 60, 10, 133,
                               32, 6, 9, and 29, respectively in these cities in the order
                               shown left).  Deaths stratified by location of death (in or
                               outside hospital) were also examined.  For each city, a GAM
                               Poisson model adjusting for temperature, dewpoint,
                               barometric pressure, day-of-week, season, and time was
                               fitted.  The data were also analyzed by season (November
                               through April as heating season).  In the second stage, the
                               PM10 coefficients were modeled as a function of city-
                               dependent covariates including copollutant to PM10
                               regression coefficient (to test confounding), unemployment
                               rate, education, poverty level, and percent non-white.
                               Threshold effects were also examined. The inverse variance
                               weighted averages of the ten cities' estimates were used to
                               combine results.
                                                      PM10 was significantly associated with total deaths, and the
                                                      effect size estimates were the same in summer and winter.
                                                      Adjusting for other pollutants did not substantially change PM10
                                                      effect size estimates.  Also, socioeconomic variables did not
                                                      modify the estimates. The effect size estimate for the deaths that
                                                      occurred outside hospitals was substantially greater than that for
                                                      inside hospitals. The effect size estimate was larger for subset
                                                      with PM10 less than 50 /ig/m3.
                                                       The total mortality RR estimates
                                                       combined across cities per 50 //g/m3
                                                       increase of mean of lag 0- and 1-days
                                                       PM10: overall 3.4 (2.7, 4.1); summer
                                                       3.4 (2.4, 4.4); winter 3.3 (2.3, 4.4); in-
                                                       hospital 2.5 (1.5, 3.4); out-of-hospital
                                                       4.5 (3.4, 5.6); days < 50 ,ug/m3 4.4 (3.1,
                                                       5.7); with SO2 2.9 (1.2, 4.6); with CO
                                                       4.6 (3.2, 6.0); with O3 3.5 (1.6, 5.3).
o

-------
                   TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS  STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
Study Description:  Outcomes, Mean outcome rate, and
ages. Concentration measures or estimates. Modeling
      methods:  lags, smoothing, and covariates.
                                                                                                                    Results and Comments.
                                                                                                       Design Issues, Uncertainties, Quantitative Outcomes.
     PM Index, lag, Excess Risk%
   (95% LCL, UCL), Co-pollutants.
OO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          United States (cont'd)

          Schwartz (2000b).
          Ten U.S. cities: New Haven,
          CT; Pittsburgh, PA;
          Birmingham, AL; Detroit, MI;
          Canton, OH; Chicago, IL;
          Minneapolis-St. Paul, MN;
          Colorado Springs, CO;
          Spokane, WA; and Seattle,
          WA. 1986-1993. PM10 means
          were 29, 35, 36, 37, 29, 37, 28,
          27, 41, and 33, respectively in
          these cities.

          Schwartz and Zanobetti (2000).
          Ten U.S. cities.
          Same as above.
          Zanobetti and Schwartz (2000).
          Four U.S. cities: Chicago, IL;
          Detroit, MI; Minneapolis-St.
          Paul, MN; Pittsburgh, PA.
          1986-1993. PM10 median = 33,
          33, 25, and 31 respectively for
          these cities.
                               The issue of distributed lag effects was the focus of this
                               study.  Daily total (non-accidental) deaths of persons 65
                               years of age and older were analyzed. For each city, a GAM
                               Poisson model adjusting for temperature, dewpoint,
                               barometric pressure, day-of-week, season, and time was
                               fitted.  Effects of distributed lag were examined using four
                               models: (1) 1-day mean at lag 0 day; (2) 2-day mean at lag 0
                               and 1 day; (3) second-degree distributed lag model using
                               lags 0 through 5 days; (4) unconstrained distributed lag
                               model using lags 0 through 5 days.
                               The inverse variance weighted averages of the ten cities'
                               estimates were used  to combine results.

                               The issue of a threshold in PM-mortality exposure-response
                               curve was the focus  of this study.  First, a simulation was
                               conducted to show that the "meta-smoothing" could
                               produce unbiased exposure-response curves.  Three
                               hypothetical curves (linear, piecewise linear, and
                               logarithmic curves) were used to generate mortality series in
                               10 cities, and GAM  Poisson models were used to estimate
                               exposure response curve. Effects of measurement errors
                               were also simulated. In the analysis of actual 10 cities data,
                               GAM Poisson models were fitted, adjusting for temperature,
                               dewpoint, and barometric pressure, and day-of-week.
                               Smooth function of PM10 with the same span (0.7) in each
                               of the cities. The predicted values of the log relative risks
                               were computed for 2 ,ug/m3 increments between 5.5 ,ug/m3
                               and 69.5 ,ug/m3 of PM10 levels.  Then, the predicted values
                               were combined across cities using inverse-variance
                               weighting.

                               Separate daily counts of total non-accidental deaths,
                               stratified by sex, race (black and white), and education
                               (education > 12yrs or not), were examined to test hypothesis
                               that people in each of these groups had higher risk of PM10.
                               GAM Poisson models adjusting for temperature, dewpoint,
                               barometric pressure, day-of-week, season, and time were
                               used. The mean of 0- and 1-day lag PM10 was used. The
                               inverse variance weighted averages of the four cities'
                               estimates were used  to combine results.
                                                      The effect size estimates for the quadratic distributed model and
                                                      unconstrained distributed lag model were similar. Both
                                                      distributed lag models resulted in substantially larger effect size
                                                      estimates than the single day lag, and moderately larger effect
                                                      size estimates than the two-day average models.
                                                      The simulation results indicated that the "meta-smoothing"
                                                      approach did not bias the underlying relationships for the linear
                                                      and threshold models, but did result in a slight downward bias
                                                      for the logarithmic model. Measurement error (additive or
                                                      multiplicative) in the simulations did not cause upward bias in
                                                      the relationship below threshold. The threshold detection in the
                                                      simulation was not very sensitive to the choice of span in
                                                      smoothing. In the analysis of real data from 10 cities, the
                                                      combined curve did not show evidence of a threshold in the
                                                      PM10-mortality associations.
                                                      The differences in the effect size estimates among the various
                                                      strata were modest. The results suggest effect modification with
                                                      the slope in female deaths one third larger than in male deaths.
                                                      Potential interaction of these strata (e.g., black and female) were
                                                      not investigated.
                                                                                                                                                               Total mortality percent increase
                                                                                                                                                               estimates combined across cities per
                                                                                                                                                               50 /^g/m3 increase in PM10: 3.3 (2.5,
                                                                                                                                                               4.1) for 1-day mean at lag 0; 5.4 (4.4,
                                                                                                                                                               6.3) 2-day mean of lag 0 and 1; 7.3
                                                                                                                                                               (5.9, 8.6) for quadratic distributed lag;
                                                                                                                                                               and 6.6 (5.3, 8.0) for unconstrained
                                                                                                                                                               distributed lag.
                                                                                                                                                               The combined exposure-response curve
                                                                                                                                                               indicates that an increase of 50 ,ug/m3 is
                                                                                                                                                               associated with about a 4% increase in
                                                                                                                                                               daily deaths.  Avg. of 0 and 1 day lags.
The total mortality RR estimates
combined across cities per 50 ,ug/m3
increase of mean of lag 0- and 1-days
PM10: white 5.0 (4.0, 6.0); black 3.9
(2.3, 5.4); male 3.8 (2.7, 4.9); female
5.5 (4.3, 6.7); education <12y 4.7 (3.3,
6.0); education> 12y3.6 (1.0, 6.3).

-------
                    TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
 to
 o
 o
 to
Reference, Location, Years,
PM Index, Mean or Median,
Study Description: Outcomes, Mean outcome rate, and
ages. Concentration measures or estimates. Modeling
     methods: lags, smoothing, and covariates.
             Results and Comments.
Design Issues, Uncertainties, Quantitative Outcomes.
  PM Index, lag, Excess Risk%
(95% LCL, UCL), Co-pollutants.
 OO
 >
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           United States (cont'd)

           Moolgavkar (2000a)
           Cook County, Illinois
           Los Angeles County, CA
           Maricopa County, AZ
           1987-1995
           PM10, CO, O3, NO2, SO2 in
           all three locations.
           PM2 5 in Los Angeles County.
           Cook Co:
           PM10 Median = 47 //g/m3.
           Maricopa Co:
           PM10 Median = 41.
           Los Angeles Co:
           PM10 Median = 44;
           PM,, Median = 22.
                               Associations between air pollution and time-series of daily
                               deaths evaluated for three U.S. metropolitan areas with
                               different pollutant mixes and climatic conditions. Daily
                               total non-accidental deaths and deaths from cardiovascular
                               disease (CVD), cerebrovascular (CrD), and chronic
                               obstructive lung disease and associated conditions (COPD)
                               were analyzed by generalized additive Poisson models in
                               relation to 24-h readings for each of the air pollutants
                               averaged over all monitors in each county. All models
                               included an intercept term for day-of-week and a spline
                               smoother for temporal trends.  Effects of weather were first
                               evaluated by regressing daily deaths (for each mortality
                               endpoint) against temp and rel. humidity with lag times of 0
                               to 5 days. Then lags that minimized deviance for temp and
                               rel. humidity were kept fixed for subsequent pollutant effect
                               analyses. Each pollutant entered linearly into the regression
                               and lags of between 0 to 5 days examined. Effects of two or
                               more pollutants were then evaluated in multipollutant
                               models.  Sensitivity analyses were used to evaluate effect of
                               degree of smoothing on results.
                                                     In general, the gases, especially CO (but not O3) were much
                                                     more strongly associated with mortality than PM.  Specified
                                                     pattern of results found  for each county were as follows.  For
                                                     Cook Co., in single pollutant analyses PM10, CO, and O3 were all
                                                     associated (PM10 most strongly on lag 0-2 days) with total
                                                     mortality, as were SO2 and NO2 (strongest association on  lag 1
                                                     day for the latter two). In joint analyses  with one of gases, the
                                                     coefficients for both PM10 and the gas were somewhat
                                                     attenuated, but remained stat. sig. for some lags. With
                                                     3-pollutant models, PM10 coefficient became small and non-sig.
                                                     (except at lag 0), whereas the gases  dominated. For Los
                                                     Angeles,  PM10, PM2 5, CO, NO2, and SO2, (but not O3), were all
                                                     associated with total mortality. In joint analyses with CO or SO2
                                                     and either PM10 or PM2 5, PM metrics were markedly reduced
                                                     and non-sig., whereas estimates for  gases remained robust. In
                                                     Maricopa Co. single-pollutant analyses,  PM10 and each of the
                                                     gases, (except O3), were associated with  total morality; in
                                                     2-pollutant models, coefficients for  CO,  NO2, SO2, were more
                                                     robust than for PM10.  Analogous patterns of more robust
                                                     gaseous pollutant effects were generally  found for cause-specific
                                                     (CVD, CrD, COPD) mortality analyses.  Author concluded that
                                                     while direct effect of individual components of air pollution
                                                     cannot be ruled out, individual components best thought of as
                                                     indices of overall pollutant mix.
                                                      In single pollutant models, estimated
                                                      daily total mortality % excess deaths per
                                                      50 /2g/m3 PM10 was mainly in range of:
                                                      0.5-1.0% lags 0-2 Cook Co.; 0.25-1.0%
                                                      lags 0-2 LA; 2.0% lag 2 Maricopa.
                                                      Percent per 25 ^g/m3 PM25 0.5% lags 0,
                                                      1 for Los Angeles.

                                                      Maximum estimated COPD % excess
                                                      deaths (95% CI) per 50 ,ug/m3 PM10:
                                                      Cook Co. 5.4  (0.3,10.7), lag 2; with O3,
                                                      3.0 (-1.8, 8.1) lag 2; LA 5.9 (-1.6,
                                                      14.0)  lag 1; Maricopa 8.2 (-4.2, 22.3)
                                                      lag 1; per 25 ,ug PM25 in LA 2.7 (-3.4,
                                                      9.1).

                                                      CVD  % per 50 ,ug/m3  PM10:
                                                      Cook 2.2 (0.4, 4.1) lag 3; with O3, SO2
                                                      1.99 (-0.06, 4.1) lag 3; LA 4.5 (1.7,
                                                      7.4) lag 2; with CO -0.56 (-3.8, 2.8)
                                                      lag 2; Maricopa 8.9 (2.7,  15.4) lag 1;
                                                      with NO2 7.4 (-0.95, 16.3) lag 1.
                                                      Percent per 25 ,ug/m3 PM2 5, LA 2.6
                                                      (0.4, 4.9) lagl; with CO 0^60 (-2.1,
                                                      3.4).

                                                      CrD % per 50 ,ug/m3 PM10:
                                                      Cook 3.3 (-0.12, 6.8) lag 2; LA 2.9
                                                      (-2.3, 8.4) lag 3;  Maricopa 11.1  (0.54,
                                                      22.8)  lag 5. Percent per 25 Mg/m3
                                                      PM25, LA 3.6 (-0.6, 7.9) lag 3.
 o

-------
                   TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
 to
 o
 o
 to
Reference, Location, Years,
PM Index, Mean or Median,
IQRin,ug/m3.
  Study Description: Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates. Modeling
        methods: lags, smoothing, and covariates.
                 Results and Comments.
     Design Issues, Uncertainties, Quantitative Outcomes.
     PM Index, lag, Excess Risk%
   (95% LCL, UCL), Co-pollutants.
 OO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          United States (cont'd)

          Ostroetal. (1999a).
          Coachella Valley, CA.
          1989-1992. PM10
          (beta-attenuation)
          Mean = 56.8 ,ug/m3.
Ostro et al. (2000).
Coachella Valley, CA.
1989-1998.
PM25= 16.8;
PM10.25 = 25.8inIndio;
PM25 = 12.7;
PM10.25 = 17.9 in Palm Springs.
Study evaluated total, respiratory, cardiovascular, non-
cardiorespiratory and age >50 yr deaths (mean = 5.4, 0.6,
1.8, 3.0, and 4.8 per day, respectively). The valley is a
desert area where 50-60% of PM10 estimated to be coarse
particles. Correlation between gravimetric and beta-
attenuation, separated by 25 miles, was high (r = 0.93).
Beta-attenuation data were used for analysis. GAM Poisson
models adjusting for temperature, humidity, day-of-week,
season, and time were used. Seasonally stratified analyses
were also conducted. Lags 0-3 days (separately) of PM10
along with moving averages of 3 and 5 days examined, as
were O3, NO2, and CO.

A follow-up study of the Coachella Valley data, with PM2 5
and PM10.2 5 data in the last 2.5 years.  Both PM2 5 and PM10.
2 5 were estimated for the remaining years to increase power
of analyses.
                                                                                     Associations were found between 2- or 3-day lagged PM10 and
                                                                                     all mortality categories examined, except non-cardiorespiratory
                                                                                     series.  The effect size estimates for total and cardiovascular
                                                                                     deaths were larger for warm season (May through October) than
                                                                                     for all year period. NO2 and CO were significant predictor of
                                                                                     mortality in single pollutant models, but in multi-pollutant
                                                                                     models, none of the gaseous pollutants were significant
                                                                                     (coefficients reduced), whereas PM10 coefficients remained the
                                                                                     same and significant.
Several pollutants were associated with all-cause mortality,
including PM2 5, CO, and NO2. More consistent results were
found for cardiovascular mortality, for which significant
associations were found for PM10_2 5 and PM10, but not PM2 5
(possibly due to low range of PM2 5 concentrations and reduced
sample size for PM25 data).
                                                          Total mortality percent excess deaths
                                                          per 50 f2g/ m3 PM10 at 2-day lag= 4.6
                                                          (0.6, 8.8).

                                                          Cardiac deaths:
                                                          8.33(2.14, 14.9)

                                                          Respiratory deaths:
                                                          13.9(3.25,25.6)
Total percent excess deaths:
PM10 = 2.0 (-1.0, 5.1) per 50 //g/m3
PM25= 11.5(0.2, 24.1) per 25 |/g/m3
PMio-2.5 = 1.3 (-0.6, 3.5) per 25 ^g/m3

Cardio deaths:
PM10 = 6.1 (2.0, 10.3) per 50 ,ug/m3
PM25 = 8.6 (-6.4, 25.8) per 25 ^g
PM10.25 = 2.6 (0.7, 4.5) per 25

Respiratory deaths:
PM10= -2.0 (-11.4, 8.4) per 50
PM25= 13.3 (-43.1,32.1) per 25 Mg/
PM10.25 = -1.3 (-6.2, 4.0) per 25 ,ug/m3
O

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
Study Description: Outcomes, Mean outcome rate, and
ages. Concentration measures or estimates. Modeling
     methods: lags, smoothing, and covariates.
             Results and Comments.
Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                               PM Index, lag, Excess Risk%
                                                                                                                                                              (95% LCL, UCL), Co-pollutants.
OO
>
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          United States (cont'd)

          Fairley(1999).
          Santa Clara County, CA
          1989-1996.
          PM25(13);PM10(34);
          PM10.25(ll);COH(0.5unit);
          NO3(3.0);SO4(1.8)
          Schwartz etal. (1999).
          Spokane, WA
          1989-1995
          PM10: "control" days:
          42 Mg/m3;
          dust-storm days: 263

          Pope etal.  (1999a).
          Ogden, Salt Lake City, and
          Provo/Orem, UT
          1985-1995
          PM10 (32 for Ogden;
          41 for SLC; 38 for P/O)
                              Total, cardiovascular, and respiratory deaths were regressed
                              on PM10, PM2 5, PM10.2 5, COH, nitrate, sulfate, O3, CO, NO2,
                              adjusting for trend, season, and min and max temperature,
                              using Poisson GAM model. Season-specific analysis was
                              also conducted. The same approach was also used to re-
                              analyze 1980-1986 data (previously analyzed by Fairley,
                              1990).
                              Effects of high concentration of coarse crustal particles were
                              investigated by comparing death counts on 17 dust storm
                              episodes to those on non-episode days on the same day of
                              the years in other years, adjusting for temperature,
                              dewpoint, and day-of-week, using Poisson regression.
                              Associations between PM10 and total, cardiovascular, and
                              respiratory deaths studied in three urban areas in Utah's
                              Wasatch Front, using Poisson GAM model and adjusting for
                              seasonality, temperature, humidity, and barometric pressure.
                              Analysis was conducted with or without dust (crustal coarse
                              particles) storm episodes, as identified on the high "clearing
                              index" days, an index of air stagnation.
                                                     PM2 5 and nitrate were most significantly associated with
                                                     mortality, but all the pollutants (except PM10_2 5) were
                                                     significantly associated in single poll, models.  In 2 and 4 poll.
                                                     models with PM2 5 or nitrate, other pollutants were not
                                                     significant.  The RRs for respiratory deaths were always larger
                                                     than those for total or cardiovascular deaths. The difference in
                                                     risk between season was not significant for PM2 5.  The 1980-
                                                     1986 results were similar, except that COH was very
                                                     significantly associated with mortality.
                                                     No association was found between the mortality and dust storm
                                                     days on the same day or the following day.
                                                     Salt Lake City (SLC), where past studies reported little PM10-
                                                     mortality associations, had substantially more dust storm
                                                     episodes. When the dust storm days were screened out from
                                                     analysis and PM10 data from multiple monitors were used,
                                                     comparable RRs were estimated for SLC and Provo/Orem (P/O).
                                                      Total mortality per 25 ,ug/m3 PM2 5 at 0
                                                      d lag:  8% in one pollutant model;
                                                      9-12
                                                               '. pollutant model; 12
                                                      4-pollutant model.  Also, 8% per
                                                      50 f/g/m3 PM10 in one pollutant model
                                                      and 2% per 25 ,ug/m3 PM10.25.

                                                      Cardiovascular mortality:
                                                      PM10 = 9% per 50 ^g/m3
                                                      PM25= 13%per25,ug/m3
                                                      PM10.25 = 3% per 25 ^g/m3

                                                      Respiratory mortality:
                                                      PM10 = 1 1% per 50 ^g/m3
                                                      PMio-2.5 = 16% per 25 ,ug/m3

                                                      0% (-4.5, 4.7) for dust storm days at 0
                                                      day lag (50 ,ug/m3 PM10) (lagged days
                                                      also reported to have no associations).
                                                      Ogden PM10
                                                      Total (0 d) = 12.0% (4.5, 20.1)
                                                      CVD(0-4 d) = 1.4% (-8.3, 12.2)
                                                      Resp. (0-4 d) = 23.8 (2.8, 49.1)
                                                      SLC PM10
                                                      Total (0 d) = 2.3% (0.47)
                                                      CVD(0-4d) = 6.5%(2.2, 11.0)
                                                      Resp. (0-4 d) = 8.2 (2.4, 15.2)

                                                      Provo/Ovem PM10
                                                      Total (Od)= 1.9% (-2.1,6.0)
                                                      CVD (0-4 d) = 8.6% (2.4, 15.2)
                                                      Resp. (0-4 d) = 2.2% (-9.8, 15.9)
                                                      Note: Above % for PM2.5 and PM10.25
                                                      all per 25 ,ug/m3; all PM10 % per
                                                      50

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
           Reference, Location, Years,
           PM Index, Mean or Median,
  Study Description:  Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates. Modeling
        methods:  lags, smoothing, and covariates.
                  Results and Comments.
     Design Issues, Uncertainties, Quantitative Outcomes.
     PM Index, lag, Excess Risk%
(95% LCL, UCL), Co-pollutants.
OO
>
OO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          United States (cont'd)

          Schwartz and Zanobetti (2000).
          Chicago 1988-1993.
          PM10. Median = 36 ,ug/m3.
           Lippmann et al. (2000).
           Detroit, MI. 1992-1994.
           PM10 = 31;
           PM25 = 18;
           PM10.25=13.
          For 1985-1990 period
          TSP, PM10, TSP-PM10,
          Sulfate from TSP (TSP-SO4')
Total (non-accidental), in-hospital, out-of-hospital deaths
(median = 132, 79, and 53 per day, respectively), as well as
heart disease, COPD, and pneumonia elderly hospital
admissions (115, 7, and 25 per day, respectively) were
analyzed to investigate possible "harvesting" effect of PM10.
GAM Poisson models adjusting for temperature, relative
humidity, day-of-week, and season were applied in baseline
models using the average of the same day and  previous
day's PM10.  The seasonal and trend decomposition
techniques called STL was applied to the health outcome
and exposure data to decompose them into different time-
scales (i.e, short-term to long-term), excluding the long,
seasonal cycles (120 day window). The associations were
examined with smoothing windows of 15, 30,  45, and 60
days.

For 1992-1994 study period, total (non-accidental),
cardiovascular, respiratory, and other deaths were analyzed
using GAM Poisson models, adjusting for season,
temperature, and relative humidity.  The air pollution
variables analyzed were: PM10, PM2 5, PM10_2 5, sulfate, H+,
O3, SO2, NO2, and CO.

For earlier 1985-1990 study period, total non-accidental,
circulatory, respiratory, and "other" (non-circulatory or
respiratory non-accidental) mortality were evaluated versus
noted PM indices and gaseous pollutants.
                                                                                                 The effect size estimate for deaths outside of the hospital is
                                                                                                 larger than for deaths inside the hospital.  All cause mortality
                                                                                                 shows an increase in effect size at longer time scales. The effect
                                                                                                 size for deaths outside of hospital increases more steeply with
                                                                                                 increasing time scale than the effect size for deaths inside of
                                                                                                 hospitals.
PM10, PM2 5, and PM10_2 5 were more significantly associated with
mortality outcomes than sulfate or H+. PM coefficients were
generally not sensitive to inclusion of gaseous pollutants. PM10,
PM2 5, and PM10_2 5 effect size estimates were comparable per
same distributional increment (5th to 95th percentile).

Both PM10 (lag 1 and 2 day) and TSP (lag 1 day) but not TSP-
PM10 or TSP-SO4" significantly associated with respiratory
mortality for 1985-1990 period.  The simultaneous inclusions of
gaseous pollutants with PM10 or TSP reduced PM effect size by 0
to 34%. Effect size estimates for total, circulatory, and "other"
categories were smaller than for respiratory mortality.
                                                           Mortality RR estimates per 50 ,ug/m3
                                                           increase of mean of lag 0- and 1-days
                                                           PM10: total deaths 4.5 (3.1, 6.0);
                                                           in-hospital 3.9 (2.1, 5.8); out-of-
                                                           hospital 6.3 (4.1, 8.6). For total deaths,
                                                           the RR approximately doubles as the
                                                           time scale changes  from 15 days to 60
                                                           days. For out-of-hospital deaths,  it
                                                           triples from 15 days to 60 days time
                                                           scale.
Total mortality percent excess deaths:
PM10(1 d) = 4.4%(-1.0, 10.1)
PM25(Od) = 3.1%(-0.6, 7.0)
PM10.25 (1 d) = 4.0% (-1.2, 9.4)
PM10(1 d) = 6.9%(-1.3, 15.7)
PM25(1 d) = 3.2%(-2.3, 8.9)
PM10.25 (1 d) = 7.8% (0.0, 16.2)

Respiratory mortality:
PM10 (0 d) = 7.8% (-10.2, 29.5)

Circulatory mortality:
PM25(od) = 2.3%(-10.3, 16,6)
PM10.25 (2 d) = 7.4% (-9.1, 26.9)
Note:  All above PM10 per 50 Mg/m3; all
PM25 and PM10.25 % per 25 //g/m3.
O

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description:  Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates. Modeling
        methods:  lags, smoothing, and covariates.
                                                                                                                   Results and Comments.
                                                                                                      Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                       PM Index, lag, Excess Risk%
                                                                                                                                                     (95% LCL, UCL), Co-pollutants.
OO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          United States (cont'd)

          Chock et al. (2000).
          1989-1991
          Pittsburgh, PA
          PM10 (daily)
          PM2 5 (every 2 days)
Klemm and Mason (2000).
Atlanta, GA
1998-1999
PM25mean=19.9;
PM2yPM10 =0.65.
Nitrate, EC, OC, and
oxygenated HC.

Gwynn et al. (2000).
Buffalo, N.Y. 1988-1990.
PM10 (24); COH (0.2 /1000ft);
SO4= (62 nmoles/m3)
Schwartz (2000c).
Boston, MA.
1979-1986.
PM25mean= 15.6.
                               Study evaluated associations between daily mortality and
                               several air pollution variables (PM10, PM2 5, CO, O3, NO2,
                               SO2) in two age groups (<75 yr., > 75 yr.) in Pittsburgh, PA,
                               during 3-yr. period. Poisson regression used, including
                               filtering of data based on cubic B-spline basis functions,
                               with adjustments for seasonal trends, day-of-week effects,
                               temp., dew point. Single- and multi-pollutant models run
                               for 0, 1, 2, and 3 day lags.  PM25/PM10 * 0.67.
Reported "interim" results for 1 yr period of observations
regarding total mortality in Atlanta, GA during 1998-1999.
Generalized additive model used to assess effects of PM25
vs PM10_2 5, and for nitrate, EC, OC and oxygenated HC
components.
                                         Total, circulatory, and respiratory mortality and unscheduled
                                         hospital admissions were analyzed for their associations
                                         with H+, SO4=, PM10, COH, O3, CO, SO2, and NO2,
                                         adjusting for seasonal cycles, day-of-week, temperature,
                                         humidity, using. Poisson and negative binomial GAM
                                         models.

                                         Non-accidental total, pneumonia,  COPD, and ischemic
                                         heart disease mortality were examined for possible
                                         "harvesting" effects of PM. The mortality, air pollution, and
                                         weather time-series were separated into seasonal cycles
                                         (longer than 2-month period), midscale, and short-term
                                         fluctuations using STL algorithm. Four different midscale
                                         components were used (15, 30, 45, and 60  days) to examine
                                         the extent of harvesting.  GAM Poisson regression analysis
                                         was performed using deaths, pollution, and weather for each
                                         of the four midscale periods.
                                                                                                 Issues of seasonal dependence of correlation among pollutants,
                                                                                                 multi-collinearity among pollutants, and instability of
                                                                                                 coefficients emphasized. Single- and multi-pollutant non-
                                                                                                 seasonal models show significant positive association between
                                                                                                 PM10 and daily mortality, but seasonal models showed much
                                                                                                 multi-collinearity, masking association of any pollutant with
                                                                                                 mortality.  Also, based on data set half the size for PM10, the
                                                                                                 PM2 5 coefficients were highly unstable and, since no
                                                                                                 consistently significant associations found in this small data set
                                                                                                 stratified by age group and season, no conclusions drawn on
                                                                                                 relative role of PM25 vs. PM10_25.

                                                                                                 No significant associations were found for any of the pollutants
                                                                                                 examined, possibly due to a relatively short study period (1-
                                                                                                 year). The coefficient and t-ratio were larger for PM2 5 than for
                                                                                                 PM10.2.5.
                                                        For total mortality, all the PM components were significantly
                                                        associated, with H+ being the most significant, and COH the
                                                        least significant predictors.  The gaseous pollutants were mostly
                                                        weakly associated with total mortality.
                                                                                       For COPD deaths, the results suggest that most of the mortality
                                                                                       was displaced by only a few months.  For pneumonia, ischemic
                                                                                       heart disease, and total mortality, the effect size increased with
                                                                                       longer time scales.
                                                                                                                   Total mortality percent increase per
                                                                                                                   25 ,ug/m3 for aged <75 yrs:
                                                                                                                   PM2 5 = 2.6% (2.0, 7.3)
                                                                                                                   PM10.25 = 0.7%(-1.7, 3.7)

                                                                                                                   Total mortality percent increase per 25
                                                                                                                   ,ug/m3 for aged >75 yrs:
                                                                                                                   PM25= 1.5%(-3.0, 6.3)
                                                                                                                   PM10.25 = 1.3%(-1.3, 3.8)
                                                                                                                                                            Total mortality percent increase per 25
                                                                                                                                                            ,ug/m3 for:
                                                                                                                                                            PM25 = 4.8%(-3.2, 13.4)
                                                                                                                                                            PM10.2.5 = 1.4% (-11.3, 15.9)
                                                                                                                                                             12% (2.6, 22.7) per 50 ,ug/m3 PM10 at 2-
                                                                                                                                                             day lag.
                                                                                                                   Total mortality percent increase per
                                                                                                                   25 ,ug/m3 increase in PM25:
                                                                                                                   5.3 (1.8, 9.0) for short-term
                                                                                                                   fluctuations; 9.6 (8.1, 11.1) for 1he 60
                                                                                                                   day window.
O

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description: Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates.  Modeling
        methods: lags, smoothing, and covariates.
                   Results and Comments.
     Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                                    PM Index, lag, Excess Risk%
                                                                                                                                                                   (95% LCL, UCL), Co-pollutants.
OO
>
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          United States (cont'd)

          Lipfert et al. (2000a).
          Philadelphia (7 county
          Metropolitan area), 1992-1995.
          Harvard PM measurements:
          PM25(17.3);PM10(24.1);
          PM10.2.5 (6.8),
          sulfate (53.1 nmol/m3);
          H+(8.0nmol/m3).
Laden et. al. (2000)
Six Cities (means):
Watertown, MA (16.5);
Kingston-Harriman, TN (21.1);
St. Louis, MO (19.2);
Steubenville, OH (30.5);
Portage, WI (11.3); Topeka, KS
(12.2). 1979-1988?. 15 trace
elements in the dichot PM2 5:
Si, S, Cl, K, Ca, V, Mn, Al, Ni,
Zn, Se, Br, Pb, Cu, and Fe.
          Levy (1998).
          King County, WA.
          1990-1994.
          PM10 Nephelometer (30);
          (0.59bspunit)
12 mortality variables, as categorized by area, age, and
cause, were regressed on 29 pollution variables (PM
components, O3, SO2, NO2, CO, and by sub-areas), yielding
348 regression results. Both dependent and explanatory
variables were pre-filtered using the!9-day-weighted
average filter prior to OLS regression.  Covariates were
selected from filtered temperature (several lagged and
averaged values), indicator variables for hot and cold days
and day-of-week using stepwise procedure. The average of
current and previous days' pollution levels were used.

Total (non-accidental), ischemic heart disease, pneumonia,
and COPD (mean daily total deaths for the six cities: 59,
12, 55, 3, 11, and  3, respectively in the order  shown left). A
factor analysis was conducted on the 15 elements in the fine
fraction of dichot samplers to obtain five common factors;
factors were rotated to maximize the projection of the single
"tracer" element (as in part identified from the past studies
conducted on these data) for each  factor; PM2 5 was
regressed on the identified factors scores so that the factor
scores could be expressed in the mass scale. Using GAM
Poisson models adjusting for temperature, humidity, day-of-
week, season, and time, mortality  was regressed on the
factor scores in the mass scale. The mean of the same-day
and previous day (increasing the sample size from 6,211 to
9,108 days) mass values were used.  The city-specific
regression coefficients were combined using inverse
variance weights.

Out-of-hospital deaths (total, respiratory, COPD, ischemic
heart disease, heart failure, sudden cardiac death screening
codes, and stroke) were related to  PM10, nephelometer (0.2 -
1.0 /an fine particles), SO2, and CO, adjusting for day-of-
week, month of the year, temperature and dewpoint, using
Poisson regression.
                                                                                        Significant associations were found for a wide variety gaseous
                                                                                        and parti culate pollutants, especially for peak O3. No systematic
                                                                                        differences were seen according particle size or chemistry.
                                                                                        Mortality for one part of the metropolitan area could be
                                                                                        associated with air quality from another, not necessarily
                                                                                        neighboring part.
Three sources of fine particles were defined in all six cities with
a representative element for each source type:  Si for soil and
crustal material; Pb for motor vehicle exhaust; and Se for coal
combustion sources. In city-specific analysis, additional sources
(V for fuel oil combustion, Cl for salt, etc.) were considered.
Five source factors were considered for each city, except Topeka
with the three sources. Coal and mobile sources account for the
majority of fine particles in each city. In all of the metropolitan
areas combined, 46% of the total fine particle mass was
attributed to coal combustion and 19% to mobile sources. The
strongest increase in daily mortality was associated with the
mobile source factor. The coal combustion factor was positively
associated with mortality in all metropolitan areas, with the
exception of Topeka. The crustal factor from the fine particles
was not associated with mortality.
                                                                                        Nephelometer data were not associated with mortality.  Cause-
                                                                                        specific death analyses suggest PM associations with ischemic
                                                                                        heart disease deaths.  Associations of mortality with SO2 and CO
                                                                                        not mentioned.  Mean daily death counts were small (e.g., 7.7 for
                                                                                        total; 1.6 for ischemic heart disease). This is an apparently
                                                                                        preliminary analysis.
                                                            The fractional Philadelphia mortality
                                                            risk attributed to the pollutant levels:
                                                            "average risk" was 0.0423 for 25 //g/m3
                                                            PM25; 0.0517 for 25 ,ug/m3 PM10.25;
                                                            0.0609 for 50 //g/m3 PM10, using the
                                                            Harvard PM indices at avg. of 0 and 1 d
                                                            lags-
                                                                                                                                                                Total mortality percent excess overall:
                                                                                                                                                                4.0 (2.8, 5.3), 2.7 (0.6, 5.0) with each
                                                                                                                                                                25 /ig/m3 increase in the two-day mean
                                                                                                                                                                of coal combustion fine PM factor; 8.7
                                                                                                                                                                (4.2,  13.4) with each 25 //g/m3 increase
                                                                                                                                                                in the two-day mean of mobile source
                                                                                                                                                                fine PM factor; -5.7 (-13.7, 3.2) with
                                                                                                                                                                each 25 ,ug/m3 increase in the two-day
                                                                                                                                                                mean of the crustal source fine PM
                                                                                                                                                                factor.
                                                            Total mortality percent excess:
                                                            5.6% (-2.4, 1.43) per 50 //g/m3 PM10 at
                                                            avg. of 2 to 4 d lag; 7.2% (-6.3, 22.8)
                                                            with SO2 CO.  1.8% (-3.5, 7.3) per
                                                            25 Mg/m3 PMj; -1.0 (-8.7,. 7.7) with
                                                            SO, and CO.
o

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
 to
 o
 o
 to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description: Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates. Modeling
        methods:  lags, smoothing, and covariates.
                  Results and Comments.
     Design Issues, Uncertainties, Quantitative Outcomes.
     PM Index, lag, Excess Risk%
   (95% LCL, UCL), Co-pollutants.
 OO
 >
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          United States (cont'd)

          Mar et al. (2000).
          Phoenix, AZ. 1995-1997.
          PM10, PM25, andPM10.25
          (TEOM), with means = 46.5,
          13.0, and 33.5, respectively;
          and PM2 5 (DFPSS),
          mean= 12.0.
Clyde et al. (2000).
Phoenix, AZ. 1995-1998.
PM10, and PM25, (from
TEOM), with means = 45.4,
and 13.8. PM10_25 computed
as PM10-PM25.
Total (non-accidental) and cardiovascular deaths (mean =
8.6 and 3.9, respectively) for only those who resided in the
zip codes located near the air pollution monitor were
included. GAM Poisson models were used, adjusting for
season, temperature, and relative humidity. Air pollution
variables evaluated included: O3, SO2, NO2, CO, TEOM
PM10, TEOM PM25, TEOM PM10.25, DFPSS PM25, S, Zn,
Pb, soil, soil-corrected K (KS), nonsoil PM, OC, EC, and
TC. Lags 0 to 4 days evaluated. Factor analysis also
conducted on chemical components of DFPSS PM25 (Al, Si,
S, Ca, Fe, Zn, Mn, Pb, Br, KS, OC, and EC); and factor
scores included in mortality regression.

Elderly (age > 65 years) non-accidental mortality for three
regions of increasing size in Phoenix urban area analyzed to
evaluate influence of spatial uniformity of PM10 and PM25.
All-age accidental deaths for the metropolitan area also
examined as a "control". GAM Poisson models adjusting
for season (smoothing splines of days), temperature, specific
humidity, and lags 0- to 3-d of weather variables. PM
indices for lags 0-3 d considered.  Bayesian Model
Averaging (BMA) produces posterior mean relative risks by
weighting each model (out of all possible model
specifications examined) based on support received from the
data.
                                                                                      Total mortality was significantly associated with CO and NO2
                                                                                      and weakly associated with SO2, PM10, PM10_2 5, and EC.
                                                                                      Cardiovascular mortality was significantly associated with CO,
                                                                                      NO2, SO2, PM25, PM10, PM10.25, OC and EC.  Combustion-
                                                                                      related factors and secondary aerosol factors were also associated
                                                                                      with cardiovascular mortality. Soil-related factors, as well
                                                                                      as individual variables that are associated with soil were
                                                                                      negatively associated with total mortality.
The BMA results suggest that a weak association was found only
for the mortality variable defined over the region with uniform
PM2 5, with a 0.91 probability that RR is greater than 1.  The
other elderly mortality variables, including the accidental deaths
("control"), had such probabilities in the range between  0.46 to
0.77. Within the results for the mortality defined over the region
with uniform PM2 5, the results suggested that effect was
primarily due to coarse particles rather than fine; only the lag  1
coarse PM was consistently included in the highly ranked
models.
                                                           Total mortality percent excess: 5.4 (0.1,
                                                           11.1) for PM10 (TEOM) 50 ,ug/m3 at lag
                                                           0  d; 3.0 (-0.5, 6.6) for PM10.25 (TEOM)
                                                           25 Mg/m3 at lag 0 d; 3.0 (-0.7, 6.9) for
                                                           PM2 5 (TEOM) 25 Mg/m3 at lag 0 d.
                                                           Cardiovascular mortality RRs: 9.9(1.9,
                                                           18.4) for PM10 (TEOM) 50 ^g/m3 at lag
                                                           0  d; 18.7 (5.7, 33.2) for PM25 (TEOM)
                                                           25 |/g/m3 at lag 1 d; and 6.4 (1.4, 11.7)
                                                           PM10 (TEOM) 25 |/g/m3 PM10.2 5 at lag 0
                                                           d.
Posterior mean RRs and 90%
probability intervals per changes of
25 |/g/m3 in all lags of fine and coarse
PM for elderly mortality for uniform
PM10 region: 1.06(1+, 1.11).
O

-------
                   TABLE 8A-1  (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
 to
 o
 o
 to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description: Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates.  Modeling
        methods: lags, smoothing, and covariates.
                   Results and Comments.
     Design Issues, Uncertainties, Quantitative Outcomes.
     PM Index, lag, Excess Risk%
   (95% LCL, UCL), Co-pollutants.
 OO
 to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           United States (cont'd)

           Smith et al. (2000).
           Phoenix, AZ.
           1995-1997
Tsai et al. (2000).
Newark, Elizabeth, and
Camden, NJ. 1981-1983.
PM15: 55.5, 47.0, 47.5; and
PM25:42.1, 37.1, 39.9, for
Newark, Elizabeth, and
Camden, respectively.
Study evaluated effects of daily and 2- to 5-day average
coarse (PM10_2 5) and fine (PM2 5) particles from an
EPA-operated central monitoring site on nonaccidental
mortality among elderly (65+ years), using time-series
analyses for residents within city of Phoenix and, separately,
for region of circa 50 mi around Phoenix. Initial model
selected to represent long-term trends and weather variables
(e.g., ave. daily temp., max daily temp., daily mean specific
humidity, etc.); then PM variables added to model one at a
time to ascertain which had strongest effect. Piecewise
linear analysis and spline analysis used to evaluate possible
nonlinear PM-mortality relationship and to evaluate
threshold possibilities. Data analyzed most likely same as
Clyde's or Mar's Phoenix data.

Factor analysis-derived source type components were
examined for their associations with mortality in this study.
Non-accidental total deaths and cardiorespiratory deaths
were examined for their associations with PM15, PM2 5
sulfate, trace metals from PM15, three fractions of
extractable organic matter, and CO. Data were analyzed
with Poisson GEE regression models with autoregressive
correlation  structure, adjusting for temperature, time-of-
week, and season indicator variables. Individual pollution
lag days from 0 to 3, as well as the average concentrations
of current and preceding 3 days were considered.  Factor
analysis of the trace elements, sulfate, and CO data was
conducted,  and mortality series were regressed on these
factor scores.
                                                                                         In linear PM effect model, a statistically significant mortality
                                                                                         association found with PM10_2 5, but not with PM2 5. In the model
                                                                                         allowing for a threshold, evidence suggestive of possible
                                                                                         threshold for PM2 5 (in the range of 20-25 ,ug/m3) found, but not
                                                                                         for PM10_25. A seasonal interaction in the PM10_25 effect was also
                                                                                         reported:  the effect being highest in spring and summer when
                                                                                         anthropogenic concentration of PM10_25 is lowest.
Factor analysis identified several source types with tracer
elements.  In Newark, oil burning factor, industrial source factor,
and sulfate factor were positively associated with total mortality;
and sulfate was associated with cardio-respiratory mortality.  In
Camden, oil burning and motor vehicle factors were positively
associated with total mortality; and, oil burning, motor vehicles,
and sulfate were associated with cardio-respiratory mortality.  In
Elizabeth, resuspended dust was not associated with total
mortality; and industrial source (traced by Cd) showed positive
associations with cardio-respiratory mortality. On the mass basis
(source-contributed mass), the RRs estimates per 10 i/g/m3 were
larger for specific sources (e.g., oil burning, industry, etc.) than
for total mass.  The choice of lag/averaging reported to be not
important.
Percent excess deaths per 50 ,ug/m3
increase in current day PM15: in
Newark, 5.7 (4.6, 6.7) for total
mortality,  7.8 (3.6,  12.1) for cardioresp.
mortality;  in Camden, 11.1 (0.7, 22.5)
and 15.0 (4.3, 26.9); and in Elizabeth,
-4.9 (-17.9, 10.9) and 3.0  (-11.0,
19.4), respectively. Percent excess
deaths per 25 ,ug/m3 PM25; in Newark,
4.3 (2.8, 5.9) for total and 5.1 (3.1, 7.2)
for cardiorespiratory mortality; in
Camden, 5.7(0.1, 11.5) and 6.2 (0.6,
12.1); in Elizabeth, 1.8 (-5.4, 9.5) and
2.3 (-5.0,  10.1), respectively.
 O

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
Study Description: Outcomes, Mean outcome rate, and
ages. Concentration measures or estimates.  Modeling
      methods:  lags, smoothing, and covariates.
             Results and Comments.
Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                                  PM Index, lag, Excess Risk%
                                                                                                                                                                (95% LCL, UCL), Co-pollutants.
OO
>
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          United States (cont'd)

          Gamble (1998).
          Dallas, TX. 1990-1994.
          PM10(25)
          Ostro (1995).
          San Bernardino and Riverside
          Counties, CA, 1980-1986.
          PM2 5 (estimated from visual
          range). Mean = 32.5.
          Kelsalletal. (1997).
          Philadelphia, PA
          1974-1988.
          TSP (67)
          Moolgavkar and Luebeck
          (1996). Philadelphia, PA.
          1973-1988. TSP (68)
                               Relationships of total, respiratory, cardiovascular, cancer,
                               and remaining non-accidental deaths to PM10, O3, NO2, SO2,
                               and CO evaluated, adjusting for temperature, dewpoint, day-
                               of-week, and seasonal cycles (trigonometric terms) using
                               Poisson regression.

                               Study evaluated total, respiratory, cardiovascular, and age >
                               = 65 deaths (mean = 40.7, 3.8, 18.7, and 36.4 per day,
                               respectively). PM2 5 estimated based on airport visual range
                               and previously published empirical formula.  Autoregressive
                               OLS (for total) and Poisson (for sub-categories) regressions
                               used, adjusting for season (sine/cosine with cycles from  1 yr
                               to 0.75 mo; prefiltering with 15-day moving ave.;
                               dichotomous variables for each year and month; smooth
                               function of day and temp.),  day-of-week, temp, and
                               dewpoint.  Evaluated lags 0, 1, and 2 of estimated PM25, as
                               well as moving averages of 2, 3, and 4 days and O3.

                               Total, cardiovascular, respiratory, and by-age mortality
                               regressed on TSP, SO2, NO2, O3, and CO, adjusting for
                               temporal trends and weather, using Poisson GAM model.
                               A critical review paper, with an analysis of total daily
                               mortality for its association with TSP, SO2, NO2, and O3,
                               adjusting for temporal trends, temperature, and also
                               conducting analysis by season, using Poisson GAM model.
                                                      O3 (avg. of 1-2 day lags), NO2 (avg.. 4 -5 day lags), and CO
                                                      (avg. of lags 5- 6 days) were significantly positively associated
                                                      with total mortality. PM10 and SO2 were not significantly
                                                      associated with any deaths.
                                                      The results were dependent on season. No PM2 5 - mortality
                                                      association found for the full year-round period. Associations
                                                      between estimated PM2 5 (same-day) and total and respiratory
                                                      deaths found during summer quarters (April - Sept.).
                                                      Correlation between the estimated PM2 5 and daily max temp.
                                                      was low (r = 0.08) during the summer quarters. Ozone was also
                                                      associated with mortality, but was also relatively highly
                                                      correlated with temp, r = 0.73). Moving averages of PM25 did
                                                      not improve the associations.
                                                      TSP, SO2, O3, and 1-day lagged CO individually showed
                                                      statistically significant associations with total mortality.  No NO2
                                                      associations unless SO2 or TSP was also considered.  The effects
                                                      of TSP and SO2 were diminished when both pollutants were
                                                      included.

                                                      RR results presented as figures, and seasonal difference noted.
                                                      TSP, SO2, O3 - mortality associations varied across season. TSP
                                                      associations were stronger in summer and fall. NO2 was the most
                                                      significant predictor.
                                                       -3.6% (-12.7, 6.6) per 50 //g/m3 PM10
                                                       at 0 lag (other lags also reported to have
                                                       no associations)
                                                       Percent excess deaths per 25 ,ug/m3 of
                                                       estimated PM25, lag 0: Full year:  0.3
                                                       (-0.6, 1.2) for total; 2.1 (-0.3, 4.5) for
                                                       respiratory; and 0.7 (-0.3, 1.7) for
                                                       circulatory.  Summer quarters: 1.6
                                                       (0.03, 3.2) for total; 5.5 (1.1,  10.0) for
                                                       respiratory; and 0 (-1.0, 1.0) for
                                                       circulatory.
                                                       Total mortality excess risk: 3.2% (0,
                                                       6.1) per 100 ,ug/m3 TSP at 0 day lag.
                                                       Total mortality excess risk: ranged ~ 0
                                                       (winter) to =4% (summer) per
                                                       100 Mg/m3 TSP at 1 day lag.
o

-------
                  TABLE 8A-1  (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description: Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates. Modeling
        methods: lags, smoothing, and covariates.
                                                                                                                     Results and Comments.
                                                                                                       Design Issues, Uncertainties, Quantitative Outcomes.
  PM Index, lag, Excess Risk%
(95% LCL, UCL), Co-pollutants.
OO
>
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          United States (cont'd)

          Neas et. al. (1999).
          Philadelphia. 1973-1980.
          TSP mean = 77.2.
Schwartz (2000d).
Philadelphia. 1974-1988.
TSP. Mean = 70//g/m3 for
warm season (April through
August) and 64 //g/m3 for
cold season.
          Levy et al. (2000).
          Years vary from study to study
          ranging between 1973 to 1994.
          21 published studies included
          U.S., Canadian, Mexican,
          European, Australian, and
          Chilean cities. PM10 levels in
          the 19 U.S. cities (in some
          cases TSP were converted to
          PM10 using factor of 0.55)
          ranged from -20 to -60 ug/m3.
Total, age over 65, cancer, and cardiovascular deaths
analyzed for association with TSP.  Conditional logistic
regression analysis with case-crossover design conducted.
Average values of current and previous days' TSP used.
Case period is the 48-hr period ending at midnight on day of
death. Control periods are 7, 14, and 21 days before and
after the case period. Other covariates included temperature
on the previous day, dewpoint on the same day, an indicator
for hot days (> SOT), an indicator for humid days
(dewpoint > 66°F), and interaction of same-day temp, and
winter season.

Total (non-accidental) deaths analyzed.  GAM Poisson
models adjusting for temperature, dewpoint, day-of-week,
and season applied to each of 15 warm and cold seasons.
Humidity-corrected extinction coefficient, derived from
airport visual range, also considered as explanatory variable.
In the second stage, resulting 30 coefficients were regressed
on regression coefficients of TSP on SO2.  Results of first
stage analysis combined using inverse variance weighting.

To determine whether across-study heterogeneity of PM
effects could be explained by regional parameters, Levy
et al. applied an empirical Bayes meta-analysis to 29 PM
estimates from 21 published studies. They considered such
city-specific variables as mortality rate, gaseous pollutants^
regression coefficients, PM10 levels, central air conditioning
prevalence, heating and cooling degree days.
                                                                                        In each set of the six control periods, TSP was associated with
                                                                                        total mortality.  A model with four symmetric reference periods 7
                                                                                        and 14 days around the case period produced a similar result.  A
                                                                                        model with only two symmetric reference periods of 7 days
                                                                                        around the case produced a larger estimate.  A larger effect was
                                                                                        seen for deaths in persons > 65 years of age and for deaths due
                                                                                        to pneumonia and to cardiovascular disease. Cancer mortality
                                                                                        was not associated with TSP.
                                                                                                  When TSP controlled for, no significant association between SO2
                                                                                                  and daily deaths. SO2 had no association with daily mortality
                                                                                                  when it was poorly correlated with TSP. In contrast, when SO2
                                                                                                  was controlled for,  TSP was more strongly associated with
                                                                                                  mortality than when it was less correlated with SO2. However,
                                                                                                  all of the association between TSP and mortality was explained
                                                                                                  by its correlation with extinction coefficient.
                                                                                        Among the city-specific variables, PM2 5/PM10 ratio was a
                                                                                        significant predictor (larger PM estimates for higher PM2 5/PM10
                                                                                        ratios) in the 19 U.S. cities data subsets. While the sulfate data
                                                                                        were not available for all the 19 cities, the investigators noted
                                                                                        that, based on their analysis of the limited data with sulfate for
                                                                                        10 estimates, the sulfate/PMlO ratio was highly correlated with
                                                                                        both the mortality (r = 0.84) and with the PM25/PM10 ratio (r =
                                                                                        0.70). This indicates that the sulfate/PM10 ratio may be even
                                                                                        better predictor of regional heterogeneity of PM RR estimates.
                                                                                                                                                              Odds Ratio (OR) for all cause mortality
                                                                                                                                                              per 100 /ig/m3 increase in 48-hr mean
                                                                                                                                                              TSP was 1.056(1.027, 1.086). The
                                                                                                                                                              corresponding number for those aged 65
                                                                                                                                                              and over was 1.074(1.037, 1.111), and
                                                                                                                                                              1.063 (1.021, 1.107) for cardiovascular
                                                                                                                                                              disease.
                                                                                                                                                              Total mortality excess risk estimates
                                                                                                                                                              combined across seasons/years:  9.0
                                                                                                                                                              (5.7, 12.5) per 100//g/m3 TSP.
                                                                                                                                                              The pooled estimate froml9 U.S. cities
                                                                                                                                                              was 0.70% (0.54, 0.84) per 10 ug/m3
                                                                                                                                                              increase in PM10.
o

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description:  Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates. Modeling
        methods: lags, smoothing, and covariates.
                  Results and Comments.
     Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                                  PM Index, lag, Excess Risk%
                                                                                                                                                                 (95% LCL, UCL), Co-pollutants.
OO
 fe
 H
 6
 o

 o
 H
O
 O
 H
 W
 O
          Canada

          Burnett etal. (1998a).
          11 Canadian cities.
          1980-1991.
          No PM index data available
          on consistent daily basis.
Burnett et al. (2000).
8 largest Canadian cities.
1986-1996. All city mean
PM1025.9;PM25  13.3;
PM10.25 12.6; sulfate 2.6.
          Burnett etal. (1998b).
          Toronto, 1980-1994.
          TSP(60);COH(0.42);
          SO4= (9.2 Mg/m3);
          PM10 (30, estimated);
          PM25(18, estimated)
Total non-accidental deaths were linked to gaseous air
pollutants (NO2, O3, SO2, and CO) using GAM Poisson
models adjusting for seasonal cycles, day-of-week, and
weather (selected from spline-smoothed functions of
temperature, dewpoint, relative humidity with 0, 1, and 2
day lags using forward stepwise procedure).  Pollution
variables  evaluated at 0, 1, 2, and up to 3-day lag averages
thereof. No PM index included in analyses because daily
PM measurements not available.  City-specific models
containing all four gaseous pollutants examined. Overall
risks computed by averaging risks across cities.

Total non-accidental deaths linked to PM indices (PM10,
PM25, PM10_25,  sulfate, 47 elemental component
concentrations for  fine and coarse fractions) and gaseous air
pollutants (NO2, O3, SO2, and CO). Each city's mortality,
pollution, and weather variables separately filtered for
seasonal trends and day-of-week patterns.  The residual
series from all the cities then analyzed in a GAM Poisson
model. The weather model was selected from spline-
smoothed functions of temperature, relative humidity, and
maximum change in barometric pressure within a day, with
0 and 1 day lags using forward stepwise procedure.
Pollution  effects were examined at lags 0 through 5 days.
To avoid unstable parameter estimates in multi-pollutant
models, principal components were also used as predictors
in the regression models.

Total,  cardiac, and other nonaccidental deaths (and by age
groups) were regressed on TSP, COH, SO4=, CO, NO2, SO2,
O3, estimated PM10 and PM2 5 (based on the relationship
between the existing every-6th-day data and SO4=, TSP and
COH), adjusting for seasonal cycles, day-of-week,
temperature, and dewpoint using Poisson GAM model.
                                                                                       NO2 had 4.1% increased risk per mean concentration; O3 had
                                                                                       1.8%; SO2 had 1.4%, and CO had 0.9% in multiple pollutant
                                                                                       regression models. A 0.4% reduction in excess mortality was
                                                                                       attributed to achieving a sulfur content of gasoline of 30 ppm in
                                                                                       five Canadian cities. Daily PM data for fine and coarse mass
                                                                                       and sulfates available on varying (not daily) schedules allowed
                                                                                       ecologic comparison of gaseous pollutant risks by mean fine
                                                                                       particle indicators mass concentrations.
O3 was weakly correlated with other pollutants and other
pollutants were "moderately" correlated with each other (the
highest was r = 0.65 for NO2 and CO). The strongest association
with mortality for all pollutants considered were for 0 or 1 day
lags. PM25 was a stronger predictor of mortality than PM10_25.
The estimated gaseous pollutant effects were generally reduced
by inclusion of PM2 5 or PM10, but not PM10_2 5. Sulfate, Fe,  Ni,
and Zn were most strongly associated with mortality. Total
effect of these four components was greater than that for PM2 5
mass alone.
                                                                                       Essentially all pollutants were significant predictors of total
                                                                                       deaths in single pollutant models, but in two pollutant models
                                                                                       with CO, most pollutants' estimated RRs reduced (all PM
                                                                                       indices remained significant). Based on results from the co-
                                                                                       pollutant models and various stepwise regressions, authors noted
                                                                                       that effects of the complex mixture of air pollutants could be
                                                                                       almost completely explained by the levels of CO and TSP.
                                                            Found suggestion of weak negative
                                                            confounding of NO2 and SO2 effects
                                                            with fine particles and weak positive
                                                            confounding of particle effects with O3.
                                                            No quantitative RR or ER estimates
                                                            reported for PM indicators.
                                                                                                                                                             Percentage increase in daily filtered
                                                                                                                                                             non-accidental deaths associated with
                                                                                                                                                             increases of 50 ,ug/m3 PM10 and
                                                                                                                                                             25 |/g/m3 PM2 5 or PM10.2 5 at lag 1  day:
                                                                                                                                                             3.5 (1.0, 6.0) for PM10; 3^0 (1.1, 5.0) for
                                                                                                                                                             PM25; and 1.8 (-0.7, 4.4) for PM10.25.
                                                                                                                                                             In the multiple pollutant model with
                                                                                                                                                             PM2 5, PM10_2 5, and the 4 gaseous
                                                                                                                                                             pollutants, 1.9 (0.6, 3.2) for PM25; and
                                                                                                                                                             1.2 (-1.3, 3.8)forPM10.25.
                                                            Total mortality percent excess: 2.3%
                                                            (0.8, 3.8) per 100 ,ug/m3 TSP; 3.5%
                                                            (1..8, 5.3) per 50 ,ug/m3 PM10; 4.8%
                                                            (3.3, 6.4) per 25 ^g/m3 PM25. 0 day lag
                                                            for TSP and PM10;  Avg. of 0 and 1 day
                                                            forPM,,.
o

-------
                 TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description:  Outcomes, Mean outcome rate, and
  ages. Concentration measures or estimates.  Modeling
       methods: lags, smoothing, and covariates.
                 Results and Comments.
     Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                         PM Index, lag, Excess Risk%
                                                                                                                                                       (95% LCL, UCL), Co-pollutants.
OO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
         Canada (cont'd)
         Goldberg et al. (2000)
         Montreal, Quebec
         1984-95 Mean
         TSP=53.1
         (14.6-211.1)Mg/m3
         PM10 = 32.2
         (6.5- 120.5),ug/m3
         PM25 = 3.3 (0.0-30.0) ,ug/m3
Goldberg etal. (2001).
Montreal, Quebec.
1984-1993.  Predicted PM2
mean = 17.6. CoH
(1000ft)) mean = 0.24,
sulfate mean =3.3.
         Goldberg etal. (2001).
         Data same as above.
         Ozkaynaketal. (1996).
         Toronto, 1970-1991.
         TSP (80); COH (0.42
         /1000ft).
                             Study aimed to shed light on population subgroups that my
                             be susceptible to PM effects. Linked data on daily deaths
                             with other health data (physician visits, pharmaceutical R,,,
                             etc.) to identify individuals with presenting health
                             conditions.  PM10 and PM2 5 measured by dichotomous
                             sampler 1 in 6 days until 1992 (2 stations), then daily
                             through 1993. PM missing days interpolated from COH,
                             ext. coefficient, sulfates. Used quasi likelihood estimation
                             in GAM's to assess PM associations with total and cause-
                             specific mortality; and, also, in subgroups by age and/or
                             preexisting health conditions.  Adjusted for CO, NO2, NO,
                             O3 and SO2  in 2-pollutant and all-pollutant models.
The investigators used the universal Quebec medicare
system to obtain disease conditions prior to deaths,
and the roles of these respiratory and cardiovascular
conditions in the PM-mortality associations were
examined.  GAM Poisson model adjusting for
temporal pattern and weather was used.
                             Cause-specific mortality (non-accidental, neoplasm,
                             lung cancer, cardiovascular, coronary artery disease,
                             diabetes, renal disease, and respiratory) series were
                             examined for their associations with O3, using GAM
                             Poisson model adjusting for temporal pattern and
                             weather. Results were also reported for models with
                             adjustments for other pollutants (SO2, CO, NO2, CoH,
                             etc.).
                             Total, cardiovascular, COPD, pneumonia, respiratory,
                             cancer, and the remaining mortality series were
                             related to TSP, SO2, COH, NO2, O3, and CO,
                             adjusting for seasonal cycles (by high-pass filtering
                             each series) temperature, humidity, day-of-week,
                             using OLS regression.  Factor analysis of multiple
                             pollutants was also conducted to extract automobile
                             related pollution, and mortality series were regressed
                             on the resulting automobile factor scores.
                                                     Significant associations found for all-cause (total non-
                                                     accidental) and cause-specific (cancer, CAD, respiratory disease,
                                                     diabetes) with PM measures. Results reported for PM25, COH
                                                     and sulfates. All three PM measures associated with increases in
                                                     total, resp., and "other nonaccidental", and diabetes-related
                                                     mortality. No PM associations found with digestive, accidental,
                                                     renal or neurologic causes of death. Also, mainly in 65+ yr
                                                     group, found consistent associations with increased total
                                                     mortality among persons who had cancer, acute lower resp.
                                                     diseases, any cardiovascular disease, chronic CAD and
                                                     congestive heart failure (CHF).
The PM-mortality associations were found for those who
had acute lower respiratory diseases, chronic coronary
diseases, and congestive heart failure.  They did not find
PM-mortality associations for those chronic upper
respiratory diseases, airways disease, cerebrovascular
diseases, acute coronary artery diseases, and
hypertension. Adjusting for gaseous pollutants generally
attenuated PM RR estimates, but the general pattern
remained. Effects were larger in summer.

The effect of O3 was generally higher in the warm season
and among persons aged 65 years and over. O3 showed
positive and statistically significant associations with
non-accidental cause, neoplasms, cardiovascular disease,
and coronary artery disease.   These associations were not
reduced when the model adjusted for SO2, CO, NO2,
CoH. simultaneously (or when CoH was replaced with
PM25 or total sulfates).
TSP (0 day lag) was significantly associated with total
and cardiovascular deaths. NO2 (0-day lag) was a
significant predictor for respiratory and COPD deaths. 2-
day lagged O3 was associated with total, respiratory, and
pneumonia deaths.  Factor analysis showed factor with
high loadings for NO2, COH, and CO  (apparently
representing automobile factor) as significant predictor
for total, cancer, cardiovascular, respiratory, and
pneumonia deaths.
                                                                                                                                                    Percent excess mortality per 25 //g/m3
                                                                                                                                                    estimated PM25:
                                                                                                                                                    Total deaths (3 d ave.) = 4.4% (2.5, 6.3)
                                                                                                                                                    CV deaths (3 d ave.) = 2.6% (-0.1, 5.5)
                                                                                                                                                    Resp deaths (3 d ave.) = 16.0% (9.7,
                                                                                                                                                    22.8)
                                                                                                                                                    Coronary artery (3 d ave.) = 3.4%
                                                                                                                                                    (-0.2,7.1)
                                                                                                                                                    Diabetes (3 d ave.) =  15.7% (4.8, 27.9)
                                                                                                                                                    Lower Resp Disease (3 d ave.) = 9.7%
                                                                                                                                                    (4.5, 15.1)
                                                                                                                                                    Airways disease (3 d ave.) = 2.7%
                                                                                                                                                    (-0.9, 6.4)
                                                                                                                                                    CHF (3 d ave.) = 8.2% (3.3, 13.4)

                                                                                                                                                    The percent excess deaths estimates
                                                                                                                                                    for non-accidental deaths per IQR
                                                                                                                                                    (average of 0-2 day lags) for CoH,
                                                                                                                                                    predicted PM25, and sulfate were:
                                                                                                                                                    1.98% (1.07, 2.90), 2.17% (1.26,
                                                                                                                                                    3.08), and 1.29% (0.68,  1.90),
                                                                                                                                                    respectively.
                                                                                                             PM RRs not reported.
                                                                                                             Total mortality excess risk: 2.8%
                                                                                                             per 100 Mg/m3 TSP at 0 day lag.

-------
                  TABLE 8A-1 (cont'd).   SHORT-TERM PARTICIPATE  MATTER EXPOSURE  MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
IQRin,ug/m3.
Study Description:  Outcomes, Mean outcome rate, and
ages. Concentration measures or estimates. Modeling
      methods:  lags, smoothing, and covariates.
             Results and Comments.
Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                                    PM Index, lag, Excess Risk%
                                                                                                                                                                  (95% LCL, UCL), Co-pollutants.
OO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          Europe (cont'd)

          Katsouyanni et al. (1997).
          12 European (APHEA) cities.
          1975-1992 (study years
          different from city to city).
          Median Black Smoke (BS)
          levels ranged from 13 in
          London to 73 in Athens
          and Kracow.
          Samolietal. (2001).
          APHEA 1 cities (see
          Katsouyanni (1997)). At least
          five years between 1980-1992.
          The PM levels are the same as
          those in Katsouyanni et al.
          (1997).
          Katsouyanni et al. (2001).
          1990-1997 (variable from city
          to city). Median PM10 ranged
          from 14 (Stockholm) to 66
          (Prague).  Median BS ranged
          from 10 (Dublin) to 64
          (Athens).
          Touloumi et al. (1997).
          6 European (APHEA) cities.
          1977-1992 (study years
          different from city to city).
          Median Black Smoke (BS)
          levels ranged from 14.6 in
          London to 84.4 in Athens.
                               Total daily deaths regressed on BS or SO2 using Poisson
                               models, adjusting for seasonal cycles, day-of-week,
                               influenza epidemic, holidays, temp., humidity. Final
                               analysis done with autoregressive Poisson models to allow
                               for overdispersion and autocorrelation.  Pollution effects
                               examined at 0 through 3 day lags and multi-day averages
                               thereof. When city-specific coefficients tested to be
                               homogeneous, overall estimates obtained by computing
                               variance-weighted means of city-specific estimates (fixed
                               effects model).  When significant heterogeneity present,
                               source of heterogeneity sought by examining a predefined
                               list of city-specific variables,  including annual and seasonal
                               means of pollution and weather variables, number of
                               monitoring sites, correlation between measurements from
                               different sites, age-standardized mortality, proportion of
                               elderly people, smoking prevalence, and geographic
                               difference (north-south, east-west). A random effects model
                               was fit when heterogeneity could not be explained.

                               In order to further investigate the source of the regional
                               heterogeneity of PM effects, and to examine the sensitivity
                               of the RRs, the APHEA data were reanalyzed by the
                               APHEA investigators themselves (Samoli et al., 2001).
                               Unlike previous model in which  sinusoidal terms for
                               seasonal control and polynomial  terms for weather, the
                               investigators this time used a  GAM model with smoothing
                               terms for seasonal trend and weather, which is more
                               commonly used approach in recent years.

                               The 2nd phase of APHEA (APHEA 2) put emphasis on the
                               effect modification by city-specific factors. The first stage
                               of city specific regressions  used GAM Poisson model. The
                               second stage regression analysis  was conducted to explain
                               any heterogeneity of air pollution effects using city-specific
                               variables.  These city-specific variables included average air
                               pollution levels, average temperature/humidity, age-
                               standardize mortality rate, region indicators, etc.


                               Results of the short-term effects of ambient NO2 and/or O3
                               on daily deaths from all causes (excluding accidents) were
                               discussed to provide a basis for comparison with estimated
                               SO2 or BS effects in APHEA  cities.  Poisson models,
                               lag/averaging of pollution,  and the computation of
                               combined effects across the cities were done in the same
                               way as done by Katsouyanni et al. (1997), as above.
                                                      Substantial variation in pollution levels (winter mean SO2 ranged
                                                      from 30 to 330 ,ug/m3), climate, and seasonal patterns were
                                                      observed across cities. Significant heterogeneity was found for
                                                      the effects of BS and SO2, but only the separation between
                                                      western and central eastern European cities resulted in more
                                                      homogeneous subgroups. Significant heterogeneity for SO2
                                                      remained in western cities. Cumulative effects of prolonged
                                                      (two to four days) exposure to air pollutants resulted in estimates
                                                      comparable with the one day effects. The effects of both SO2
                                                      and BS were stronger during the summer and were independent.
                                                      T he estimated relative risks for central-eastern cities were larger
                                                      than those obtained from the previous model.  Also, restricting
                                                      the analysis to days with concentration < 150ug/m3 further
                                                      reduced the differences between the western and central-eastern
                                                      European cities.  The authors concluded that part of the
                                                      heterogeneity in the estimated air pollution effects between
                                                      western and central eastern cities in previous publications was
                                                      caused by the statistical approach and the data range.


                                                      The authors found several effect modifiers. The cities with
                                                      higher NO2 levels showed larger PM effects. The cities with
                                                      warmer climate showed larger PM effects.  The cities with low
                                                      standardized mortality rate showed larger PM  effects.  The
                                                      combined estimate of mortality RRs per 10ug/m3 PM10 or BS
                                                      was: 0.6% (0.4, 0.8). The PM RR estimates for cities with low
                                                      vs. high NO2 levels were 0.19% (0, 0.41) and  0.80% (0.67,
                                                      0.93);  0.29% (0.16, 0.42) for cities with cold  climate and 0.82%
                                                      (0.69, 0.96) for warm climate, respectively.

                                                      Significant positive associations found between daily deaths and
                                                      both NO2 and O3. Tendency for larger effects  of NO2 in cities
                                                      with higher levels of BS. When BS included in the model,
                                                      pooled estimate for O3 effect only slightly reduced, but
                                                      coefficient for NO2 reduced by half. Authors speculated that
                                                      short-term effects of NO2 on mortality confounded by other
                                                      vehicle-derived pollutants.
                                                       Total mortality excess deaths per
                                                       25 |/g/m3 increase in single day BS for
                                                       western European cities: 1.4(1.0, 1.8);
                                                       and 2 (1, 3) per 50 ,ug/m3 PM10
                                                       increase. In central/eastern Europe
                                                       cities, corresponding figure was 0.3
                                                       (0.05, 0.5) per 25 ^g/m3 BS.
                                                       Total mortality RRs per 50ug/m3 BS for
                                                       all cities, western cities, and central-
                                                       eastern cities using the GAM approach
                                                       were: 2.2% (1.8, 2.6); 3.1% (2.4, 3.9);
                                                       and, 2.2% (1.4, 2.3), respectively. In
                                                       contrast, those with old method were:
                                                       1.3% (0.9, 1.7); 2.9% (2.1, 3.7); and,
                                                       0.6% (0.1, 1.1), respectively.
                                                       NO2 and/or O3 estimates only.

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS  STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description: Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates. Modeling
        methods: lags, smoothing, and covariates.
             Results and Comments.
Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                                    PM Index, lag, Excess Risk%
                                                                                                                                                                  (95% LCL, UCL), Co-pollutants.
OO
OO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          Europe (cont'd)

          Zmirouetal. (1998).
          10 European (APHEA) cities.
          1977-1992 (study years
          different from city to city).
          Median Black Smoke (BS)
          levels ranged from 13 in
          London to 73 in Kracow.
          Bremneretal. (1999).
          London, UK, 1992-1994.
          BS (13), PM10 (29).
Prescottetal. (1998).
Edinburgh, UK, 1981-1995.
PM10 (21, by TEOM only for
1992-1995); BS (8.7).
          Rooneyetal. (1998).
          England and Wales, and
          Greater London, UK
          PM10 (56, during the worst heat
          wave; 39, July-August mean)
          Wordleyetal. (1997).
          Birmingham, UK,
          1992-1994.
          PM10 (apparently
          beta-attenuation, 26)
                               Cardiovascular, respiratory, and digestive mortality series
                               in 10 European cities analyzed to examine cause-specificity
                               of air pollution. The mortality series were analyzed for
                               associations with PM (BS, except TSP in Milan and
                               Bratislava; PM13 in Lyon), NO2, O3, and SO2. Poisson
                               models, lag/averaging of pollution, and computation of
                               combined effects across the cities done in the same way as
                               by Katsouyanni et al.  (1997), above.

                               Total, cardiovascular, and respiratory (by age) mortality
                               series were regressed on PM10, BS, O3, NO2, CO, and SO2,
                               adjusting for seasonal cycles, day-of-week, influenza,
                               holidays, temperature, humidity, and autocorrelation using
                               Poisson model.
Both mortality (total, cardiovascular, and respiratory) and
emergency hospital admissions (cardiovascular and
respiratory), in two age groups (<65 and >= 65), were
analyzed for their associations with PM10, BS, SO2, NO2, O3,
and CO, using Poisson regression adjusting for seasonal
cycles, day-of-week, temperature, and wind speed.

Excess deaths, by age, sex, and cause, during the 1995 heat
wave were estimated by taking the difference between  the
deaths during heat wave and the 31-day moving averages
(for 1995 and 1993-94 separately).  The pollution effects,
additively for O3, PM10, and NO2, were estimated based on
the published season-specific coefficients from the  1987-
1992 study (Anderson et al., 1996).

Mortality data were analyzed for COPD, pneumonia, all
respiratory diseases, all circulatory diseases, and all causes.
Mortality associations with PM10, NO2, SO2, and O3 were
examined using OLS (with some health outcomes log- or
square-root transformed), adjusting for day-of-week, month,
linear trend, temperature and relative humidity. The study
also analyzed hospital admission data.
                                                         The cardiovascular and respiratory mortality series were
                                                         associated with BS and SO2 in western European cities, but not
                                                         in the five central European cities. NO2 did not show consistent
                                                         mortality associations.  RRs for respiratory causes were at least
                                                         equal to, or greater than those for cardiovascular causes. No
                                                         pollutant exhibited any association with digestive mortality.
                                                       Pooled cardiovascular mortality percent
                                                       excess deaths per 25 ,ug/m3 increase in
                                                       BS for western European cities: 1.0
                                                       (0.3, 1.7); for respiratory mortality, it
                                                       was 2.0 (0.8, 3.2) in single lag day
                                                       models (the lags apparently varied
                                                       across cities).
All effect size estimates (except O3) were positive for total deaths  1.9% (0.0, 3.8) per 25 ^g/m3 BS at lag
(though not significant for single lag models).  The effects of O3
found in 1987-1992 were not replicated, except in
cardiovascular deaths.  Multiple day averaging (e.g., 0-1, 0-2
days) tend to give more significant effect size estimates. The
effect size for PM10 and BS were similar for the same
distributional increment.

Among all the pollutants, BS was most significantly associated
with all cause, cardiovascular, and respiratory mortality series.
In the subset in which PM10 data were available, the RR
estimates for BS and PM10 for all cause elderly mortality were
comparable. Other pollutants' mortality associations were
generally inconsistent.

Air pollution levels at all the locations rose during the heat wave.
8.9% and 16.1% excess deaths were estimated for England and
Wales, and Greater London, respectively.  Of these excess
deaths, up to 62% and 38%, respectively for these locations, may
be attributable to combined pollution effects.
                                                                                                                     1 day; 1.3% (-1.0, 3.6) per 50 ^g/
                                                                                                                     PM10 at lag 1 d for total deaths.  Resp.
                                                                                                                     deaths (3 d) = 4.9% (0.5, 9.4). CVD
                                                                                                                     deaths (1 d) = 3.0%(0.3, 5.7).
                                                       3.8 (1.3, 6.4) per 25 i/g/m3 increase in
                                                       BS for all cause mortality in age 65+
                                                       group, avg. of 1-3 day lags.
                                                                                                                                                    2.6% increase for PM10 in Greater
                                                                                                                                                    London during heat wave.
                                                                                        Total, circulatory, and COPD deaths were significantly
                                                                                        associated with 1-day lag PM10. The gaseous pollutants "did not
                                                                                        have significant associations independent from that of PM10",
                                                                                        and the results for gaseous pollutants were not presented. The
                                                                                        impact of reducing PM10 to below 70 ,ug/m3 was estimated to be
                                                                                        "small" (0.2% for total deaths), but the PM10 level above 70
                                                                                        ,ug/m3 occurred only once during the study period.
                                                       5.6% (0.5, 11.0) per 50 ^g/m3 PM10 at 1
                                                       d lag for total deaths. COPD (1 d lag)
                                                       deaths = 27.6 (5.1,54.9).
                                                       Circulatory (1 d) deaths = 8.8 (1.9,
                                                       17.1)

-------
                  TABLE 8A-1  (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
Study Description:  Outcomes, Mean outcome rate, and
ages. Concentration measures or estimates.  Modeling
      methods:  lags, smoothing, and covariates.
             Results and Comments.
Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                                  PM Index, lag, Excess Risk%
                                                                                                                                                                 (95% LCL, UCL), Co-pollutants.
OO
vo
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Europe (cont'd)

          Hoek et al. (2000).
          The Netherlands, 1986-1994.
          PM10 (median 34);
          BS (median 10).
          Hoek etal. (2001).
          The Netherlands. 1986-1994.
          PM10 (median 34);
          BS (median 10).
          Ponkaetal. (1998).
          Helsinki, Finland, 1987-1993.
          TSP (median 64);
          PM10 (median 28)
          Peters etal. (1999a).
          A highly polluted coal basin
          area in the Czech Republic and
          a rural  area in Germany,
          northeast Bavaria districts.
          1982-1994. TSP: mean = 121.1
          and 51.6, respectively, for these
          two regions. PM10 and PM2 5
          were also measured in the coal
          basin during 1993-1994 (mean
          = 65.9  and 51.0, respectively).
                               Total, cardiovascular, COPD, and pneumonia mortality
                               series were regressed on PM10, BS, sulfate, nitrate, O3, SO2,
                               CO, adjusting for seasonal cycles, day-of-week, influenza,
                               temperature, and humidity using Poisson GAM model.
                               Deaths occurring inside and outside hospitals were also
                               examined.

                               This study of the whole population of the Netherlands,  with
                               its large sample size (mean daily total deaths ~ 330, allowed
                               examination of specific cardiovascular cause of deaths.
                               GAM Poisson regression models, adjusting for seasonal
                               cycles, temperature, humidity, day-of-week was used.
                               Total and cardiovascular deaths, for age groups < 65 and 65
                               +, were related to PM10, TSP, SO2, NO2, and O3, using
                               Poisson model adjusting for temperature, relative humidity,
                               day-of-week, temporal patterns, holiday and influenza
                               epidemics.
                               Non-accidental total and cardiovascular deaths (mean =
                               18.2 and 12.0 per day, for the Czech and Bavaria areas,
                               respectively). The APHEA approach (Poisson model with
                               sine/cosine, temperature as a quadratic function, relative
                               humidity, influenza, day-of-week as covariates), as well as
                               GAM Poisson models were considered. Logarithm of TSP,
                               SO2, NO2, O3, and CO (and PM10 and PM25 for 1993-1994)
                               were examined at lags 0 through 3 days.
                                                      Particulate air pollution was not more consistently associated
                                                      with mortality than were the gaseous pollutants SO2 and NO2.
                                                      Sulfate, nitrate, and BS were more consistently associated with
                                                      total mortality than was PM10. The RRs for all pollutants were
                                                      larger in the summer months than in the winter months.
                                                      Deaths due to heart failure, arrhythmia, cerebrovascular causes,
                                                      and thrombocytic causes were more strongly (~ 2.5 to 4 times
                                                      larger relative risks) associated with air pollution than the overall
                                                      cardiovascular deaths (CVD) or myocardial infarction (MI) and
                                                      other ischemic heart disease (IHD).
                                                      No pollutant significantly associated with mortality from all
                                                      cardiovascular or CVD causes in 65+ year age group. Only in
                                                      age <65 year group, PM10 associated with total and CVD deaths
                                                      with 4 and 5 d lags, respectively. The "significant" lags were
                                                      rather "spiky". O3 was also associated with CVD mortality <65
                                                      yr. group with inconsistent signs and late and spiky lags (neg. on
                                                      d 5 and pos. on d 6).

                                                      In the coal basin (i.e., the Czech Republic polluted area), on the
                                                      average, 68% of the TSP was PM10, and most of PM10 was PM25
                                                      (75%).  For the coal basin, associations were found between the
                                                      logarithm of TSP and all-cause mortality at lag 1 or 2 days. SO2
                                                      was also associated with all-cause mortality with slightly lower
                                                      significance. PM10 and PM25 were both associated with all-
                                                      cause mortality in 1993-1994 with a lag of 1-day.  NO2, O3 and
                                                      CO were positively but more weakly associated with mortality
                                                      than PM indices or SO2. In the Bavarian region, neither TSP nor
                                                      SO2 was associated with mortality, but CO (at lag 1-day) and O3
                                                      (at lag 0-day) were associated with all-cause mortality.
                                                       0.9 (0.1, 1.7) per 50 ,ug/m3 PM10; 1.0
                                                       (0.5, 1.5) per 25 ^g/m3 BS; 3.2 (0.6,
                                                       5.9) per 25 ,ug/m3 sulfate; and 4.1 (1.4,
                                                       6.9) per 25 ^g/m3 nitrate, all at 1 day
                                                       lag.
                                                       For PM10 (7-day mean), RRs for total
                                                       CVD, MI/IHD, arrhythmia, heart
                                                       failure, cerebrovascular, and
                                                       thrombocytic mortality per 80 ug/m3
                                                       increase were: 1.2% (-1.6, 4.1), 0.5% (-
                                                       3.6, 4.8), 4.1% (-6.8, 16.3), 3.6% (-4.0,
                                                       11.8), 3.1%(-2.9, 9.4), and 1.0%(-
                                                       10.6, 14.3), respectively. The RRs for
                                                       BS were larger and more significant
                                                       than those for PM10.
                                                       18.8% (5.6, 33.2) per 50 ^g/m3 PM10 4
                                                       day lag (other lags negative or zero).
                                                       Total mortality excess deaths per 100
                                                       //g/m3 increase in TSP for the Czech
                                                       region: 3.8 (0.8, 6.9) at lag 2-day for
                                                       1982-1994 period. For period 1993-
                                                       1994, 9.5 (1.2, 18.5) per 100 ,ug/m3
                                                       increase in TSP at lag 1-day, and 4.8
                                                       (0.7, 9.0) per 50 ,ug/m3 increase in
                                                       PM10; and 1.4 (-0.5, 3.4) per 25  ,ug/m3
                                                       PM,,.
O

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description:  Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates. Modeling
        methods:  lags, smoothing, and covariates.
                  Results and Comments.
     Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                                 PM Index, lag, Excess Risk%
                                                                                                                                                                (95% LCL, UCL), Co-pollutants.
OO
to
o
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Europe (cont'd)

          Hoeketal. (1997).
          Rotterdam, the Netherlands,
          1983-1991. TSP (median 42);
          BS (median 13).
          Kotesovec et al. (2000).
          Northern Bohemia, Czech
          Republic, 1982-1994.
          TSP (121.3).
Zanobetti et al. (2000a).
Milan, Italy.  1980-1989.
TSP mean =  142.
          Anderson et al. (1996).
          London, UK, 1987-1992.
          BS(15)
                               Total mortality (also by age group) was regressed on TSP,
                               Fe (from TSP filter), BS, O3, SO2, CO, adjusting for
                               seasonal cycles, day-of-week, influenza, temperature, and
                               humidity using Poisson GAM model.
Total (excluding accidents and children younger than 1 yr),
cause specific (cardiovascular and cancer), age (65 and less
vs. otherwise), and gender specific mortality series were
examined for their associations with TSP and SO2 using
logistic model, adjusting for seasonal cycles, influenza
epidemics, linear and quadratic temperature terms. Lags 0
through 6 days, as well as a 7 day mean values were
examined.

The focus of this study was to quantify mortality
displacement using GAM distributed lag models. Non-
accidental total deaths were regressed on smooth function of
TSP  distributed over the same day and the previous 45 days
using penalized splines for the smooth terms and seasonal
cycles, temperature, humidity, day-of-week, holidays, and
influenza epidemics. The mortality displacement was
modeled as the initial positive increase, negative rebound
(due  to depletion), followed by another positive coefficients
period, and the sum of the three phases were considered as
the total cumulative effect.

Total, cardiovascular, and respiratory mortality series were
regressed on BS, O3, NO2, and SO2, adjusting for seasonal
cycles, day-of-week, influenza, holidays, temperature,
humidity, and autocorrelation using Poisson model.
Daily deaths were most consistently associated with TSP. TSP
and O3 effects were "independent" of SO2 and CO. Total iron
(from TSP filter) was associated "less consistently" with
mortality than TSP was. The estimated RRs for PM indices were
higher in warm season than in cold season.

For the total mortality, TSP, but not SO2, was associated. There
were apparent differences in associations were found between
men and women. For example, for age below 65 cardiovascular
mortality was associated with TSP for men but not for women.
TSP was positively associated with mortality up to 13 days,
followed by nearly zero coefficients between 14 and 20 days,
and then followed by smaller but positive coefficients up to the
45th day (maximum examined).  The sum of these coefficients
was over three times larger than that for the single-day estimate.
                                                                                       Both O3 (0 day lag) and BS (1 day lag) were significant
                                                                                       predictors of total deaths. O3 was also positively significantly
                                                                                       associated with respiratory and cardiovascular deaths.  The effect
                                                                                       size estimates per the same distributional increment (10% to
                                                                                       90%) were larger for O3 than for BS.  These effects were larger
                                                                                       in warm season.  SO2 and NO2 were not consistently associated
                                                                                       with mortality.
                                                                                                                   5.5 (1.1, 9.9) per 100 //g/m3 TSP at 1
                                                                                                                   day lag.
                                                                                                                                                  Total mortality percent excess deaths
                                                                                                                                                  per 100 |/g/m3 increase in TSP at 2 day
                                                                                                                                                  lag was 3.4 (0.5, 6.4).
                                                                                                                                                            Total mortality percent increase
                                                                                                                                                            estimates per IQR increase in TSP: 2.2
                                                                                                                                                            (1.4, 3.1) for single-day model; 6.7 (3.8,
                                                                                                                                                            9.6) for distributed lag model.
                                                           2.8% (1.4, 4.3) per 25 ^g/m3
                                                           lag for total deaths.
                                                           CVD(1 d)= 1.0 (-1.1, 3.1).
                                                           Resp. (1 d)= 1.1 (-2.7, 5.0).
                                                                                                                                                                                      BS at 1-d
O

-------
                  TABLE 8A-1  (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description: Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates. Modeling
        methods: lags, smoothing, and covariates.
                  Results and Comments.
     Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                                   PM Index, lag, Excess Risk%
                                                                                                                                                                 (95% LCL, UCL), Co-pollutants.
OO
>
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Europe (cont'd)

          Michelozzietal. (1998).
          Rome, Italy, 1992-1995.
          TSP ("PM13" beta attenuation,
          84).
          Garcia-Aymerich et al. (2000).
          Barcelona, Spain.  1985-1989.
          Black Smoke no data
          distribution was reported).
Rahlenbeck and Kahl (1996).
East Berlin,
1981-1989.
"SP" (beta attenuation, 97)
          Rossi etal. (1999).
          Milan, Italy, 1980-1989
          TSP ("PM13" beta attenuation,
          142)
          Sunyer et al. (2000).
          Barcelona, Spain.
          1990-1995.
          BS means: 43.9 for case period,
          and 43.1 for control period.
Total mortality was related to PM13, SO2, NO2, CO, and O3,
using Poisson GAM model, adjusting for seasonal cycles,
temperature, humidity, day-of-week, and holiday.  Analysis
of mortality by place of residence, by season, age,  place of
death (in or out of hospital), and cause was also conducted.

Daily total (mean = 1.8/day), respiratory, and cardiovascular
mortality counts of a cohort (9,987 people) with COPD or
asthma were associated with black smoke (24-hr), SO2 (24-
hr and 1-hr max), NO2 (24-hr and 1-hr max), O3 (1-hr max),
temperature, and relative humidity. Poisson regression
models using APHEA protocol were used. The resulting
RRs were compared with those of the general population.

Total mortality (as well as deviations from long-wave
cycles) was regressed on SP and SO2, adjusting for day-of-
week, month, year, temperature, and relative humidity,
using OLS, with options to log-transform pollution, and w/
and w/o days with pollution  above 150 //g/m3.

Specific causes of death (respiratory, respiratory infections,
COPD, circulatory, cardiac, heart failure, and myocardial
infarction) were related to TSP, SO2, and NO2, adjusting for
seasonal cycles, temperature, and humidity, using  Poisson
GAM model.
                               Those over age 35 who sought emergency room services for
                               COPD exacerbation during 1985-1989 and died during
                               1990-1995 were included in analysis. Total, respiratory,
                               and cardiovascular deaths were analyzed using a conditional
                               logistic regression analysis with a case-crossover design,
                               adjusting for temperature, relative humidity, and influenza
                               epidemics. Bi-directional control period at 7 days was used.
                               Average of the same and previous 2 days used for pollution
                               exposure period.  Data also stratified by potential effect
                               modifiers (e.g., age, gender, severity and number of ER
                               visits, etc.).
                                                                                        PM13 and NO2 were most consistently associated with mortality.
                                                                                        CO and O3 coefficients were positive, SO2 coefficients negative.
                                                                                        RR estimates higher in the warmer season. RRs similar for in-
                                                                                        and out-of hospital deaths.
                                                                                        Daily mortality in COPD patients was associated with all six
                                                                                        pollution indices. This association was stronger than in the
                                                                                        general population only for daily 1-hr max of SO2, daily 1-hr
                                                                                        max and daily means of NO2.  BS and daily means of SO2
                                                                                        showed similar or weaker associations for COPD patients than
                                                                                        for the general population.
Both SP and SO2 were significantly associated with total
mortality with 2 day lag in single pollutant model. When both
pollutants were included, their coefficients were reduced by 33%
and 46% for SP and SO2, respectively.
All three pollutants were associated with all cause mortality.
Cause-specific analysis was conducted for TSP only.
Respiratory infection and heart failure deaths were both
associated with TSP  on the concurrent day, whereas the
associations for myocardial infarction and COPD deaths were
found for the average of 3 to 4 day prior TSP.

BS levels were associated with all cause deaths. The association
was stronger for respiratory causes. Older women, patients
admitted to intensive care units, and patients with a higher rate
of ER visits were at greater risk of deaths associated with BS.
                                                            1.9% (0.5, 3.4) per 50 //g/m3 PM13 atO
                                                            day lag.
                                                            Total mortality percent increase per
                                                            25 //g/m3 increase in avg. of 0-3 day
                                                            lags of BS:  2.76 (1.31, 4.23) in general
                                                            population, and 1.14 (-4.4, 6.98) in the
                                                            COPD cohort.
                                                                                                                                                              6.1% per 100 //g/m3 "SP" at 2 day lag.
                                                                                                                                                   3.3% (2.4, 4.3) per 100 ,ug/m3 TSP at 0
                                                                                                                                                   day lag.
                                                                                                                    Percent increase per 25 //g/m3 increase
                                                                                                                    in 3-day average BS: 14.2(1.6,28.4)
                                                                                                                    for all causes; 9.7 (-10.2, 34.1) for
                                                                                                                    cardiovascular deaths; 23.2 (3.0, 47.4)
                                                                                                                    for respiratory deaths.
O

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
           Reference, Location, Years,
           PM Index, Mean or Median,
  Study Description:  Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates.  Modeling
        methods:  lags, smoothing, and covariates.
             Results and Comments.
Design Issues, Uncertainties, Quantitative Outcomes.
  PM Index, lag, Excess Risk%
(95% LCL, UCL), Co-pollutants.
          Europe (cont'd)

          Sunyer and Basagana (2001).
          Barcelona, Spain. 1990-1995.
          See Sunyer et al. (2000) for
          PM levels.
                                         The analysis assessed any "independent" particle effects,
                                         after controlling for gaseous pollutants, on a cohort of
                                         patients with COPD (see the summary description for
                                         Sunyer et al. (2000) for analytical approach).  PM10, NO2,
                                         O,, and CO were analyzed.
                                                        PM10, but not gaseous pollutants were associated with mortality    Odds ratio for all cause mortality per
                                                        for all causes. In the two-pollutant models, the PM10-mortality
                                                        associations were not diminished, whereas those with gaseous
                                                        pollutants were.
                                                      IQR PM10 on the same-day (27 ug/m3)
                                                      was 11% (0, 24). In two pollutant
                                                      models, the PM10 RRs were 10.5%,
                                                      12.9%, and 10.8% with NO2, O3, and
                                                      CO, respectively.
OO
to
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Tobias and Campbell (1999).
          Barcelona, Spain.
          1991-1995.
          Black Smoke (BS)
          (no data distribution
          was reported).
           Alberdi Odriozola et al. (1998).
           Madrid, Spain, 1986-1992.
           "TSP" (beta attenuation,
           47 for average of 2 stations)

           Diaz etal. (1999).
           Madrid, Spain. 1990-1992.
           TSP (no data distribution
           was reported).
Study examined the sensitivity of estimated total mortality
effects of BS to different approaches to modeling influenza
epidemics:  (1) with a single dummy variable; (2) with three
dummy variables; (3) using daily number of cases of
influenza. Poisson regression used to model total daily
mortality, adjusting for weather, long-term trend, and
season, apparently following APHEA protocol.

Total, respiratory, and cardiovascular deaths were related to
TSP and SO2. Multivariate autoregressive integrated
moving average models used to adjust for season,
temperature, relative humidity, and influenza epidemics.

Non-accidental, respiratory, and cardiovascular deaths
(mean = 62.4, 6.3, and 23.8 per day, respectively).  Auto-
regressive Integrated Moving Average (ARIMA) models fit
to both depend, and independ. variables first to remove
auto-correlation and seasonality (i.e., pre-whitening"),
followed by examining cross-correlation to find optimal
lags. Multivariate OLS models thus included ARIMA
components, seasonal cycles (sine/cosine), V-shaped temp.,
and optimal lags found for pollution and weather variables.
TSP, SO2, NO2, and O3 examined. Season-specific analyses
also conducted.
                                                                                                 Using the reported daily number of influenza cases resulted in a
                                                                                                 better fit (i.e., a lower AIC) than those using dummy variables.
                                                                                                 In the "better" model, the black smoke coefficient was about
                                                                                                 10% smaller than those in the models with dummy influenza
                                                                                                 variables, but remained significant. Lags not reported.
                                                                                                 TSP (1-day lag) and SO2 (3-day lagged) were independently
                                                                                                 associated with mortality.
                                                                                                 TSP was significantly associated with non-accidental mortality
                                                                                                 at lag 0 for year around and winter, but with a 1-day lag in
                                                                                                 summer. A similar pattern was seen for circulatory deaths.  For
                                                                                                 respiratory mortality, a significant association with TSP was
                                                                                                 found only in summer (0-day lag).  SO2, NOx, and NO2 showed
                                                                                                 similar associations with non-accidental deaths at lag 0 day. O3'
                                                                                                 associations with non-accidental mortality was U-shaped, with
                                                                                                 inconsistent lags (1, 4, and 10).
                                                      Total mortality excess deaths per 25
                                                      Mg/m3 increase in BS:  1.37 (0.20,  2.56)
                                                      for model using the daily case of
                                                      influenza; 1.71 (0.53, 2.91) for model
                                                      with three influenza dummy variables.
                                                      4.8% (1.8, 7.7) per 100 ,ug/m3 TSP at
                                                      lag 1 day.
                                                      For non-accidental mortality, excess
                                                      deaths was 7.4% (confidence bands not
                                                      reported; p < 0.05) per 100 ,ug/m3 TSP
                                                      at 0 day lag.
O

-------
                  TABLE 8A-1  (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description: Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates. Modeling
        methods: lags, smoothing, and covariates.
                                                                                                                     Results and Comments.
                                                                                                       Design Issues, Uncertainties, Quantitative Outcomes.
     PM Index, lag, Excess Risk%
   (95% LCL, UCL), Co-pollutants.
OO
>
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
Europe (cont'd)

Wichmann et al., (2000).
Erfurt, Germany.
1995-1998.
Number counts (NC) & mass
concentrations (MC) of
ultrafme particles in three size
classes, 0.01 to 0.1 fan, and
fine particles in three size
classes from 0.1 to 2.5  fan.
diameter, using Spectrometryll
Mobile Aerosol Spectrometry
(MAS).
MAS MC PM2.5-0.01  (mean
25.8, median 18.8, IQR 19.9).
Filter measurements of PM10
(mean 38.2, median 31.0, IQR
27.7) and PM25 (mean  26.3,
median 20.2, IQR 18.5).  MAS
NC2.5-0.01 (mean 17,966 per
cu.cm, median 14,769, IQR
13,269).
          Zeghnoun et al. (2001).
          Rouen and Le Havre, France.
          1990-1995.  PM13 mean = 32.9
          for Rouen, 36.4 for Le Havre.
          BS mean = 18.7 for Rouen,
          16.3 for Le Havre.
          Roemeretal. (2001).
          Amsterdam. 1987-1998.
          BS and PM10 means in
          "background" = 10 and 39;
          BS mean in "traffic" area = 21.
          (No PM10 measurements
          available at traffic sites)
Total non-accidental, cardiovascular, and respiratory deaths
(mean 4.88, 2.87, 1.08 per day, respectively) were related to
particle mass concentration and number counts in each size
class, and to mass concentrations of gaseous co-pollutants
NO2, CO, SO2, using GAM regression models adjusted for
temporal trends, day of week, weekly national influenza
rates, temperature and relative humidity.  Data analyzed by
season, age group, and cause of death separately.  Single-
day lags and polynomial distributed lag models (PDL) used.
Particle indices and pollutants fitted using linear, log-
transformed, and LOESS transformations. Two-pollutant
models with a particle index and a gaseous pollutant were
fitted. The "best" model  as used by Wichmann et al. (2000)
was that having the highest t-statistic, since other criteria
(e.g., log-likelihood for nested models) and AIC for non-
nested models could not be applied due to different numbers
of observations in each model. There should be little
difference between these  approaches and resulting
differences in results should be small in practice.
Sensitivity analyses included stratifying data by season,
winter year, age, cause of death, or transformation of the
pollution variable (none,  logarithmic, non-parametric
smooth).

Total, cardiovascular, and respiratory mortality series were
regressed on BS, PM13, SO2, NO2, and O3 in 1- and 2-
pollutant models using GAM Poisson models adjusting for
seasonal trends, day-of-week, and weather.
                               Daily deaths for those who lived along roads with more than
                               10,000 motor vehicle, as well as deaths for total population,
                               were analyzed using data from background and traffic
                               monitors. Poisson GAM model was used adjusting for
                               season, day-of-week, and weather. BS, PM10, SO2, NO2,
                               CO, and O3 were analyzed.
                                                                                                  Loss of stat. power by using a small city with a small number of
                                                                                                  deaths was offset by advantage of having good exposure
                                                                                                  representation from single monitoring site. Since ultrafme
                                                                                                  particles can coagulate into larger aggregates in a few hours,
                                                                                                  ultrafme particle size and numbers can increase into the fine
                                                                                                  particle category, resulting in some ambiguity. Significant
                                                                                                  associations were found between mortality and ultrafme particle
                                                                                                  number concentration (NC), ultrafme particle mass
                                                                                                  concentration (MC), fine particle mass concentration, or SO2
                                                                                                  concentration. The correlation between MCO.01-2.5 and
                                                                                                  NCO.01-0.1 is only moderate, suggesting it may be possible to
                                                                                                  partially separate effects of ultrafme and fine particles. The most
                                                                                                  predictive single-day effects are either immediate (lag 0 or 1) or
                                                                                                  delayed (lag 4 or 5  days), but cumulative effects characterized by
                                                                                                  PDL are larger than single-day effects. The significance of SO2
                                                                                                  is robust, but hard to explain as a true causal factor since its
                                                                                                  concentrations are very low.  Age is an important modifying
                                                                                                  factor, with larger effects at ages < 70 than > 70 years.
                                                                                                  Respiratory mortality has a higher RR than cardio- vascular
                                                                                                  mortality. A large number of models were fitted, with some
                                                                                                  significant findings of association between mortality and particle
                                                                                                  mass or number indices.
                                                                                        In Rouen, O3, SO2, and NO2 were each significantly associated
                                                                                        with total, respiratory, and cardiovascular mortality, respectively.
                                                                                        In Le Havre, SO2 and PM13 were associated with cardiovascular
                                                                                        mortality. However, the lack of statistical significance reported
                                                                                        for most of these results may be in part due to the relatively
                                                                                        small population size of these cities (430,000 and 260,000,
                                                                                        respectively).
                                                         Correlations between the background monitors and traffic
                                                         monitors were moderate for BS (r = 0.55) but higher for NO2 (r =
                                                         0.79) and O3 (r = 0.80). BS and NO2 were associated with
                                                         mortality in both total and traffic population. Estimated RR for
                                                         traffic population using background sites was larger than the RR
                                                         for total population using background sites. The RR for total
                                                         pop. using traffic sites was smaller that RRs for total population
                                                         using background sites. This is not surprising since the mean BS
                                                         for traffic sites were larger that for background sites.
Total mortality excess deaths:
Filter PM10 (0-4 d lag) = 6.6 (0.7, 12.8)
per 50 Mg/m3. Filter PM25 (0-1 d) = 3.0
(-1.7,7.9). MCforPM001.256.2%(1.4,
11.2) for all year; by season,
Winter = 9.2% (3.0, 15.7)
Spring = 5.2% (-2.0, 12.8)
Summer = -4.7% (-18.7, 11.7)
Fall = 9.7% (1.9, 18.1)

For ultrafme PM, NC 0.01-0.1 (0-4 d
lag):
All Year =8.2% (0.3, 16.9)
Winter = 9.7% (0.3, 19.9)
Spring = 10.5% (-1.4, 23.9)
Summer = -13.9% (-29.8, 5.7)
Fall = 12.0% (2.1,22.7)
PM13 total mortality RRs per IQR were
0.5% (-1.1, 2.1) in Rouen (IQR=20.6,
1-day lag) and 1.9% (-0.8, 7.4) in Le
Havre (IQR=23.9, 1-day lag ). BS total
mortality RRs per IQR were 0.5% (-1.8,
2.9) in Rouen (IQR=14.2, 1-day lag)
and 0.3% (-1.6, 2.2) in Le Havre
(IQR=11.5, 0-1 day lag avg.).

The RRs per 100 ug/m3 BS (at lag
1-day) were 1.383 (1.153, 1.659),
1.887 (1.207, 2.949), and 1.122 (1.023,
1.231) for total population using
background sites, traffic population
using background sites, and total
population using traffic sites,
respectively.  Results for traffic pop.
using traffic sites not reported)

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS  STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
Study Description:  Outcomes, Mean outcome rate, and
ages. Concentration measures or estimates. Modeling
      methods:  lags, smoothing, and covariates.
             Results and Comments.
Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                                    PM Index, lag, Excess Risk%
                                                                                                                                                                  (95% LCL, UCL), Co-pollutants.
OO
>
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          Europe (cont'd)

          Anderson et al. (2001).
          The west Midlands
          conurbation, UK.  1994-1996.
          PM means: PM10 = 23,
          PM25 = 15, PM10.25 = 9,
          BS = 13.2, sulfate = 3.7.
          Keatinge and Donaldson
          (2001).  Greater London,
          England, 1976-1995.
          BS mean = 17.7.
          Latin America

          Cifuentes et al. (2000).
          Santiago, Chile.
          1988-1996.
          PM25(64.0), andPM1025
          (47.3).
          Castillejosetal. (2000).
          Mexico City.
          1992-1995.
          PM10 (44.6), PM25 (27.4),
          andPM10.25(17.2).
          Loomisetal. (1999).
          Mexico-City, 1993-1995.
          PM25 (mean: 27.4 ,ug/m3)
                               Non-accidental cause, cardiovascular, and respiratory
                               mortality (as well as hospital admissions) were analyzed for
                               their associations with PM indices and gaseous pollutants
                               using GAM Poisson models adjusting for seasonal cycles,
                               day-of-week, and weather.
                               The study examined potential confounding effects of
                               atypical cold weather on air pollution/mortality
                               relationships. First, air pollution variables (SO2, CO and
                               BS) were modeled as a function of lagged weather variables
                               These variables were deseasonalized by regressing on seine
                               and cosine variables. Mortality regression included various
                               lagged and averaged weather and pollution variables.
                               Analyses were conducted in the linear range of
                               mortality/temperature relationship (15 to 0 degrees C).
                               Non-accidental total deaths (56.6 per day) were examined
                               for associations with PM2 5, PM10.2 5, O3, CO, SO2, and NO2.
                               Data analyzed using GAM Poisson regression models,
                               adjusting for temperature, seasonal cycles. Single and two
                               pollutant models with lag days from 0 to 5, as well as the
                               2- to 5-day average concentrations evaluated.

                               Non-accidental total deaths, deaths for age 65 and over, and
                               cause-specific (cardiac, respiratory, and the other
                               remaining) deaths were examined for their associations with
                               PM10, PM2 5, PM10_2 5, O3, and NO2. Data were analyzed
                               using GAM Poisson regression models, adjusting for
                               temperature (average of 1-3 day lags) and seasonal cycles.
                               Individual pollution lag days from 0 to 5, and average
                               concentrations of previous 5 days were considered.

                               Infant mortality (avg. ~ 3/day) related to PM2 5, O3, and
                               NO2, adjusting for temperature and smoothed time, using
                               Poisson GAM model.
                                                      Daily non-accidental mortality was not associated with PM
                                                      indices or gaseous pollutants in the all-year analysis.  However,
                                                      all the PM indices (except coarse particles) were positively and
                                                      significantly associated with non-accidental mortality (age over
                                                      65) in the warm season. Of gaseous pollutants, NO2 and O3
                                                      were positively and significantly associated with non-accidental
                                                      mortality in warm season. Two pollutant models were not
                                                      considered because "so few associations were found".

                                                      Polluted days were found to be colder and less windy and rainy
                                                      than usual. In the regression of mortality on the multiple-lagged
                                                      temperature, wind, rain, humidity, sumshine, SO2, CO, and BS,
                                                      cold temperature was associated with mortality increase, but not
                                                      SO2 or CO. BS suggestive evidence, though not statistically
                                                      significant, of association at 0- and 1-day lag.
                                                      Both PM size fractions associated with mortality, but different
                                                      effects found for warmer and colder months.  PM2 5 and PM10_2 5
                                                      both important in whole year, winter, and summer. In summer,
                                                      PM10_25 had largest effect size estimate. NO2 and CO also
                                                      associated with mortality, as was O3 in warmer months. No
                                                      consistent SO2-mortality associations.

                                                      All three particle size fractions were associated individually with
                                                      mortality.  The effect size estimate was largest for PM10_25. The
                                                      effect size estimate was stronger for respiratory causes than for
                                                      total, cardiovascular,  or other causes of death. The results were
                                                      not sensitive to additions of O3 and NO2.  In the model with
                                                      simultaneous inclusion of PM25 and PM10_25, the effect size for
                                                      PM10_2 5 remained about the same, but the effect size for PM2 5
                                                      became negligible.

                                                      Excess infant mortality associated with PM2 5, NO2, and O3 in the
                                                      same average/lags. NO2 and O3 associations less consistent in
                                                      multi-pollutant models.
                                                       Percent excess mortality for PM10,
                                                       PM25, and PM10.25 (avg. lag 0 and 1
                                                       days) were 0.2% (-1.8, 2.2) per 24.4
                                                       ug/m3 PM10, 0.6% (-1.5, 2.7) per 17.7
                                                       ug/m3 PM25, and -0.6% (-4.2,  2.3) per
                                                       11.3 ug/m3 PM10_25 in all-year analysis.
                                                       The results for season specific  analysis
                                                       were given only as figures.

                                                       3% (95% CI not reported) increase in
                                                       daily mortality per 17.7 ug/m3 of BS
                                                       (lag 0 and 1).
                                                       Percent excess total deaths per 25 ,ug/m3
                                                       increase in the average of previous two
                                                       days for the whole year: 1.8(1.3, 2.4)
                                                       for PM2.5 and 2.3 (1.4, 3.2) for PM10.25
                                                       in single pollutant models.
                                                       Total mortality percent increase
                                                       estimates per increase for average of
                                                       previous 5 days:  9.5 (5.0, 14.2) for
                                                       50 |/g/m3 PM10; 3.7 (0, 7.6) for
                                                       25 Mg/m3 PM25; and 10.5 (6.4, 14.8) for
                                                       25,ug/m3PM10.25.
                                                       Infant mortality excess risk: 18.2% (6.4,
                                                       30.7) per 25 ,ug/m3 PM25 at avg. 3-5 lag
                                                       days.

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
  Study Description: Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates.  Modeling
        methods: lags, smoothing, and covariates.
                  Results and Comments.
     Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                                    PM Index, lag, Excess Risk%
                                                                                                                                                                  (95% LCL, UCL), Co-pollutants.
OO
>
to
          Latin America (cont'd)

          Borja-Aburto et al. (1998).
          Mexico-City,
          1993-1995.
          PM2.5(mean:  27)
          Borja-Aburto et al. (1997).
          Mexico-City,
          1990-1992.
          TSP (median: 204)
Tellez-Rojo et al. (2000).
Mexico City.  1994.
PM10 mean = 75.1.
                               Total, respiratory, cardiovascular, other deaths, and age-
                               specific (age >= 65) deaths were related to PM2 5, O3, and
                               NO2, adjusting for 3-day lagged temperature and periodic
                               cycles, using Poisson GAM model.
Total, respiratory, cardiovascular, and age-specific (age >=
65) deaths were related to O3, TSP, and CO, adjusting for
minimum temperature (temperature also fitted seasonal
cycles) using Poisson models.  The final models were
estimated using the iteratively weighted and filtered least
squares method to account for overdispersion and
autocorrelation.

One year of daily total respiratory and COPD mortality
series were analyzed for their associations with PM10 and
O3 using Poisson model adjusting for cold or warm months,
and 1-day lagged minimum temperature. The data were
stratified by the place of deaths.
PM2 5, O3, and NO2 were associated with mortality with different
lag/averaging periods (1 and 4 day lags; 1-2 avg.; 1-5 avg.,
respectively). PM25 associations were most consistently
significant. SO2 was available, but not analyzed because of its
"low" levels.

O3, SO2, and TSP were all associated with total mortality in
separate models, but in multiple pollutant model, only TSP
remained associated with mortality.  CO association weak.
The average number of daily respiratory deaths, as well as that of
COPD deaths, was similar for in and out of hospital. They found
that the estimated PM10 relative risks were consistently larger for
the deaths that occurred outside medical units.  The results are
apparently  consistent  with  the  assumption that the  extent of
exposure misclassification may  be smaller for those  who died
outside medical units.
                                                                                                                                                               For total excess deaths, 3.4% (0.4, 6.4)
                                                                                                                                                               per 25 ,ug/m3 PM2 5 for both 0 and 4 d
                                                                                                                                                               lags. For respiratory (4 d) = 6.4 (-2.6,
                                                                                                                                                               16.2); for
                                                                                                                                                               CVD(4d) = 5.6(-0.1, 11.5)

                                                                                                                                                               Total deaths:
                                                                                                                                                               6% (3.3, 8.3) per 100 ,ug/m3 TSP at 0 d
                                                                                                                                                               lag.
                                                                                                                                                               CVD deaths:
                                                                                                                                                               5.2% (0.9, 9.9).
                                                                                                                                                               Resp. deaths:
                                                                                                                                                               9.5% (1.3, 18.4).

                                                                                                                                                               Percent excess for total respiratory and
                                                                                                                                                               COPD mortality were 2.9% (0.9, 4.9)
                                                                                                                                                               and 4.1% (1.3, 6.9) per 10 ug/m3
                                                                                                                                                               increase in 3-day lag PM10,
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          Pereiraetal. (1998).
          Sao Paulo, Brazil, 1991-1992.
          PM10 (beta-attenuation, 65)
          Gouveia and Fletcher (2000).
          Sao Paulo, Brazil. 1991-1993.
          PM10mean = 64.3.
          Conceicao et al. (2001).
          Sao Paulo, Brazil. 1994-1997.
          PM10mean = 66.2
                               Intrauterine mortality associations with PM10, NO2, SO2,
                               CO, and O3 investigated using Poisson regression adjusting
                               for season and weather. Ambient CO association with
                               blood carboxyhemoglobin sampled from umbilical cords of
                               non-smoking pregnant mothers studied in separate time
                               period.

                               All non-accidental causes, cardiovascular, and respiratory
                               mortality were analyzed for their associations with air
                               pollution (PM10, SO2, NO2, O3, and CO) using Poisson
                               model adjusting for trend, seasonal cycles, and weather.
                               Potential roles of age and socio-economic status were
                               examined by stratifying data by these factors.


                               Daily respiratory deaths for children under 5 years of age
                               were analyzed for their  associations with air pollution
                               (PM10, SO2, O3, and CO) using GAM Poisson model
                               adjusting for seasonal cycles and weather.
                                                         NO2, SO2, and CO were all individually significant predictor of
                                                         the intrauterine mortality.  NO2 was most significant in multi-
                                                         pollutant model. PM10 and O3 were not significantly associated
                                                         with the mortality. Ambient CO levels were associated with and
                                                         carboxyhemoglobin of blood sampled from the umbilical cords.
                                                         There was an apparent effect modification by age categories.
                                                         Estimated PM10 effects were higher for deaths above age 65
                                                         (highest for the age 85+ category), and no associations were
                                                         found in age group < 65 years. Respiratory excess deaths were
                                                         larger than those for cardiovascular or non-accidental deaths.
                                                         Other pollutants were also associated with the elderly mortality.
                                                         Significant mortality associations were found for CO, SO2, and
                                                         PM10 in single pollutant models. When all the pollutants were
                                                         included, PM10 coefficient became negative and non-significant.
                                                            Intrauterine mortality excess risk: 4.1%
                                                            (-1.8, 10.4) per 50 ,ug/m3 PM10 at 0 day
                                                            lag.
                                                            Percent excess for total non-accidental,
                                                            cardiovascular, and respiratory
                                                            mortality for those with age > 65 were
                                                            3.3% (0.6, 6.0), 3.8% (0.1, 7.6), and 6.0
                                                            (0.5, 11.8), respectively, per 64.2 ug/m3
                                                            increase in PM10 (0-, 0-, and 1-day lag,
                                                            respectively).

                                                            Percent excess for child (age < 5)
                                                            respiratory deaths:  9.7% (1.5, 18.6) per
                                                            66.2 ug/m3 PM10 (2-day lag) in single
                                                            pollutant model.

-------
                  TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE  MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
IQRin,ug/m3.
  Study Description:  Outcomes, Mean outcome rate, and
   ages. Concentration measures or estimates. Modeling
        methods: lags, smoothing, and covariates.
                  Results and Comments.
     Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                                  PM Index, lag, Excess Risk%
                                                                                                                                                                 (95% LCL, UCL), Co-pollutants.
OO
>
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Australia

          Morgan etal. (1998).
          Sydney, 1989-1993.
          Nephelometer (0.30
          bscat/104m).
          Site-specific conversion:
          PM25 = 9; PM10 = 18

          Simpson etal. (1997).
          Brisbane, 1987-1993.
          PM10 (27, not used in analysis).
          Nephelometer
          (0.26 bscat/104m,
          size range: 0.01-2 ,um).
Asia

Hong etal. (1999).
Inchon, South Korea,
1995-1996 (20 months).
PM10mean = 71.2.
          Lee etal. (1999).
          Seoul and Ulsan, Korea,
          1991-1995. TSP(beta
          attenuation, 93 for Seoul
          and 72 for Ulsan)
                               Total, cardiovascular, and respiratory deaths were related to
                               PM (nephelometer), O3, and NO2, adjusting for seasonal
                               cycles, day-of-week, temperature, dewpoint, holidays, and
                               influenza, using Poisson GEE to adjust for autocorrelation.
                               Total, cardiovascular, and respiratory deaths (also by age
                               group) were related to PM (nephelometer), O3, SO2, and
                               NO2, adjusting for seasonal cycles, day-of-week,
                               temperature, dewpoint, holidays, and influenza, using
                               Poisson GEE to adjust for autocorrelation. Season-specific
                               (warm and cold) analyses were also conducted.
Non-accidental total deaths, cardiovascular, and respiratory
deaths were examined for their associations with PM10, O3,
SO2, CO, and NO2. Data were analyzed using GAM
Poisson regression models, adjusting for temperature,
relative humidity, and seasonal cycles. Individual pollution
lag days from 0 to 5, as well as the average concentrations
of previous 5 days were considered.

Total mortality series was examined for its association with
TSP, SO2, and O3, in Poisson GEE (exchangeable
correlation for days in the same year), adjusting for season,
temperature, and humidity.
                                                        PM, O3, and NO2 all showed significant associations with total
                                                        mortality in single pollutant models. In multiple pollutant
                                                        models, the PM and O3 effect estimates for total and
                                                        cardiovascular deaths were marginally reduced, but the PM
                                                        effect estimate for respiratory deaths was substantially reduced.
                                                        Same-day PM and O3 were associated most significantly with
                                                        total deaths. The O3 effect size estimates for cardiovascular and
                                                        respiratory deaths were consistently positive (though not
                                                        significant), and larger in summer. PM's effect size estimates
                                                        were comparable for warm and cold season for cardiovascular
                                                        deaths, but larger in warm season for respiratory deaths. NO2
                                                        and SO2 were not associated with mortality.
A greater association with mortality was seen with the 5-day
moving average and the previous day's exposure than other
lag/averaging time. In the models that included a 5-day moving
average of one or multiple pollutants, PM10 was a significant
predictor of total mortality, but gaseous pollutants were not
significant.  PM10 was also a significant predictor of
cardiovascular and respiratory mortality.

All the pollutants were significant predictors of mortality in
single pollutant models. TSP was not significant in multiple
pollutant models, but SO2 and O3 remained significant.
                                                            4.7% (1.6, 8.0) per 25 ,ug/m3 estimated
                                                            PM25 or 50 |/g/m3 estimated PM10 at
                                                            avg. ofO and 1 day lags.
                                                            (Note: converted from nephelometry
                                                            data)
                                                            3.4% (0.4, 6.4) per 25 ,ug/m3 1-h PM25
                                                            increment at 0 d lag; and 7.8% (2.5,
                                                            13.2) per 25 ^g/m3 24-h PM25
                                                            increment.
                                                                                                                                                             Percent excess deaths (t-ratio) per 50
                                                                                                                                                             //g/m3 increase in the 5-day moving
                                                                                                                                                             average of PM10: 4.1 (0.1, 8.2) for total
                                                                                                                                                             deaths; 5.1 (0.1, 10.4) for
                                                                                                                                                             cardiovascular deaths; 14.4 (-3.2, 35.2)
                                                                                                                                                             for respiratory deaths.
                                                                                                                                                   5.1% (3.1, 7.2) for Seoul, and -0.1% (-
                                                                                                                                                   3.9, 3.9) for Ulsan, per lOO^g/m3 TSP
                                                                                                                                                   at avg. of 0, 1, and 2 day lags.
O

-------
               TABLE 8A-1 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
to
o
o
to
Reference, Location, Years,
PM Index, Mean or Median,
Study Description: Outcomes, Mean outcome rate, and
ages. Concentration measures or estimates. Modeling
     methods: lags, smoothing, and covariates.
             Results and Comments.
Design Issues, Uncertainties, Quantitative Outcomes.
                                                                                                                                                            PM Index, lag, Excess Risk%
                                                                                                                                                          (95% LCL, UCL), Co-pollutants.
OO
>
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           Asia (cont'd)

           Lee and Schwartz (1999).
           Seoul, Korea. 1991-1995.
           TSPmean = 925.
          Xu et al. (2000).
          Shenyang, China, 1992
          TSP (430).
           Ostroetal. (1998).
           Bangkok, Thailand,
           1992-1995
           PM10 (beta attenuation, 65)
                              Total deaths were analyzed for their association with TSP,
                              SO2, and O3. A conditional logistic regression analysis with
                              a case-crossover design was conducted. Three-day moving
                              average values (current and two past days) of TSP and SO2,
                              and l-hrmaxO3 were analyzed separately. The control
                              periods are 7 and 14 days before and/or after the case
                              period. Both unidirectional and bi-directional controls (7 or
                              7 and 14 days) were examined, resulting in six sets
                              of control selection schemes. Other covariates included
                              temperature and relative humidity.

                              Total (non-accidental), CVD, COPD, cancer and other
                              deaths examined for their associations with TSP and
                              SO2,using Poisson (GAM, and Markov approach to adjust
                              for mortality serial dependence) models, adjusting for
                              seasonal cycles, Sunday indicator, quintiles of temp, and
                              humidity.  Ave. pollution values of concurrent and
                              3 preceding days used.
                              Total (non-accidental), cardiovascular, respiratory deaths
                              examined for associations with PM10 (separate
                              measurements showed =50% of PM10 was PM25),using
                              Poisson GAM model adjusting for seasonal cycles, day-of-
                              week, temp., humidity.
                                                     Among the six control periods, the two unidirectional
                                                     retrospective control schemes resulted in odds ratios less than
                                                     1; the two unidirectional prospective control schemes resulted
                                                     in larger odds ratios (e.g., 1.4 for 50 ppb increase in SO2); and
                                                     bi-directional control schemes resulted in odds ratios between
                                                     those for uni-directional schemes.  SO2 was more significantly
                                                     associated with mortality than TSP.
                                                     Total deaths were associated with TSP and SO2 in both single
                                                     and two pollutant models.  TSP was significantly associated
                                                     with CVD deaths, but not with COPD.  SO2 significantly
                                                     associated with COPD, but not with CVD deaths. Cancer
                                                     deaths not associated with TSP or SO,.
                                                     All the mortality series were associated with PM10 at various
                                                     lags. The effects appear across all age groups. No other
                                                     pollutants were examined.
                                                     OR for non-accidental mortality
                                                     per 100 ,ug/m3 increase in 3-day
                                                     average TSP was 1.010 (0.988,
                                                     1.032).
                                                     Percent total excess deaths per
                                                     100 /^g/m3 increase in 0-3 day
                                                     ave. of TSP = 1.75(0.65,2.85);
                                                     with SO2 = 1.31(0.14, 2.49)
                                                     COPD TSP = 2.6 (-0.58, 5.89);
                                                     with SO2 = 0.76 (-2.46, 4.10).
                                                     CVD TSP = 2.15 (0.56, 3.71);
                                                     with SO2 = 1.95 (1.19, 3.74).
                                                     Cancer TSP = 0.87 (-1.14,
                                                     2.53); with SO2 = 1.07 (-1.05,
                                                     3.23).
                                                     Other deaths TSP = 3.52 (0.82,
                                                     6.30); with SO2 = 2.40 (-0.51,
                                                     5.89).

                                                     Total mortality excess risk: 5.1%
                                                     (2.1, 8.3) per 50 ,ug/m3 PM10 at
                                                     3 d lag (0 and 2 d lags also
                                                     significant).
                                                     CVD (3d ave.) = 8.3 (3.1, 13.8)
                                                     Resp. (3d ave.) = 3.0 (-8.4,
                                                     15.9)
O

-------
                TABLE 8A-1  (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE MORTALITY EFFECTS STUDIES
 to
 o
 o
 to
Reference, Location, Years,
PM Index, Mean or Median,
IQRin//g/m3.
Study Description: Outcomes, Mean outcome rate, and
ages. Concentration measures or estimates. Modeling
     methods: lags, smoothing, and covariates.
             Results and Comments.
Design Issues, Uncertainties, Quantitative Outcomes.
  PM Index, lag, Excess Risk%
(95% LCL, UCL), Co-pollutants.
 OO
 to
 OO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           Asia (cont'd)

           Cropper etal. (1997).
           Delhi, India, 1991-1994
           TSP(375)
           Kwonetal. (2001).
           Seoul, South Korea,
           1994-1998.
           PM10 mean = 68.7.
           Lee et al. (2000).
           Seven major cities, Korea.
           1991-1997.
           TSP mean = 77.9.
                              Total (by age group), respiratory and CVD deaths related to
                              TSP, SO2, and NOx, using GEE Poisson model (to control
                              for autocorrelation), adjusting for seasonal cycles
                              (trigonometric terms), temperature, and humidity.  70%
                              deaths occur before age 65 (in U.S., 70% occur after age
                              65).

                              The study was planned to test the hypothesis that patients
                              with congestive heart failure are more susceptible to the
                              harmful effects of ambient air pollution than the general
                              population. GAM Poisson regression models, adjusting for
                              seasonal cycles, temperature, humidity, day-of-week, as
                              well as the case-crossover design, with 7 and 14 days before
                              and after the case period, were applied
                              All non-accidental deaths were analyzed for their
                              associations with TSP, SO2, NO2, O3, and CO using GAM
                              Poisson model adjusting for trend, seasonal cycles, and
                              weather. Pollution relative risk estimates were obtained for
                              each city, and then pooled.
                                                      TSP was significantly associated with all mortality series
                                                      except with the very young (age 0-4) and the "very old" (age
                                                      >=65).  The results were reported to be unaffected by addition
                                                      of SO2 to the model. The authors note that, because those who
                                                      are affected are younger (than Western cities), more life-years
                                                      are likely to be lost per person from air pollution impacts.

                                                      The estimated effects were larger among the congestive heart
                                                      failure patients than among the general population (2.5 — 4.1
                                                      times larger depending on the pollutants). The case-crossover
                                                      analysis showed similar results.  In two pollutant models, the
                                                      PM10 effects were much lower when CO, NO2,  or SO2 were
                                                      included. O3  had little impact on the effects of the other
                                                      pollutants.
                                                      In the results of pooled estimates for multiple pollutant models,
                                                      the SO2 relative  risks were not affected by addition of other
                                                      pollutants,  whereas the relative risks  for other pollutants,
                                                      including TSP, were. The SO2 levels in these Korean cities were
                                                      much higher than the levels observed in the current U.S.  For
                                                      example, the 24-hr mean SO2 levels in the Korean cities ranged
                                                      from 12.1 to 31.4 ppb, whereas, in Samet et al.'s 20 largest U.S.
                                                      cities, the range of 24-hr mean SO2 levels were 0.7 to  12.8 ppb.
                                                      2.3% (significant at 0.05, but SE
                                                      of estimate not reported) per 100
                                                      Mg/m3 TSP at 2 day lag.
                                                      The RRs for PM10 (same-day)
                                                      using the GAM approach for the
                                                      general population and for the
                                                      cohort with congestive heart
                                                      failure were 1.4% (0.6, 2.2) and
                                                      5.8 (-1.1, 13.1), respectively, per
                                                      42.1ug/m3. Corresponding
                                                      ORs using the case-crossover
                                                      approach were 0.1% (-0.9, 1.2)
                                                      and 7.4% (-2.2, 17.9),
                                                      respectively.

                                                      Percent excess deaths for all non-
                                                      accidental deaths was 1.7% (0.8,
                                                      2.6) per 100 ug/m3 2-day moving
                                                      average TSP.
 O

-------
                  APPENDIX 8B
  PARTICULATE MATTER-MORBIDITY STUDIES:
               SUMMARY TABLES
April 2002                 8B-1     DRAFT-DO NOT QUOTE OR CITE

-------
         Appendix 8B.1: PM-Cardiovascular Admissions Studies
April 2002                         8B-2      DRAFT-DO NOT QUOTE OR CITE

-------
               TABLE 8B-1.  ACUTE PARTICIPATE MATTER EXPOSURE AND  CARDIOVASCULAR HOSPITAL ADMISSIONS
 to
 o
 o
 to
           Reference citation.  Location, Duration
           PM Index, Mean or Median, IQR
                                          Study Description:  Health outcomes or codes.
                                          Mean outcome rate, sample or population size, ages.
                                          Concentration measures or estimates.
                                          Modeling methods: lags, smoothing, co-pollutants,
                                          covariates, concentration-response
Results and Comments. Design Issues,
Uncertainties, Quantitative Outcomes
PM Index, Lag, Excess
Risk % (95% LCL, UCL),
Co-Pollutants
 OO
 td
UnitedStates

Samet et al. (2000a,b)
14 US cities
1985-1994, but range of years varied by city

PM10 (,ug/m3) mean, median, IQR:
Birmingham, AL:  34.8, 30.6, 26.3
Boulder, CO: 24.4, 22.0, 14.0
Canton, OH: 28.4, 25.6, 15.3
Chicago, II:  36.4, 32.6, 22.4
Colorado Springs,  CO:  26.9, 22.9, 11.9
Detroit, MI:  36.8,  32.0, 28.2
Minneapolis/St. Paul, MN:  27.4, 24.1, 17.9
Nashville, TN: 31.6, 29.2, 17.9
New Haven, CT: 29.3,26.0,20.2
Pittsburgh, PA: 36.0, 30.5,  27.4
Provo/Orem, UT: 38.9, 30.3, 22.8
Seattle, WA: 31.0,26.7,20.0
Spokane, WA: 45.3,36.2,33.5
Youngstown, OH:  33.1, 29.4, 18.6
                                                     Daily medicare hospital admissions for total
                                                     cardiovascular disease, CVD (ICD9 codes 390-429), in
                                                     persons 65 or greater. Mean CVD counts ranged from 3 to
                                                     102/day in the 14 cities. Covariates: SO2, NO2, O3, CO,
                                                     temperature, relative humidity, barometric pressure. Stats:
                                                     In first stage, performed city-specific, PM10-ONLY,
                                                     generalized additive robust Poisson regression with
                                                     seasonal, weather, and day of week controls.  Repeated
                                                     analysis for days with PM10 less than 50 ,ug/m3 to test for
                                                     threshold. Lags of 0-5 considered, as well as the quadratic
                                                     function of lags 0-5. Individual cities analyzed first.  The
                                                     14 risk estimates were then analyzed in several second
                                                     stage analyses: combining risks across cities using inverse
                                                     variance weights, and regressing risk estimates on
                                                     potential effect-modifiers and slopes of PM10 on co-
                                                     pollutants.
City-specific risk estimates for a 10 i/g/m3
increase in PM10 ranged from -1.2% in Canton to
2.2% in Colorado Springs. Across-city weighted
mean risk estimate was largest at lag 0,
diminishing rapidly at other lags.  Only the mean
of lags 0 and 1 was significantly associated with
CVD. There was no evidence of statistical
heterogeniety in risk estimates across cities for
CVD. City-specific risk estimates were not
associated with the percent of the population that
was non-white, living in poverty, college
educated, nor unemployed. No evidence was
observed that PM10 effects were modified by
weather. No association was observed between
the city-specific PM10 risk estimates and the city-
specific correlation between PM10 and co-
pollutants. However, due to the absence of multi-
pollutant regression results, it is not clear
whether this study demonstrates an independent
effect of PM10.
Percent Excess CVD Risk (95% CI),
combined over cities per 50 ,ug/m3
change in PM10.

PM10: Odlag.
 5.5% (4.7, 6.2)
PM10: 0-Id lag.
 6.0% (5.1,6.8)
PM10 < 50 Mg/m3: 0-Id lag.
 7.6% (6.0, 9.1)
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
o

-------
                            TABLE 8B-1 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                                HOSPITAL ADMISSIONS
to
o
o
to
           Reference citation. Location, Duration
           PM Index, Mean or Median, IQR
                                           Study Description:  Health outcomes or codes,
                                           Mean outcome rate, sample or population size,
                                           ages.  Concentration measures or estimates.
                                           Modeling methods: lags, smoothing, co-pollutants,
                                           covariates, concentration-response
Results and Comments. Design Issues,
Uncertainties, Quantitative Outcomes
                                                                                                                                                            PM Index, Lag, Excess
                                                                                                                                                            Risk % (95% LCL, UCL),
                                                                                                                                                            Co-Pollutants
OO
td
 fe
 H
 6
 o

 o
 H
O
 O
 H
 W
 O
 O
United States (cont'd)

Janessen et al. (2002)
14 U.S. cities studied in Samet et al.
(2000a,b) above

                     Mean
                  Summer/Winter   Ratio
Birmingham
Boulder*
Canton
Chicago
Colorado Springs*
Detroit
Minneapolis
Nashville
New Haven
Pittsburgh
Seattle*
Spokane*
Provo-Urem*
Youngstown
40.0/27.4
26.8/36.3
36.6/25.8
42.5/30.4
21.3/37.3
42.8/32.8
30.5/23.0
40.1/31.9
30.3/31.6
46.6/29.4
23.8/43.3
32.7/42.2
31.4/66.3
40.7/30.1
Zanobetti et al. (2000b)
10 US cities
1986-1994

PM10 (Mg/m3) median, IQR:
Canton, OH: 26, 15
Birmingham, AL:  31,26
Chicago, II:  33,23
Colorado Springs, CO: 23, 13
Detroit, MI: 32, 28
Minneapolis/St. Paul, MN: 24, 18
New Haven, CT:  26,21
Pittsburgh, PA: 30, 28
Seattle, WA: 27,21
Spokane, WA:  36, 34
                                   0.69
                                   1.35
                                   0.70
                                   0.71
                                   1.75
                                   0.77
                                   0.75
                                   0.80
                                   1.04
                                   0.63
                                   1.82
                                   1.29
                                   2.11
                                   0.74
                                                      Examined same database as Samet et al. (2000a,b) to
                                                      evaluate whether differences in prevalence in air
                                                      conditioning (AC) and/or the contribution of different
                                                      sources to total PM10 emissions could partially explain the
                                                      observed variability in exposure effect relations. Variables
                                                      included 24-hr means of temperature.  Cities were
                                                      characterized and analyzed as either winter or nonwinter
                                                      peaking. Rations between mean concentrations during
                                                      summer (June, July August) and winter (January,
                                                      February, March) were calculated. ('Winter peaking PM10
                                                      concentration.)
                                                      Derived from the Samet et al. (2000a,b) study, but for a
                                                      subset of 10 cities. Daily hospital admissions for total
                                                      cardiovascular disease, CVD (ICD9 codes 390-429), in
                                                      persons 65 or greater. Median CVD counts ranged from 3
                                                      to 103/dayinthe 10 cities. Covariates: SO2, O3, CO,
                                                      temperature, relative humidity, barommetric pressure.
                                                      Stats:  In first stage, performed single-pollutant generalized
                                                      additive robust Poisson regression with seasonal, weather,
                                                      and day of week controls. Repeated analysis for days with
                                                      PM10 less than 50 ,ug/m3 to test for threshold. Lags of 0-5
                                                      considered, as well as the quadratic function of lags 0-5.
                                                      Individual cities analyzed first.  The 10 risk estimates were
                                                      then analyzed in several second stage analyses:  combining
                                                      risks across cities using inverse variance weights, and
                                                      regressing risk estimates on potential effect-modifiers and
                                                      pollutant confounders.
Analysis of city groups of winter peaking, PM10
and nonwinter peaking PM10 yielded coefficients
for CVD-related hospitalization admissions that
decreased significantly with increasing
percentage of central AC for both city groups.
Four source related variables coefficients for
hospital admissions for CVD increased
significantly with increasing percentage of PM10
from highway vehicles, highway diesels, oil
combustion,  metal processing, increasing
population, and vehicle miles traveled (VMT)
per sq mg and with decreasing percentage of
PM10 from fugitive dust.  For COPD and
pneumonia association were less significant but
the pattern of association were similar to that for
CVD.
                                                                                                  Same basic pattern of results as in Samet et al.
                                                                                                  (2000a,b).  For distributed lag analysis, lag 0 had
                                                                                                  largest effect, lags 1 and 2 smaller effects, and
                                                                                                  none at larger lags. City-specific slopes were
                                                                                                  independent of percent poverty and percent non-
                                                                                                  white. Effect size increase when data were
                                                                                                  restricted to days with PM10 less than 50 ,ug/m3.
                                                                                                  No multi-pollutant models reported; however, no
                                                                                                  evidence of effect modification by co-pollutants
                                                                                                  in second stage analysis.  As with Samet et al.
                                                                                                  2000., it is not clear whether this study
                                                                                                  demonstrates an independent effect of PM10.
                                                                                                                                                            Homes with AC
                                                                                                                                                            PCVD
                                                                                                                                                            % change (SE)

                                                                                                                                                            All cities
                                                                                                                                                            -15.2(14.8)
                                                                                                                                                            Nonwinter peak cities
                                                                                                                                                            -50.3" (17.4)
                                                                                                                                                            Winter peak cities
                                                                                                                                                            -51.7" (13.8)
                                                                                                                                                            Source PM10 from
                                                                                                                                                              highway vehicles
                                                                                                                                                            % change (SE) p CVD
                                                                                                                                                            58.0*(9.9)
                                                                                                                                                            [**p<0.05]
                                              Percent Excess Risk (SE) combined
                                              over cities:
                                              Effects computed for 50 ,ug/m3 change
                                              in PM10.

                                              PM10: Od.
                                               5.6 (4.7, 6.4)
                                              PM10: 0-1 d.
                                               6.2(5.4,7.0)
                                              PM10 < 50//g/m3:  0-ld.
                                               7.8 (6.2, 9.4)

-------
to
o
o
to
                            TABLE 8B-1 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                               HOSPITAL ADMISSIONS
           Reference citation. Location, Duration
           PM Index, Mean or Median, IQR
Study Description: Health outcomes or codes,
Mean outcome rate, sample or population size,
ages.  Concentration measures or estimates.
Modeling methods: lags, smoothing, co-pollutants,
covariates, concentration-response
Results and Comments. Design Issues,
Uncertainties, Quantitative Outcomes
PM Index, Lag, Excess
Risk % (95% LCL, UCL),
Co-Pollutants
OO
>
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           United States (cont'd)

           Schwartz (1999)
           8 US metropolitan counties
           1988-1990
           median, IQR for PM10 (//g/m3):
           Chicago, IL: 35, 23
           Colorado Springs, CO: 23, 14
           Minneapolis, MN:  28, 15
           New Haven, CT: 37,25
           St. Paul,  MN: 34, 23
           Seattle, WA: 29,20
           Spokane, WA: 37,33
           Tacoma, WA:  37,27
           Linn et al. (2000)
           Los Angeles
           1992-1995
           mean,  SD:
           PM10est(,ug/m3):45, 18
Daily hospital admissions for total cardiovascular diseases
(ICD9 codes 390-429) among persons over 65 years.
Median daily hospitalizations: 110, 3, 14, 18, 9, 22, 6, 7,
alphabetically by city. Covariates: CO, temperature,
dewpoint temp.  Stats: robust Poisson regression after
removing admission outliers; generalized additive models
with LOESS smooths for control of trends, seasons, and
weather.  Day of week dummy variables. Lag 0 used for
all covariates.
Hospital admissions for total cardiovascular diseases
(CVD), congestive heart failure (CHE), myocardial
infarction (MI), cardiac arrhythmia (CA) among all
persons 30 years and older, and by sex, age, race, and
season. Mean hospital admissions for CVD: 428.
Covariates: CO, NO2, O3, temperature, rainfall.  Daily
gravimetric PM10 estimated by regression of every sixth
day PM10 on daily real-time PM10 data collected by
TEOM.  Poisson regression with controls for seasons and
day of week.  Reported results for lag 0 only. Results
reported as Poisson regression coefficients and their
standard errors. The number of daily CVD admissions
associated  with the mean PM10 concentration can be
computed by multiplying the PM10 coefficient by the PM10
mean and then exponentiating.  Percent effects are
calculated  by dividing this result by the mean daily
admission  count for CVD.
In single-pollutant models, similar PM10 effect
sizes obtained for each county. Five of eight
county-specific effects were statistically
significant, as was the PM10 effect pooled across
locations.  CO effects significant in six of eight
counties. The PM10 and CO effects were both
significant in a two pollutant model that was run
for five counties where the PM10/CO correlation
was less than 0.5. Results reinforce those of
Schwartz, 1997.
In year-round, single-pollutant models,
significant effects of CO, NO2, and PM10
on CVD were reported. PM10 effects appeared
larger in winter and fall than in spring and
summer.  No consistent differences in PM10
effects across sex, age, and race. CO risk was
robust to including PM10 in the model; no results
presented on PM10 robustness to co-pollutants.
Percent Excess Risk (95% CI):
Effects computed for 50 ,ug/m3 change
in PM10.

PM10: Od.
Individual counties:
Chicago: 4.7(2.6,6.8)
COSpng:  5.6 (-6.8, 19.0)
Minneap: 4.1 (-3.6, 12.5)
NewHav:  5.8(2.1,9.7)
St. Paul: 8.6 (2.9, 14.5)
Seattle:  3.6 (-0.1, 7.4)
Spokane: 6.7 (0.9, 12.8)
Tacoma:  5.3(3.1,7.6)

Pooled:  5.0(3.7,6.4)
3.8(2.0, 5.5) w. CO

% increase with PM10 change of
50 ,ug/m3:

PM10est:  Od.
CVD ages 30+
 3.25% (2.04, 4.47)

MI ages  30+
 3.04% (0.06, 6.12)

CHE ages 30+
 2.02% (-0.94, 5.06)

CA ages 30+
 1.01% (-1.93, 4.02)
o

-------
to
o
o
to
                           TABLE 8B-1 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                              HOSPITAL ADMISSIONS
           Reference citation. Location, Duration
           PM Index, Mean or Median, IQR
                                          Study Description: Health outcomes or codes,
                                          Mean outcome rate, sample or population size,
                                          ages.  Concentration measures or estimates.
                                          Modeling methods: lags, smoothing, co-pollutants,
                                          covariates, concentration-response
                                                      Results and Comments. Design Issues,
                                                      Uncertainties, Quantitative Outcomes
                                             PM Index, Lag, Excess
                                             Risk % (95% LCL, UCL),
                                             Co-Pollutants
OO
td
           United States (cont'd)

           Morris and Naumova (1998)
           Chicago, IL
           1986-1989
           mean, median, IQR, 75th percentile:
           PM10 Cug/m3):  41,38,23,51
Schwartz (1997)
Tucson, AZ
1988-1990
mean, median, IQR:
PM10 (,ug/m3): 42, 39, 23
Daily hospital admissions for congestive heart failure,
CHF (ICD9 428), among persons over 65 years. Mean
hospitalizations: 34/day.  Covariates: O3, NO2, SO2,
CO, temperature, relative humidity.  Gases measured at up
to eight sites; daily PM10 measured at one site. Stats:
GLM for time series data. Controlled for trends and cycles
using dummy variables for day of week, month, and year.
Residuals were modeled as negative binomial distribution.
Lags of 0-3 days examined.

Daily hospital admissions for total cardiovascular diseases
(ICD9 codes 390-429) among persons over 65 years.
Mean hospitalizations:  13.4/day. Covariates: O3, NO2,
CO, SO2, temperature, dewpoint temperature. Gases
measured at multiple sites; daily PM10 at one site.  Stats:
robust Poisson regression; generalized additive model with
LOESS smooth for controlling trends and seasons, and
regression splines to control weather. Lags of 0-2 days
examined.
                                                                                                CO was only pollutant statistically significant in
                                                                                                both single- and multi-pollutant models.
                                                                                                Exposure misclassification may have been larger
                                                                                                for PM10 due to single site. Results suggest
                                                                                                effects of both CO and PM10 on congestive heart
                                                                                                failure hospitalizations among  elderly, but CO
                                                                                                effects appear more robust.
Both PM10 (lag 0) and CO significantly and
independently associated with admissions,
whereas other gases were not. Sensitivity
analyses reinforced these basic results. Results
suggest independent effects of both PM10 and
CO for total cardiovascular hospitalizations
among the elderly.
                                             Percent Excess Risk (95% CI)
                                             per 50 ,ug/m3 change in PM10.

                                             PM10: Od.
                                              3.92% (1.02, 6.90)
                                              1.96% (-1.4, 5.4) with
                                                4 gaseous pollutants
                                                                                                                                                        Percent Excess Risk (95% CI)
                                                                                                                                                        per 50 Mg/m3 change in PM10.

                                                                                                                                                        PM10: Od.
                                                                                                                                                         6.07% (1.12, 1.27)
                                                                                                                                                         5.22% (0.17, 10.54) w. CO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           Gwynn et al (2000)
           Buffalo, NY
           mn/max
           PM10 = 24.1/90.8,ug/m3
           SO4- = 2.4/3.9
           H+ = 36.4/38.2 nmol/m3
           CoH = 0.2/0.9 10'3 ft
                                          Air pollution health effects associations with total,
                                          respiratory, and CVD hospital admissions (HA's)
                                          examined using Poisson model controlling for weather,
                                          seasonality, long-wave effects, day of week, holidays.
                                                      Positive, but non-significant assoc. found
                                                      between all PM indices and circulatory hospital
                                                      admissions. Addition of gaseous pollutants to
                                                      the model had minimal effects on the PM RR
                                                      estimates.
                                             Percent excess CVD HA risks (95%
                                             CI) per PM10 = 50 ,ug/m3; SO4 =
                                             15,ug/m3;
                                             H+ = 75 nmoles/m3; COH = 0.5
                                             units/1,000 ft:
                                             PM10 (lag 3) = 5.7% (-3.3, 15.5)
                                             SO4 (lag 1) = 0.1% (-0.1, 0.4)
                                             H+(lag 0) = 1.9% (-0.3, 4.2)
                                             COH (lag 1) = 2.2% (-1.9, 6.3)
o

-------
 to
 o
 o
 to
                           TABLE 8B-1 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                             HOSPITAL ADMISSIONS
           Reference citation.  Location, Duration
           PM Index, Mean or Median, IQR ,ug/m3
Study Description: Health outcomes or codes,
Mean outcome rate, sample or population size,
ages. Concentration measures or estimates.
Modeling methods: lags, smoothing, co-pollutants,
covariates, concentration-response
Results and Comments. Design Issues,
Uncertainties, Quantitative Outcomes
PM Index, Lag, Excess
Risk % (95% LCL, UCL),
Co-Pollutants
 OO
 td
           United States (cont'd)

           Lippmann et al. (2000)
           Detroit, MI
           1992-1994
           mean, median, IQR:
           PM25(Mg/m3): 18, 15, 11
           PM10(,ug/m3):31, 28, 19
           PM10.25(//g/m3):13, 12,9
Various cardiovascular (CVD)-related hospital admissions
(HA's) for persons 65+ yr. analyzed, using GAM Poisson
models, adjusting for season, day of week, temperature,
and relative humidity. The air pollution variables analyzed
were: PM10, PM2 5, PM10.2 5, sulfate, H+, O3, SO2, NO2, and
CO. However, this study site/period had very low acidic
aerosol levels. As noted by the authors 85% of H+ data
was below detection limit (8 nmol/m3).
For heart failure, all PM metrics yielded
significant associations. Associations for IHD,
dysrhythmia, and stroke were positive but
generally non-sig. with all PM indices. Adding
gaseous pollutants had negligible effects on
various PM metric RR estimates.  The general
similarity of the PM25 and PM10_25 effects per
,ug/m3 in this study suggest similarity in human
toxicity of these two inhalable mass components
in study locales/periods where PM2 5 acidity not
usually present.  However, small sample size
limits power to distinguish between pollutant-
specific effects.
Percent excess CVD HA risks (95%
CI) per 50 //g/m3 PM10, 25 ,ug/m3
PM25andPM10.25:
IHD:
  PM25(lag2)4.3(-1.4, 10.4)
  PM10 (lag 2) 8.9 (0.5, 18.0)
  PM10.2.5 (lag 2) 10.5 (2.7, 18.9)
Dysrhythmia:
  PM25 (lag 1)3.2 (-6.5, 14.0)
  PM10 (lag 1)2.9 (-6.8, 13.7)
  PM10.2.5 (lag 0)0.2 (-12.2, 14.4)
Heart Failure:
  PM25 (lag 1)9.1(2.4,  16.2)
  PM10 (lag 0)9.7 (0.2, 20.1)
  PM10.2.5 (lag 0)5.2 (-3.3, 14.5)
Stroke:
  PM25 (lag 0)1.8 (-5.3, 9.4)
  PM10 (lag 1)4.8 (-5.5, 16.2)
  PM10.25 (lag 1)4.9 (-4.7, 15.5)
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
O

-------
                           TABLE 8B-1 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                              HOSPITAL ADMISSIONS
 to
 o
 o
 to
           Reference citation. Location, Duration
           PM Index, Mean or Median, IQR
Study Description: Health outcomes or codes,
Mean outcome rate, sample or population size,
ages. Concentration measures or estimates.
Modeling methods: lags, smoothing, co-pollutants,
covariates, concentration-response
Results and Comments. Design Issues,
Uncertainties, Quantitative Outcomes
PM Index, Lag, Excess
Risk % (95% LCL, UCL),
Co-Pollutants
 OO
 td
 OO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           United States (cont'd)

           Moolgavkar (2000b)
           Three urban counties: Cook, IL; Los
           Angeles, CA; Maricopa, AZ.
           1987-1995

           Pollutant median, IQR:
           Cook: PM10: 35, 22
           LA: PM10: 44, 26
               PM25: 22, 16
           Maricopa: PM10: 41, 19
Analysis of daily hospital admissions for total
cardiovascular diseases, CVD, (ICD9 codes 390-429) and
cerebrovascular diseases, CRD, (ICD9 430-448) among
persons aged 65 and over. For Los Angeles, a second age
group, 20-64, was also analyzed.  Median daily CVD
admissions were 110, 172, and 33 in Cook, LA, and
Maricopa counties, respectively.  PM10 available only
every sixth day in LA and Maricopa counties. In LA,
every-sixth-day PM2 5 also was available.  Covariates: CO,
NO2, O3, SO2, temperature, relative humidity.  Stats:
generalized additive Poisson regression, with controls for
day of week and smooth temporal variability.  Single-
pollutant models estimated for individual lags from 0 to 5.
Two-pollutant models also estimated, with both pollutants
at same lag.
In single-pollutant models in Cook and LA
counties, PM was significantly associated with
CVD admissions at lags 0, 1, and 2, with
diminishing effects over lags. PM2 5 also was
significant in LA for lags 0 and 1.  For the 20-64
year old age group in LA, risk estimates were
similar to those for 65+. In Maricopa county, no
positive PM10 associations were observed at any
lag. In two-pollutant models in Cook and LA
counties, the PM10/PM25 risk estimates
diminished and/or were rendered non-
significant. Little evidence observed for
associations between CRD admissions and PM.
These results suggest that PM is not
independently associated with CVD or CRD
hospital admissions.
Percent Excess CVD Risk (95% CI)
Effects computed for 50 ,ug/m3 change
in PM10 and 25 ,ug/m3 change in
PM25.

Cook 65+:
PM10, 0 d.
 4.2(3.0,5.5)
PM10, 0 d. w/NO2.
 1.8(0.4,3.2)
LA 65+:
PM10, 0 d.
 3.2(1.2,5.3)
PM10, 0 d. w/CO
 -1.8 (-4.4, 0.9)

PM25, Od.
 4.3(2.5,6.1)
PM25, Od. w/CO
 0.8 (-1.3, 2.9)
LA 20-64 years  old:
PM10, 0 d.
 4.4 (2.2, 6.7)
PM10, 0 d. w/CO
 1.4 (-1.3, 4.2)

PM25, Od.
 3.5(1.8,5.3)
PM25, Od., w/CO)
 2.3 (-0.2, 4.8)
Maricopa:
PM10, 0 d.
 -2.4 (-6.9, 2.3)
O

-------
                            TABLE 8B-1 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                                HOSPITAL ADMISSIONS
 to
 o
 o
 to
           Reference citation.  Location, Duration
           PM Index, Mean or Median, IQR
                                           Study Description: Health outcomes or codes,
                                           Mean outcome rate, sample or population size,
                                           ages. Concentration measures or estimates.
                                           Modeling methods: lags, smoothing, co-pollutants,
                                           covariates,  concentration-response
                                                       Results and Comments. Design Issues,
                                                       Uncertainties, Quantitative Outcomes
                                              PM Index, Lag, Excess
                                              Risk % (95% LCL, UCL),
                                              Co-Pollutants
 OO
 td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
United States (cont'd)

Zanobetti et al. (2000a)
Cook County, IL
1985-1994
Median, IQR:
PM10 Og/m3): 33, 23

Tolbert et al. (2000a)
Atlanta
Period 1: 1/1/93-7/31/98
Mean, median, SD:
PM10 Og/m3):  30.1,28.0, 12.4

Period 2: 8/1/98-8/31/99
Mean, median, SD:
PM10 Og/m3):  29.1,27.6, 12.0
PM25Og/m3): 19.4, 17.5,9.35
CP Og/m3):  9.39, 8.95, 4.52 10-100 nm PM
counts
(count/cm3):  15,200,  10,900, 26,600
10-100 nm PM surface area (umVcm3):
62.5,43.4, 116
PM25 soluble metals Og/m3):  0.0327,
0.0226, 0.0306
PM2.5 Sulfates Og/m3):  5.59, 4.67, 3.6
PM2 5 Acidity Og/m3): 0.0181,0.0112,
0.0219
PM25 organic PM Og/m3):  6.30, 5.90, 3.16
PM2 5 elemental carbon Og/m3): 2.25, 1.88,
1.74
Total cardivascular hospital admissions in persons 65 and
older (ICD 9 codes390-429) in relation to PM10. Data
were analyzed to examine effect modification by
concurrent or preexisting cardiac and/or respiratory
conditions, age, race, and sex. No co-pollutants included.

Preliminary analysis of daily emergency department (ED)
visits for dysrhythmias, DYS,  (ICD 9 code 427) and all
cardiovascular diseases, CVD, (codes 402, 410-414, 427,
428, 433-437, 440, 444, 451-453) for persons aged 16 and
older in the period before (Period 1) and during (Period 2)
the Atlanta superstation study. ED data analyzed here
from just 18 of 33 participating hospitals; numbers of
participating hospitals increased during period 1.  Mean
daily ED visits for dysrhythmias and all CVD in period 1
were 6.5 and 28.4, respectively.  Mean daily ED visits for
dysrhythmias and all CVD in period 2 were 11.2 and 45.1,
respectively. Covariates: NO2, O3, SO2, CO temperature,
dewpoint, and, in period 2 only, VOCs. PM measured by
both TEOM and Federal Reference Method; unclear which
used in analyses. For epidemiologic analyses, the two time
periods were analyzed separately. Poisson regression
analyses were conducted with  cubic splines for time,
temperature and dewpoint. Day of week and hospital
entry/exit indicators also included. Pollutants were treated
a-priori as three-day moving averages of lags 0, 1, and 2.
Only single-pollutant results reported.
                                                                                                            Evidence seen for increased CVD effects among
                                                                                                            persons with concurrent respiratory infections or
                                                                                                            with previous admissions for conduction
                                                                                                            disorders.
In period 1, significant negative association
(p=0.02) observed between CVD and 3-day
average PM10. There was ca. 2% drop in CVD
per 10 ,ug/m3 increase in PM10. CVD was
positively associated with NO2 (p=0.11) and
negatively associated with SO2 (p=0.10). No
association observed between dysrhythmias and
PM10 in period 1. However, dysrhythmias were
positively associated with NO2 (p=0.06). In
period 2, i.e., the first year of operation of the
superstation, no  associations seen with PM10 or
PM25.  However, significant positive
associations observed between CVD and
elemental carbon (p=0.005) and organic matter
(p=0.02), as well as with CO (p=0.001).
For dysrhythmias,  significant positive
associations observed with elemental carbon
(p=0.004), CP (p=0.04), and CO (p=0.005).
These preliminary results should be interpreted
with caution given the incomplete and variable
nature of the databases analyzed.
Percent Excess CVD Risk (95% CI)
Effects computed for 50 ,ug/m3

PM10, 0-1 D. AVG.
CVD: 6.6 (4.9-8.3)

Percent Excess Risk (p-value):
Effects computed for 50 ,ug/m3 change
in PM10; 25 ^g/m3 for CP and PM2 5;
25,000 counts/cm3 for 10-100 nm
counts.

Period 1:
PM10, 0-2 d. avg.
CVD: -8.2(0.02)
DYS: 4.6 (0.58)

Period 2:
0-2 d. avg. in all cases
CVD % effect; DYS % effect:
PM10: 5.1 (-7.9, 19.9); 13.1 (-14.1,
50.0)
PM25: 6.1 (-3.1, 16.2); 6.1 (-12.6,
28.9)
CP: 17.6 (-4.6, 45.0); 53.2 (2.1,
129.6)
10-100 nm counts:  -11.0
(0.17); 3.0 (0.87)
 O

-------
to
o
o
to
                           TABLE 8B-1  (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                                HOSPITAL ADMISSIONS
           Reference citation. Location, Duration
           PM Index, Mean or Median, IQR
Study Description:  Health outcomes or codes,
Mean outcome rate, sample or population size,
ages.  Concentration measures or estimates.
Modeling methods: lags, smoothing, co-pollutants,
covariates, concentration-response
Results and Comments. Design Issues,
Uncertainties, Quantitative Outcomes
PM Index, Lag, Excess
Risk % (95% LCL, UCL),
Co-Pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           Canada

           Burnett etal. (1995)
           Ontario, Canada
           1983-1988

           Sulfate
           Mean: 4.37 ,ug/m3
           Median: 3.07,ug/m3
           95th percentile:  13 ^
           Burnett etal. (1997a)
           Canada's 10 largest cities
           1981-1994

           COH daily maximum
           Mean: 0.7 103 In feet
           Median: 0.6 103 In feet
           95th percentile:  1.5103lnfeet
168 Ontario hospitals.  Hospitalizations for coronary artery
disease, CAD (ICD9 codes 410,413), cardiac
dysrhythmias, DYS (code 427), heart failure, HF (code
428), and all three categories combined (total CVD).
Mean total CVD rate: 14.4/day.  1986 population of study
area: 8.7 million. All ages, <65, >=65. Both sexes, males,
females. Daily sulfates from nine monitoring stations.
Ozone from 22 stations. Log hospitalizations filtered with
19-day moving average prior to GEE analysis. Day of
week effects removed.  0-3 day lags examined.
Covariates: ozone, ozone2, temperature, temperature2.
Linear and quadratic sulfate terms included in model.
Daily hospitalizations for congestive heart failure (ICD9
code 427) for patients over 65 years at 134 hospitals.
Average hospitalizations: 39/day. 1986 population of
study area:  12.6 million. Regressions on air quality using
generalized estimating equations, controlling for long-term
trends, seasonality, day of week, and inter-hospital
differences.  Models fit monthly and pooled over months.
Log hospitalizations filtered with 19-day moving average
prior to GEE analysis. 0-3 day lags examined.  Covariates:
CO, SO2, NO2, O3, temperature, dewpoint temperature.
Sulfate lagged one day significantly assoc. with
total CVD admissions with and without ozone in
the model. Larger associations observed for
coronary artery disease and heart failure than for
cardiac dysrhythmias. Suggestion of larger
associations for males and the sub-population 65
years old and greater.  Little evidence for
seasonal differences in sulfate effects after
controlling for covariates.
COH significant in single-pollutant models with
and without weather covariates. Only InCO and
In NO2 significant in multi-pollutant models.
COH highly colinear with CO and NO2.
Suggests no particle effect independent of gases.
However, no gravimetric PM data were included.
Effects computed for 95th percentile
change in SO4

SO4, Id, no covariates:

Total CVD: 2.8(1.8,3.8)
CAD: 2.3(0.7,3.8)
DYS: 1.3 (-2.0, 4.6)
HF:  3.0(0.6,5.3)

Males: 3.4(1.8, 5.0)
Females: 2.0(0.2,3.7)

<65: 2.5(0.5,4.5)
>=65: 3.5(1.9,5.0)

SO4, Id, w. temp and O3:

Total CVD: 3.3 (1.7,4.8)

Effects computed for 95% change in
COH:

0 d lag:
  5.5% (2.5, 8.6)
0 d lag w/weather:
  4.7% (1.3, 8.2)
0 d lag w/CO, NO2, SO2, O3:
  -2.26 (-6.5, 2.2)
O

-------
 to
 o
 o
 to
                           TABLE 8B-1 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                             HOSPITAL ADMISSIONS
           Reference citation.  Location, Duration
           PM Index, Mean or Median, IQR
Study Description: Health outcomes or codes,
Mean outcome rate, sample or population size,
ages. Concentration measures or estimates.
Modeling methods: lags, smoothing, co-pollutants,
covariates, concentration-response
Results and Comments. Design Issues,
Uncertainties, Quantitative Outcomes
PM Index, Lag, Excess
Risk % (95% LCL, UCL),
Co-Pollutants
 OO
 td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           Canada (cont'd)

           Burnett etal. (1997b)
           Metro-Toronto, Canada
           1992-1994

           Pollutant: mean, median, IQR:
           COH (103 In ft):  0.8,0.8,0.6
           H+ (nmol/m3): 5, 1, 6
           SO4 (nmol/m3): 57, 33, 57
           PM10 0/g/m3):  28,25,22
           PM25  (,ug/m3): 17, 14, 15
           PM10.25 Cug/m3):  12,10,7
Daily unscheduled cardiovascular hospitalizations (ICD9
codes 410-414,427, 428) for all ages. Average hospital
admissions: 42.6/day.  Six cities of metro-Toronto
included Toronto, North York, East York, Etobicoke,
Scarborough, and York, with combined 1991 population
of 2.36 million. Used same stat model as in Burnett et al.,
1997c.  0- 4 day lags examined, as well as multi-day
averages. Covariates: O3, NO2, SO2, CO, temperature,
dewpoint temperature.
Relative risks > 1 for all pollutants in univariate
regressions including weather variables; all but
H+ and FP statistically significant. In
multivariate models, the gaseous pollutant
effects were generally more robust than were
particulate effects.  However, in contrast to
Burnett et al. (1997A), COH remained
significant in multivariate models. Of the
remaining particle metrics, CP was the most
robust to the inclusion of gaseous covariates.
Results do not support independent effects of FP,
SO4, or H+ when gases are controlled.
Percent excess risk (95% CI) per
50 Mg/m3 PM10, 25 ,ug/m3 PM25 and
PM10_2 5, and IQR for other indicators.

COH: 0-4 d.
 6.2(4.0, 8.4)
 5.9(2.8, 9.1) w. gases
H+: 2-4 d.
 2.4(0.4,4.5)
 0.5 (-1.6, 2.7) w. gases
S04: 2-4 d.
 1.7 (-0.4, 3.9)
 -1.6 (-4.4, 1.3) w. gases
PM10: l-4d.
 7.7(0.9, 14.8)
 -0.9 (-8.3, 7.1) w. gases
PM25 :2-4d.
 5.9(1.8, 10.2)
 -1.1 (-7.8, 6.0) w. gases
PM10.25:0-4d.
 13.5(5.5,22.0)
 8.1 (-1.3, 18.3) w. gases
O

-------
 to
 o
 o
 to
                            TABLE 8B-1  (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                               HOSPITAL ADMISSIONS
           Reference citation. Location, Duration
           PM Index, Mean or Median, IQR
Study Description: Health outcomes or codes,
Mean outcome rate, sample or population size,
ages.  Concentration measures or estimates.
Modeling methods: lags, smoothing, co-pollutants,
covariates, concentration-response
Results and Comments.  Design Issues,
Uncertainties, Quantitative Outcomes
PM Index, Lag, Excess
Risk % (95% LCL, UCL),
Co-Pollutants
 OO
 td
 to
 fe
 H
 6
 o

 o
 H
O
 O
 H
 W
 O
           Canada (cont'd)

           Burnett etal. (1999)
           Metro-Toronto, Canada
           1980-1994

           Pollutant: mean, median, IQR:
           FPe!t 0/g/m3):  18,16,10
           CPe!t Cug/m3): 12, 10, 8
           PM10 ea (Mg/m3):  30,27, 15
Daily hospitalizations for dysrhythmias, DYS (ICD9 code
427; mean 5/day); heart failure, HF (428; 9/d); ischemic
heart disease, IHD (410-414; 24/d); cerebral vascular
disease, CVD (430-438; 10/d); and diseases of the
peripheral circulation, DPC (440-459; 5/d) analyzed
separately in relation to environmental covariates. Same
geographic area as in Burnett et al., 1997b. Three size-
classified PM metrics were estimated, not measured, based
on a regression on TSP, SO4, and COH in a subset of every
6th-day data. Generalized additive models used and non-
parametric LOESS prefilter applied to both pollution and
hospitalization data.  Day of week controls.  Tested 1-3
day averages of air pollution ending on lags 0-2.
Covariates: O3, NO2, SO2, CO, temperature, dewpoint
temperature, relative humidity.
In univariate regressions, all three PM metrics
were associated with increases in cardiac
outcome (DVS, HF, IHD).  No associations with
vascular outcomes, except for CPest  with DPC.
In multi-pollutant models, PM effects estimates
reduced by variable amounts (often >50%) for
specific endpoints and no statistically significant
(at p<0.05) PM associations seen with any
cardiac or circulatory outcome (results not
shown). Use of estimated PM metrics limits
interpretation of pollutant-specific results.
However, results suggest that linear combination
of TSP, SO4, and COH does not have a strong
independent association with cardiovascular
admissions when full range of gaseous pollutants
also modeled.
Single pollutant models:
Percent excess risk (95% CI) per
50 ,ug/m3 PM10; 25 ,ug/m3 PM25; and
25 Mg/m3 PM10.2.5.

All cardiac HA (lags 2-5 d):
PM25 1-poll = 8.1(2.45, 14.1)
PM25 w/4 gases = -1.6 (-10.4, 8.2);
w/CO = 4.60 (-3.39, 13.26)
PM10 1-poll = 12.07 (1.43, 23.81)
w/4 gases = -1.40 (-12.53, 11.16)
w/CO= 10.93 (0.11,22.92)
PM10.25 1-poll = 20.46 (8.24, 34.06)
w/4 gases = 12.14 (-1.89, 28.2);
w/CO=19.85(7.19, 34.0)
DYS:
FPest(Od): 6.1 (1.9, 10.4)
CPest(Od): 5.2 (-0.21, 1.08)
PM10 est: (0 d):  8.41 (2.89, 14.2)
HF:
FPe!t(0-2d): 6.59(2.50, 10.8)
CPest (0-2 d):  7.9 (2.28, 13)
PM10 est (0-2 d): 9.7(4.2, 15.5)
IHD:
FPest(0-2d): 8.1 (5.4, 10.8)
CPest(Od): 3.7(1.3,6.3)
PM10est(0-ld): 8.4(5.3, 11.5)
O

-------
                            TABLE 8B-1 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
 to
 o
 o
 to
           Reference citation. Location, Duration
           PM Index, Mean or Median, IQR
Study Description:  Health outcomes or codes,
Mean outcome rate, sample or population size,
ages.  Concentration measures or estimates.
Modeling methods: lags, smoothing, co-pollutants,
covariates, concentration-response
Results and Comments. Design Issues,
Uncertainties, Quantitative Outcomes
PM Index, Lag, Excess
Risk % (95% LCL, UCL),
Co-Pollutants
 OO
 td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           Canada (cont'd)

           Stieb et al. (2000)
           Saint John, Canada
           7/1/92-3/31/96
           mean and S.D.:
           PM10 (,ug/m3):  14.0,9.0
           PM25(,ug/m3): 8.5, 5.9
           HOSPITAL ADMISSIONS

           H+(nmol/m3): 25.7,36.8
           Sulfate (nmol/m3): 31.1,29.7
           COHmean(103lnft): 0.2,0.2
           COHmax(103lnft): 0.6,0.5
Study of daily emergency department (ED) visits for
angina/myocardial infarction (mean 1.8/day), congestive
heart failure (1.0/day), dysrhythmia/conduction
disturbance (0.8/day), and all cardiac conditions (3.5/day)
for persons of all ages. Covariates included CO, H2S,
NO2, O3, SO2, total reduced sulfur (TRS), a large number
of weather variables, and 12 molds and pollens.  Stats:
generalized additive models with LOESS prefiltering of
both ED and pollutant variables, with variable window
lengths. Also controlled for day of week and LOESS-
smoothed functions of weather.  Single-day, and five day
average, pollution lags tested out to lag 10. The strongest
lag, either positive or negative, was chosen for final
models. Both single  and multi-pollutant models reported.
Full-year and May-Sep models reported.
In single-pollutant models, significant positive
associations observed between all cardiac ED
visits and PM10, PM25, H2S, O3, and SO2.
Significant negative associations observed with
H+, sulfate, and COH max. PM results were
similar when data were restricted to May-Sep.  In
multi-pollutant models, no PM metrics were
significantly associated with all cardiac ED visits
in full year analyses, whereas both O3 and SO2
were. In the May-Sep subset,  significant
negative association found for sulfate. No
quantitative results presented for non-significant
variables in these multi-pollutant regressions. In
cause-specific, single-pollutant models, PM
tended to be positively associated with
dysrhythmia/conductive disturbances but
negatively associated with congestive heart
failure (no quantitative results presented).  The
objective decision rule used for selecting lags
reduced the risk of data mining; however, the
biological  plausibility of lag effects beyond 3-5
days is open to question.  Rich co-pollutant data
base.  Results imply no effects of PM
independent of co-pollutants.
Percent Excess Risk (p-value)
computed for 50 ,ug/m3 PM10, 25
f/g/m3 PM2 5 and  mean levels of
sulfate and COH.

Full year results for all cardiac
conditions, single pollutant models:

PM10: 3d.
29.3  (P=0.003)

PM25: 3d.
14.4(P=0.055)

H+: 4-9 d. avg.
 -1.8(0.010)
Sulfate: 4d.
 -6.0(0.001)
COH max: 7d.
 -5.4(0.027)

Full year results for all cardiac
conditions, multi-pollutant models:

No significant PM associations.
 O

-------
                           TABLE 8B-1 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                               HOSPITAL ADMISSIONS
to
o
o
to
           Reference citation. Location, Duration
           PM Index, Mean or Median, IQR
                                          Study Description:  Health outcomes or codes,
                                          Mean outcome rate, sample or population size,
                                          ages. Concentration measures or estimates.
                                          Modeling methods: lags, smoothing, co-pollutants,
                                          covariates, concentration-response
Results and Comments.  Design Issues,
Uncertainties, Quantitative Outcomes
PM Index, Lag, Excess
Risk % (95% LCL, UCL),
Co-Pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
Europe

Le Tertre et al. (2002)
Eight-City - APHEA 2
Study mean (SD) PM10 ,ug/m3
Barcelona- 1/94-12/96
  55.7(18.4)
Birmingham - 3/92-12/94
  24.8(13.1)
London - 1/92-12/94
  28.4(12.3)
Milan - No PM10
Netherlands - 1/92-9/95
  39.5(19.9)
Paris - 1/92-9/96
  PM13-22.7 (10.8)
Rome - No PM10
Stockholm - 3/94-12/96
  15.5(7.2)
           Atkinson etal. (1999a)
           Greater London, England
           1992-1994

           Pollutant: mean, median, 90-10 percentile
           range:
           PM10 (Mg/m3): 28.5, 24.8, 30.7
           Black Smoke (Mg/m3): 12.7,10.8,16.1
                                                     Examined the association between measures of PM to
                                                     include PM10 and hospital admissions for cardiac causes in
                                                     eight European cities with a combined population of
                                                     38 million.  Examined age factors and ischemic heart
                                                     disease and studies also stratified by age using
                                                     autoregressive Poisson models controlled for long-term
                                                     trends, season, influenza, epidemics, and meteorology, as
                                                     well as confounding by other pollutants. In a second
                                                     regression examined, pooled city-specific results for
                                                     sources of heterogeneity.
                                          Daily emergency hospital admissions for total
                                          cardiovascular diseases, CVD (ICD9 codes 390-459), and
                                          ischemic heart disease, IHD (ICD9 410-414), for all ages,
                                          for persons less than 65, and for persons 65 and older.
                                          Mean daily admissions for CVD: 172.5 all ages, 54.5 <65,
                                          117.8 >65; for IHD: 24.5 <65, 37.6 >65. Covariates: NO2,
                                          O3, SO2, CO, temperature, relative humidity. Poisson
                                          regression using APHEA methodology; sine and cosine
                                          functions for seasonal control; day of week dummy
                                          variables.  Lags of 0-3, as well as corresponding multi-day
                                          averages ending on lag 0, were considered.
Pooled results were reported for the cardiac
admissions results in table format.  City-specific
and pooled results were depicted in figures only.
Found a significant effect of PM10 and black
smoke on admissions for cardiac causes (all
ages) and cardiac causes and ischemic heart
disease for people over 65 years with the impact
of PM10 per unit of pollution being half that
found in the United States. PM10 did not seem to
be confounded by O3 or SO2. The effect was
reduced when CO was incorporated in the
regression model and eliminated when
controlling for NO2.  There was little  evidence of
an impact of particles on hospital admissions for
ischemic heart disease for people below 65 years
or stroke for people over 65 years.  The authors
state results were consistent with a role for traffic
exhaust/diesel in Europe.

In single-pollutant models, both PM metrics
showed positive associations with both CVD and
IHD admissions across age groups. In Two-
pollutant models, the BS effect, but not the PM10
effect, was robust. No quantitative results
provided for two-pollutant models. Study does
not support a PM10 effect independent of co-
pollutants.
For a 10 ,ug/m3 increase in PM10

Cardiac admissions/all ages
0.5% (0.2, 0.8)

Cardiac admissions/over 65 years
0.7% (0.4, 1.0)

Ischemic heart disease/over 65 years
0.8% (0.3, 1.2)

For cardiac admissions for people
over 65 years: All the city-specific
estimates were positive with London,
Milan, and Stockholm significant at
1he 5% level.
Effects computed for 50 ,ug/m3 PM10
and 25 ^g/m3 BS
PM10 0 d.
All ages:
CVD: 3.2(0.9, 5.5)
0-64 yr:
CVD: 5.6(2.0,9.4)
IHD: 6.8(1.3, 12.7)
65+ yr:
CVD: 2.5 (-0.2, 5.3)
IHD: 5.0(0.8,9.3)

Black Smoke 0 d.
All ages:
CVD: 2.95(1.00,4.94)
0-64 yr:
CVD: 3.12(0.05,6.29)
IHD: 2.78 (-1.88, 7.63)
65+ yr:
CVD: 4.24(1.89,6.64)
IHD (lag 3): 4.57 (0.86, 8.42)

-------
                           TABLE 8B-1 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                               HOSPITAL ADMISSIONS
to
o
o
to
           Reference citation. Location, Duration
           PM Index, Mean or Median, IQR
Study Description: Health outcomes or codes,
Mean outcome rate, sample or population size,
ages.  Concentration measures or estimates.
Modeling methods: lags, smoothing, co-pollutants,
covariates, concentration-response
Results and Comments.  Design Issues,
Uncertainties, Quantitative Outcomes
PM Index, Lag, Excess
Risk % (95% LCL, UCL),
Co-Pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           Europe (cont'd)

           Prescottetal. (1998)
           Edinburgh, Scotland
           1981-1995 (BS and SO2)
           1992-1995 (PM10, NO2, O3, CO)
           Means for long and short series:
           BS: 12.3, 8.7
           PM10:  NA, 20.7
           Wordleyetal. (1997)
           Birmingham, UK
           4/1/92-3/31/94
           mean, min, max:
           PM10 (,ug/m3): 26,3, 131
           Diaz etal. (1999)
           Madrid, Spain
           1994-1996

           TSP by beta attenuation
           Summary statistics not given.
Daily emergency hospital admissions for cardiovascular
disease (ICD9 codes 410-414, 426-429, 434-440) for
persons less than 65 years and for persons 65 or older.
Separate analyses presented for long (1981-1995) and
short (1992-1995) series. Mean hospital admissions
for long and short series: <65, 3.5, 3.4; 65+, 8.0, 8.7.
Covariates: SO2, NO2, O3, CO, wind speed, temperature,
rainfall.  PM10 measured by TEOM. Stats: Poisson
log-linear regression; trend and seasons controlled by
monthly dummy variables over entire series; day of week
dummy variables; min daily temperature modeled using
octile dummies. Pollutants expressed as cumulative lag 1-
3 day moving avg.
Daily hospital admissions for acute ischemic heart disease
(ICD9 codes 410-429) for all ages. Mean hospitalizations:
25.6/day. Covariates: temperature and relative humidity.
Stats: Linear regression with day of week and monthly
dummy variables, linear trend term. Lags of 0-3
considered, as well as the mean of lags 0-2.
Daily emergency hospital admissions for all cardiovascular
causes (ICD9 codes 390-459) for the Gregorio Maranon
University Teaching Hospital. Mean admissions: 9.8/day.
Covariates: SO2, NO2, O3, temperature, pressure, relative
humidity, excess sunlight.  Stats: Box-Jenkins time-series
methods used to remove autocorrelations, followed by
cross-correlation analysis; sine and cosine terms for
seasonality; details unclear.
In long series, neither BS nor NO2 were
associated with CVD admissions in either age
group.  In the short series, only 3-day moving
average PM10 was positively and significantly
associated with CVD admissions in single-
pollutant models, and only for persons 65 or
older. BS, SO2, and CO also snowed positive
associations in this subset, but were not
significant at the 0.05 level. The PM10 effect
remained largely unchanged when all other
pollutants were added to the model, however
quantitative results were not given.  Results
appear to show an effect of PM10 independent of
co-pollutants.
No statistically significant effects observed for
PM10 on ischemic heart disease admissions for
any lag. Note that PM10 was associated with
respiratory admissions and with cardiovascular
mortality in the same study (results not shown
here).
No significant effects of TSP on CVD reported.
Percent Excess Risk (95% CI):
Effects computed for 50 //g/m3 change
in PM10 and 25 ,ug/m3 change in BS.

Long series:
BS, 1-3d. avg.
<65: -0.5 (-5.4, 4.6)
65+: -0.5 (-3.8, 2.9)

Short series:
BS, 1-3 d. avg.
<65: -9.5 (-24.6, 8.0)
65+: 5.8 (-4.9, 17.8)

PM10,  1-3d. avg.
<65: 2.0 (-12.5, 19.0)
65+: 12.4(4.6,20.9)

% change (95% CI) per
50 ,ug/m3 change PM10
IHD admissions:
PM10 0-dlag:
  1.4% (-4.4, 7.2)
PM10  1-dlag:
  -1.3% (-7.1, 4.4)

No quantitative results presented for
PM.
O

-------
                           TABLE 8B-1 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND CARDIOVASCULAR
                                                                               HOSPITAL ADMISSIONS
to
o
o
to
           Reference citation. Location, Duration
           PM Index, Mean or Median, IQR
Study Description: Health outcomes or codes,
Mean outcome rate, sample or population size,
ages.  Concentration measures or estimates.
Modeling methods: lags, smoothing, co-pollutants,
covariates, concentration-response
Results and Comments.  Design Issues,
Uncertainties, Quantitative Outcomes
PM Index, Lag, Excess
Risk % (95% LCL, UCL),
Co-Pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Australia

          Morgan etal. (1998)
          Sydney, Australia
          1990-1994

          mean, median, IQR, 90-10 percentile range:
          Daily avg. bscat/104m: 0.32, 0.26, 0.23,
          0.48
          Daily max l-hrbscat/104m: 0.76, 0.57, 60,
          1.23
          Asia

          Wong etal. (1999)
          Hong Kong
          1994-1995
          median, IQR for PM10 (|/g/m3): 45.0, 34.8
Daily hospital admissions for heart disease (ICD9 codes
410, 413, 427, 428) for all ages, and separately for persons
less than 65 and persons 65 or greater. Mean daily
admissions: all ages, 47.2; <65, 15.4; 65+, 31.8.  PM
measured by nephelometry (i.e., light scattering), which is
closely associated with PM25.  Authors give conversion for
Sydney as PM2 5 =30 x bscat.  Covariates: O3, NO2,
temperature, dewpoint temperature. Stats:  Poisson
regression; trend and seasons controlled with linear time
trend and monthly dummies; temperature and dewpoint
controlled with dummies for eight levels of each variable;
day of week and holiday dummies. Single and cumulative
lags from 0-2 considered. Both single and multi-pollutant
models were examined.
Daily emergency hospital admissions for cardiovascular
diseases, CVD (ICD9 codes 410-417, 420-438, 440-444),
heart failure, HF (ICD9 428), and ischemic heart disease,
IHD (ICD9 410-414) among all ages and in the age
categories  5-64, and 65+. Median daily CVD admissions
for all ages: 101. Covariates: NO2, O3, SO2, temperature,
relative humidity. PM10 measured by TEOM.  Stats:
Poisson regression using the APHEA protocol; linear and
quadratic control of trends; sine and cosine control for
seasonality; holiday and day of week dummies;
autoregressive terms. Single and cumulative lags from 0-5
days considered.
In single-pollutant models, NO2 was strongly
associated with heart disease admissions in all
age groups. PM was more weakly, but still
significantly associated with admissions for all
ages and for persons 65+. The NO2 association
in the 65+ age group was unchanged in
the multi-pollutant model, whereas the PM effect
disappeared when NO2 and O3 were added to the
model.. These results suggest that PM is not
robustly associated with heart disease admissions
when NO2 is included, similar to the sensitivity
of PM to CO in other studies.
In single-pollutant models, PM10, NO2, SO2, and
O3 all significantly associated with CVD
admissions for all ages and for those 65+. No
multi-pollutant risk coefficients were presented;
however, the PM10 effect was larger when O3 was
elevated (i.e., above median). A much larger
PM10 effect was observed for HF than for CVD
or IHD.  These results confirm the presence of
PM10 associations with cardiovascular
admissions in single-pollutant models, but do not
address the independent role of PM10.
Percent Excess Risk (95% CI):
Effects computed for 25 //g/m3 PM2
(converted from bscat).

24-hr avg. PM25Od.
  <65:  1.8 (-2.9, 6.7)
  65+: 4.9(1.6,8.4)
  All:   3.9(1.1,6.8)

24-hr PM2 5, 0 d w. NO2 and O3.
  65+: 0.12 (-1.3, 1.6)

l-hrPM25, Od.
  <65: (X19(-1.6, 2.0)
  65+:  1.8(0.5,3.2)
  All:   1.3(0.3,2.3)
Percent Excess Risk (95% CI):
Effects computed for 50 ,ug/m3 change
in PM10.

PM10, 0-2 d. avg.

CVD:
  5-64:  2.5 (-1.5, 6.7)
  65+:    4.1(1.3,6.9)
  All:    3.0(0.8,5.4)

HF (PM10, 0-3 d ave.):
  All: 26.4(17.1,36.4)

IHD (PM10, 0-3 d ave.):
  All: 3.5 (-0.5, 7.7)	
O

-------
            Appendix 8B.2.  PM-Respiratory Hospitalization Studies
April 2002                         8B-17      DRAFT-DO NOT QUOTE OR CITE

-------
 to
 o
 o
 to
                 TABLE 8B-2. ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation.
 Location, Duration
 PM Index/Concentrations
                                                        Study Description:
                                                  Results and Comments
                                                PM Index, Lag, Excess Risk %,
                                             (95% CI = LCI, UCL) Co-Pollutants
 oo
 td
 oo
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          United States

          Samet et al, (2000a,b)
          Study Period.: 84-95
          14 U.S. Cities: Birmingham,
          Boulder, Canton, Chicago, Col.
          Springs, Detroit, Minn./St. Paul,
          Nashville, New Haven, Pittsburgh,
          Provo/Orem, Seattle, Spokane,
          Youngstown.  Mean pop. aged
          65+yr per city =143,000
          PM10 mean = 32.9 /j,g/m3
          PM10IQR = NR
 Zanobetti et al. (2000b)
 10 U.S. Cities
                                    Hospital admissions for adults 65+ yrs. for
                                    CVD (mean=22.1/day/city), COPD
                                    (mean=2.0/day/city), and Pneumonia
                                    (mean=5.6/day/city) related to PM10, SO2,
                                    O3, NO2, and CO. City-specific Poisson
                                    models used with adjustment for season,
                                    mean temperature (T) and relative humidity
                                    (RH) (but not their interaction), as well as
                                    barometric pressure (BP) using LOESS
                                    smoothers (span usually 0.5).  Indicators for
                                    day-of-week and autoregressive terms also
                                    included.
Derived from the Samet et al. (2000a,b)
study, but for a subset of 10 cities. Daily
hospital admissions for total cardiovascular
and respiratory disease in persons aged >65
yr.  Covariates: SO2, O3, CO, temperature,
relative humidity, barometric pressure. In
first stage, performed single-pollutant
generalized additive robust Poisson
regression with seasonal, weather, and day
of week controls. Repeated analysis for
days with PM10 less than 50 /-ig/m3 to test
for threshold. Lags of 0-5 d considered, as
well as the quadratic function of lags 0-5.
Individual cities analyzed first. The 10 risk
estimates were then analyzed in several
second stage analyses:  combining risks
across cities using inverse variance weights,
and regressing risk estimates on potential
effect-modifiers and pollutant confounders.
PM10 positively associated with all three
hospital admission categories, but city
specific results ranged widely, with less
variation for outcomes with higher daily
counts.  PM10 effect estimates not found to
vary with co-pollutant correlation,
indicating that results appear quite stable
when controlling for confounding by
gaseous pollutants.  Analyses found little
evidence that key socioeconomic factors
such as poverty or race are modifiers, but
it is noted that baseline risks may differ,
yielding differing impacts for a given RR.

Same basic pattern of results as in Samet
et al. (2000a,b). For distributed lag
analysis, lag 0 had largest effect, lags 1
and 2 smaller effects, and none at larger
lags. City-specific slopes were
independent of percent poverty and
percent non-white. Effect size increase
when data were restricted to days with
PM10 less than 50^g/m3.  No multi-
pollutant models reported; however, no
evidence of effect modification by co-
pollutants in second stage analysis.
Suggests association between PM10 and
total respiratory hospital admissions
among the elderly.
                                                                                  COPD HA's for Adults 65+ yrs.
                                                                                  LagOER = 7.4%(CI:5.1,9.8)
                                                                                  Lag 1 ER = 7.5% (CI: 5.3, 9.8)
                                                                                  2 day mean (lagOJagl) ER = 10.3%
                                                                                           (CI: 7.7, 13)
                                                                                  Pneumonia HA's for Adults 65+ yrs.
                                                                                  Lag 0 ER =8.1% (CI: 6.5, 9.7)
                                                                                  LaglER = 6.7%(CI:5.3, 8.2)
                                                                                  2 day mean (lagO, lagl) = 10.3%
                                                                                           (CI: 8.5, 12.1)
Percent excess respiratory risk (95% CI) per
50 Mg/rn3 PM10 increase:
COPD (0-1 d lag) = 10.6 (7.9, 13.4)
COPD (unconstrained dist. lag) = 13.4 (9.4,
17.4)
Pneumonia (0-1 d lag) = 8.1 (6.5, 9.7)
Pneumonia (unconstrained dist. lag) =10.1
(7.7, 12.6)
O

-------
to
o
o
to
            TABLE 8B-2 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                       Study Description:
         Results and Comments
                                                                                        PM Index, Lag, Excess Risk %,
                                                                                      (95% CI = LCI, UCL) Co-Pollutants
oo
td
 fe
 H
 6
 o

 o
 H
O
 O
 H
 W
 O
          United States (cont'd)

          Jamason et al. (1997)
          New York City, NY (82 - 92)
          Population = NR
          PM10 mean = 38.6 ,ug/rn3
          Chen et al. (2000)
          Reno-Sparks, NV (90 - 94)
          Population = 307,000
          B-Gauge PM10 mean=36.5 /-
          PM10 IQR = 18.3-44.9 Mg/m
          PM10 maximum = 201.3 /-ig/
          Gwynn et al. (2000)
          Buffalo, NY (5/88-10/90)
          PM10 mn./max. = 24.1/90.8 Mg/
          PM10IQR = 14.8-29.2 Mg/m3
          SO4= mn./max. = 2.4/3.9 Mg/m3
          SO4=IQR = 23.5-7.5 Mg/m3
          H+ mn/max = 36.4/382 nmol/m3
          H+ IQR = 15.7-42.2 nmol/m3
          CoH mn/max = 0.2/0.9 10~3 ft.
          CoH IQR =0.1-0.3
          Gwynn etal. (2001)
          New York City, NY
          1988,89,90
          PM10 37.4 Mg/ni3 mean
Weather/asthma relationships examined
using a synoptic climatological multivariate
methodology. Procedure relates
homogenous air masses to daily counts of
overnight asthma hospital admission.
                                   Log of COPD (mean=1.72/day) and
                                   gastroenteritis (control) admissions from 3
                                   hospitals analyzed using GAM regression,
                                   adjusting for effects of day-of-week,
                                   seasons, Weather effects (T, WS), and long-
                                   wave effects. No co-pollutants considered.
                                   Air pollutant-health effect associations
                                   with total, respiratory, and circulatory
                                   hospital admissions and mortality examined
                                   using Poisson methods controlling for
                                   weather, seasonality, long-wave effects, day
                                   of week, holidays,
                                   Respiratory hospital admissions, race
                                   specific for PM10, H+, O3, SO4=. Regression
                                   model used to model daily variation in
                                   respiratory hospital admissions, day-week,
                                   seasonal, and weather aspects addressed in
                                   modeling.
Air pollution reported to have little role in
asthma variations during fall and winter.
During spring and summer, however, the
high risk categories are associated with
high concentration of various pollutants
(i.e., PM10, SO2,NO2, O3).

PM10 positively associated with COPD
admissions, but no association with
gastroenteritis (GE) diseases, indicating
biologically plausible specificity of the
PM10-health effects association.
Association remained even after excluding
days with PM10 above 150 ^g/m3.

Strongest associations found between
SO4= and respiratory hospital admissions,
while secondary aerosol H+ and SO4=
demonstrated the most coherent
associations across both respiratory
hospital admissions and mortality.
Addition of gaseous pollutants to the
model had minimal effects on the PM RR
estimates. CoH weakness in associations
may reflect higher toxicity by acidic sulfur
containing secondary particles versus
carbonaceous primary particles.

Greatest difference between the white and
non-white subgroups was observed for O3.
However, within race analyses by
insurable coverage suggested that most of
the higher effects of air pollution found
for minorities were related to
socio-economic studies.
                                                                                                                    NR
                                                                                 COPD All age Admissions
                                                                                 50 Mg/m3 IQR PM10 (single pollutant):
                                                                                 ER = 9.4% (CI: 2.2, 17.1)
                                                                                 Respiratory Hospital Admissions(all ages) PM
                                                                                 Index (using standardized cone, increment)
                                                                                 -Single Pollutant Models
                                                                                 H+ = 75nmoles/m3;COH = 0.5 units/1 000ft
                                                                                 PM10(lag 0) ER = 11% (CI: 4.0, 18)
                                                                                 S04=(lag 0) ER = 8.2% (CI:  4.1, 12.4)
                                                                                 H+(lag 0) ER = 6% (CI: 2.8, 9.3)
                                                                                 CoH(lagO) ER = 3% (CI: - 1.2, 7.4)
                                                                                 PM10 (max-min) increment
                                                                                 1 day lag
                                                                                 white 1.027 (0.971-1.074)
                                                                                 non-white (1.027 (0.988-1.069)
O

-------
to
o
o
to
           TABLE 8B-2 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                      Study Description:
                                                Results and Comments
                                              PM Index, Lag, Excess Risk %,
                                            (95% CI = LCI, UCL) Co-Pollutants
         United States (cont'd)

         Jacobs etal. (1997)
         Butte County, CA (83 - 92)
         Population = 182,000
         PM10 mean = 34.3 ,ug/rn3
         PM10 mm/max = 6.6 / 636 /-ig/m3
         CoH mean = 2.36 per 1000 lin. ft.
         CoH mm/max = 0/16.5
                                   Association between daily asthma HA's
                                   (mean = 0.65/day) and rice burning using
                                   Poisson model with a linear term for
                                   temperature, and indicator variables for
                                   season and yearly population.
                                   Co-pollutants were O3 and CO. PM10
                                   estimated for 5 of every 6 days from CoH.
                                        Increases in rice straw bum acreage found
                                        to correlate with asthma HA's over time.
                                        All air quality parameters gave small
                                        positive elevations in RR. PM10 showed
                                        the largest increase in admission risk.
                                       Asthma HA's (all ages)
                                       For an increase of 50 ,ug/ni3 PM10:
                                       ER = 6.11% (not statistically significant)
oo
td
to
o
fe
H
6
o
o
H
O
O
H
W
O
 Linn et al. (2000)
 Los Angeles, CA (92-95)
 Population = NR
 PM10 mean = 45.5 /^g/m3
 PM10 Min/Max = 5/132 Mg/
         Moolgavkar et al. (1997)
         Minneapolis-St. Paul 86-91
         Populations NR
         Birmingham, AL '86-'91
         Population. = NR
         PM10 mean = 34 ,ug/m3 (M-SP)
         PM10IQR =22-41 Mg/m3 (M-SP)
         PM10 mean =43.4 ^g/m3(Birm)
         PM10 IQR =26-56 Mg/m3(Birm)
Pulmonary hospital admissions (HA's)
(mean=74/day) related to CO, NO2, PM10,
and O3 in Los Angeles using Poisson model
with long-wave, day of week, holidays, and
weather controls.
                                   Investigated associations between air
                                   pollution (PM10, SO2, NO2 O3, and CO) and
                                   hospital admissions for COPD
                                   (mean/day=2.9 in M-SP; 2.3 in Birm) and
                                   pneumonia (mean=7.6 in M-SP; 6.0 in
                                   Birm) among older adults (>64 yrs.).
                                   Poisson GAM's used, controlling for day-
                                   of-week, season, LOESS of temperature
                                   (but neither RH effects nor T-RH
                                   interaction considered).
PM10 positively associated with
pulmonary admissions year-round,
especially in winter. No association with
cerebro-vascular or abdominal control
diseases. However, use of linear
temperature, and with no RH interaction,
may have biased effect estimates
downwards for pollutants here most
linearly related to temperature  (i.e., O3
and PM10).

In the M-SP area, PM10 significantly and
positively associated with total daily
COPD and pneumonia admissions among
elderly, even after simultaneous inclusion
of O3. When four pollutants included in
the model (PM10, SO2, O3, NO2), all
pollutants remained positively associated.
In Birm., neither PM10 nor O3 showed
consistent associations across lags.  The
lower power (fewer counts) and lack of T-
RH interaction weather modeling in this
Southern city vs. M-SP may have
contributed to  the differences seen
between cities.
                                                                                                                          Pulmonary HA's (>29 yrs.)
                                                                                                                          PM10 = 50 Mg/m3
                                                                                                                          (Lag 0)ER = 3.3% (CI: 1.7, 5)
                                                                               COPD + Pneumonia Admissions (>64yrs.)

                                                                               In M-SP, For PM10 = 50 /j,g/m3 (max Ig)
                                                                               ER(lgl) = 8.7%(CI:4.6, 13)
                                                                               With O3 included simultaneously:
                                                                               ER(lgl)= 6.9%(95 CI: 2.7, 11.3)

                                                                               In Birm, For PM10=50 /-ig/m3 (max Ig.)
                                                                               ER(lgO)=1.5%(CI: -1.5,4.6)
                                                                               With O3 included simultaneously:
                                                                               ER(lgO) = 3.2% (CI: -0.7, 7.2)
O

-------
to
o
o
to
            TABLE 8B-2 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                       Study Description:
                                                 Results and Comments
                                               PM Index, Lag, Excess Risk %,
                                             (95% CI = LCI, UCL) Co-Pollutants
oo
td
to
fe
H
6
o
o
H
O
O
H
W
O
          United States (cont'd)

          Nauenberg and Basu (1999)
          Los Angeles (9 1-94)
          Wet Season =11/1-3/1
          Dry Season =5/1 -8/1 5
          Population .= 2.36 Million
         PM10SE= 17.23,ug/m3
 Schwartz et al. (1996b)
 Cleveland (Cayahoga County), Ohio
 (88 - 90)
 PM10 mean = 43 /j.g/m3
 PM10IQR = 26-56 ,ug/m3
         Zanobetti, et al. (2000a)
         Study Period: 86 - 94
         Chicago (Cook Count), IL
         Population = 633,000 aged 65+
         PM10 mean = 33.6 Mg/m3
         PM10 range = 2.2,  157.3 Mg/m3
The effect of insurance status on the
association between asthma-related hospital
admissions and exposure to PM10 and O3
analyzed, using regression techniques with
same day and 8-day weighted moving
average levels, after removing trends using
Fourier series. Compared results during wet
season for all asthma HA's (mean = 8.7/d),
for the uninsured (mean=0.77/d), for
MediCal (poor) patients (mean = 4.36/d),
and for those with other private health or
government insurance (mean = 3.62/d).


Review paper including an example drawn
from respiratory hospital admissions of
adults aged 65 yr and older (mean = 22/day)
in Cleveland, OH. Categorical variables for
weather and sinusoidal terms for filtering
season employed.

Analyzed HA's for older adults (65 + yr)
for COPD (mean = 7.8/d), pneumonia
(mean = 25.5/d), and CVD, using Poisson
regression controlling for temperature, dew
point, barometric pressure, day of week,
long wave cycles and autocorrelation, to
evaluate whether previous admission or
secondary diagnosis for associated
conditions increased risk from air pollution.
Effect modification by race, age, and sex
also evaluated.
                                                                            No associations found between asthma
                                                                            admissions and O3. No O3 or PM10
                                                                            associations found in dry season. PM10
                                                                            averaged over eight days associated with
                                                                            increase in asthma admissions, with even
                                                                            stronger increase among MediCal asthma
                                                                            admissions in wet season. The authors
                                                                            conclude that low income is useful
                                                                            predictor of increased asthma
                                                                            exacerbations associated with air
                                                                            pollution. Non-respiratory HA's showed
                                                                            no such association with PM10.
Hospital admissions for respiratory illness
of persons aged 65 yr and over in
Cleveland strongly associated with PM10
and O3, and marginally associated with
SO2 after control for season, weather, and
day of the week effects.

Air pollution- associated CVD HA's were
nearly doubled for those with concurrent
respiratory infections (RI) vs. those
without concurrent RI. For COPD and
pneumonia admissions, diagnosis of
conduction disorders or dysrhythmias
(Dyshr.) increased PM10 RR estimate.  The
PM10 RR effect size did not vary
significantly by sex, age, or race, but
baseline risks across these groups differ
markedly, making such sub-population RR
inter-comparisons difficult to interpret.
                                                                                                                            All Age Asthma HA's
                                                                                                                            PM10 = 50 Mg/m3, no co-pollutant, during wet
                                                                                                                            season (Jan. 1 - Mar. 1):

                                                                                                                            All Asthma Hospital Admissions
                                                                                                                            0-d lag PM10 ER = 16.2 (CI: 2.0, 30)
                                                                                                                            8-d avg. PM10 ER = 20.0 (CI: 5.3, 35)

                                                                                                                            MediCal Asthma Hospital Admissions
                                                                                                                            8-d avg. PM10 ER = 13.7 (3.9, 23.4)

                                                                                                                            Other Insurance Asthma HA's
                                                                                                                            8-davg. PM10ER=6.2(-3.6, 16.1)

                                                                                                                            Respiratory HA's for persons 65+ years
                                                                                                                            50 Mg/m3 PM10
                                                                                                                            ER=5.8%(CI:0.5,  11.4)
                                                                                                                   PM10 = 50 ^g/m3(average of lags 0,1)
                                                                                                                   COPD (adults 65+ vrs.)
                                                                                                                   W/o prior RI. ER = 8.8% (CI: 3.3, 14.6)
                                                                                                                   With prior RI ER = 17.1% (CI: -6.7, 46.9)
                                                                                                                   COPD (adults 65+ vrs.)
                                                                                                                   W/o concurrent Dys. ER = 7.2% (CI: 1.3, 13.5)
                                                                                                                   With concurrent Dys. ER = 16.5%(CI: 3.2,
                                                                                                                   31.5)
                                                                                                                   Pneumonia (adults 65+ vrs.)
                                                                                                                   W/o pr. Asthma ER =11% (CI: 7.7, 14.3)
                                                                                                                   Withpr. Asthma ER = 22.8% (CI: 5.1, 43.6)
                                                                                                                   Pneumonia (adults 65+ vrs.)
                                                                                                                   W/o pr. Dyshr. ER = 10.4% (CI: 6.9, 14)
                                                                                                                   With pr. Dyshr. ER = 18.8% (CI: 6.3, 32.7)
O

-------
to
o
o
to
                   TABLE 8B-2 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
                                                                     ADMISSIONS STUDIES
          Reference/Citation
          Location, Duration
          PM Index/Concentrations
                                                     Study Description:
        Results and Comments
       PM Index, Lag, Excess Risk %,
     (95% CI = LCI, UCL) Co-Pollutants
oo
td
to
to
fe
H
6
o
o
H
O
O
H
W
O
O
          United States (cont'd)
          Lippmann et al. (2000)
          Detroit, MI ('92-'94)
          Population = 2.1 million
          PM10 Mean = 31 /-ig/m3
          (IQR=19, 38Mg/m3;
          max=105,ug/m3)
          PM25Mean= IS^g/m3
          (IQR= 10, 21  Mg/m3; max=86
          Mg/m3)
          PM10.25 Mean = 12 /-ig/m3
          (IQR= 8, 17 Mg/m3; max=50 M
          SO4TVIean = 5 Mg/m3
          (IQR=1.8, 6.3,ug/m3;
          max=34.5 ^g/m3)
          H+ Mean =8.8 nmol/m3 = 0.4
          (IQR=0, 7nmol/m3;max=279)
                                          Respiratory (COPD and Pneumonia) HA's
                                          for persons 65 + yr. analyzed, using GAM
                                          Poisson models, adjusting for season, day
                                          of week, temperature, and relative humidity.
                                          The air pollution variables analyzed were:
                                          PM10, PM2 5, PM10.2 5, sulfate, H+, O3, SO2,
                                          NO2, and CO. However, this study
                                          site/period had very low acidic aerosol
                                          levels. As noted by the authors 85% of H+
                                          data was below detection limit (8 nmol/m3).
For respiratory HA's, all PM metrics
yielded RR's estimates >1, and all were
significantly associated in single pollutant
models for pneumonia. For COPD, all
PM metrics gave RR's >1, with H+ being
associated most significantly, even after
the addition of O3 to the regression.
Adding gaseous pollutants had negligible
effects on the various PM metric RR
estimates. The most consistent effect of
adding co-pollutants was to widen the
confidence bands on the PM metric RR
estimates: a common statistical artifact of
correlated predictors. Despite usually
non-detectable levels, H+ had strong
association with respiratory admissions on
the few days it was present.  The general
similarity of the PM25 and PM10_25 effects
per Mg/ni3 in this study suggest similarity
in human toxicity of these two inhalable
mass components in study locales/periods
where PM2 5 acidity is usually not present.
Pneumonia HA's for 65+ yrs.
No co-pollutant:
PM10 (50 Mg/m3) Id lag
  ER = 22%(CI: 8.3,36)
PM25(25Mg/m3)ldlag:
  ER=13%(CI: 3.7,22)
PM25.10(25Mg/m3)ldlag:
  ER=12%(CI:0.8,24)
H+ (75 nmol/m3) 3d lag:
  ER=12%(CI:0.8,23)
 O, co-pollutant (lag 3) also in model:
PM10 (50 Mg/m3) Id lag,
  ER = 24% (CI: 8.2, 43)
PM25(25Mg/m3)ldlag:
  ER=12%(CI: 1.7,23)
PM25.10(25Mg/m3)ldlag:
  ER = 14% (CI: 0.0, 29)
H+ (75 nmol/m3) 3d lag:
  ER=11%(CI: -0.9,24)
COPD Hospital Admissions for 65+ yrs.
 No co-pollutant:
                                                                                                                         PM10 (50
                                                                                                                                       3d lag
                                                                                                                           ER = 9.6%(CI: -5.1,27)
                                                                                                                        PM25(25,ug/m3)3dlag:
                                                                                                                           ER = 5.5%(CI: -4.7,17)
                                                                                                                        PM25.10(25Aig/m3)3dlag:
                                                                                                                           ER = 9.3%(CI: -4.4,25)
                                                                                                                        H+ (75 nmol/m3) 3d lag:
                                                                                                                           ER=13%(CI:0.0,28)
                                                                                                                         O, co-pollutant (lag 3) also in model:
                                                                                                                        PM10 (50 Mg/m3) 3d lag,
                                                                                                                           ER=1.0%(-15,20)
                                                                                                                        PM25(25Mg/m3)3dlag:
                                                                                                                           ER = 2.8%(CI: -9.2,16)
                                                                                                                        PM25.10(25Mg/m3)3dlag:
                                                                                                                           ER = 0.3%(CI: -14, 18)
                                                                                                                        H+ (75 nmol/m3) 3d lag:
                                                                                                                           ER=13%(CI: -0.6,28)

-------
to
o
o
to
           TABLE 8B-2 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                       Study Description:
        Results and Comments
       PM Index, Lag, Excess Risk %,
     (95% CI = LCI, UCL) Co-Pollutants
oo
td
to
fe
H
6
o
o
H
O
O
H
W
O
          United States (cont'd)

          Lumley and Heagerty (1999)
          Seattle (King Cty.), WA (87-94)
          Population = NR
          PM[ daily mean = NR
          PM^K) daily mean = NR
          From Sheppard et al,  1999:
          PM10 mean = 31.5 /-ig/m3
          PM10IQR=19-39Mg/m3
Moolgavkar et al. (2000)
King County, WA (87 - 95)
Population = NR
PM10 mean = 30.0 /-ig/m3
PM10IQR =18.9-37.3 Mg/m3
PM2 5 mean =18.1 ,ug/m3
PM2^IQR=10-23,ug/m3
                                   Estimating equations based on marginal
                                   generalized linear models applied to
                                   respiratory HA's for persons <65 yrs. of age
                                   (mean ~ 8/day) using class of variance
                                   estimators based upon weighted empirical
                                   variance of the estimating functions.
                                   Poisson regression used to fit a marginal
                                   model for the log of admissions with linear
                                   temperature, day of week, time trend, and
                                   dummy season variables. No co-pollutants
                                   considered.
                                   Association between air pollution and
                                   hospital admissions (HA's) for COPD
                                   (all age mean=7.75/day; 0-19 yrs.
                                   mean=2.33/day) investigated using Poisson
                                   GAM's controlling for day-of-week,
                                   season, and LOESS of temperature. Co-
                                   pollutants addressed:  O3, SO2, CO, and
                                   pollens. PM25 only had one monitoring site
                                   versus multiple sites averaged for other
                                   pollutants.
PM[ at lag 1 day associated with
respiratory HA's in children and younger
adults (<65), but not PM1(ri, suggesting a
dominant role by the submicron particles
in PM2 5-asthma HA associations reported
by Sheppard et al. (1999). 0-day lag PM;
and 0 and 1 day lag PM^^ had RR near 1
and clearly non-significant. Authors note
that model residuals correlated at r=0.2,
suggesting the need for further long-wave
controls in the model (e.g., inclusion of
the LOESS of HA's).

Of the PM metrics, PM10 showed the most
consistent associations across lags (0-4 d).
PM25 yielded the strongest positive PM
metric association at Iag3  days, but gave a
negative association at Iag4 days.  That
PM2 5 only had one monitoring site may
have contributed to its effect estimate
variability. Residual autocorrelations (not
reported) may also be a factor. Adding
gaseous co-pollutants or pollens decreased
the PM2 5 effect estimate less than PM10.
Analyses indicated that asthma HA's
among the young were driving the overall
COPD-air pollution associations.
Respiratory HA's for persons <65 vrs. old
PM[ = 25 Mg/m3, no co-pollutant:

1-dlag ER= 5.9 (1.1, 11.0)
COPD HA's all ages (no co-pollutant)
PM10 (50 Mg/m3, lag 2)
      ER = 5.1%(CI: 0,10.4)
PM25(25^g/m3,lag3)
      ER = 6.4% (CI: 0.9, 12.1)

COPD HA's all ages (CO as co-pollutant)
PM10 (50 Mg/m3, lag 2)
      ER = 2.5%(CI: -2.5,7.8)
PM25(25Mg/m3,lag3)
      ER = 5.6% (CI: 0.2, 11.3)
o

-------
to
o
o
to
            TABLE 8B-2 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                       Study Description:
                                                 Results and Comments
                                               PM Index, Lag, Excess Risk %,
                                             (95% CI = LCI, UCL) Co-Pollutants
oo
td
to
fe
H
6
o
o
H
O
O
H
W
O
 United States (cont'd)

 Moolgavkar (2000a)
 Study Period: 1987-1995

 Chicago (Cook County), IL
 Population = NR
 PM10 median = 35 /j.g/w?
 PM10IQR = 25-47 ,ug/m3

 Los Angeles (LA County), CA
 Population = NR
 PM10 median = 44 /-ig/m3
 PM10 IQR = 33-59 Mg/m3
 PM2 5 median = 22 /-ig/m3
 PM2^5IQR=15-3lMg/m3

 Phoenix (Maricopa County), AZ
 Population = NR
 PM10 median = 41 /j.g/w?
 PM10 IQR = 32-5 l,ug/m3
Investigated associations between air
pollution (PM10, O3, SO2, NO2, and CO)
and COPD Hospital Admissions (HA's).
PM25 also analyzed in Los Angeles. HA's
for adults >65 yr.:  median=12/day in
Chicago, =4/d in Phoenix; =20/d in LA.
In LA, analyses also conducted for children
0-19 yr.  (med.=17/d) and adults 20-64
(med.=24/d). Poisson GAM's used
controlling for day-of-week, season, and
splines of temperature and RH (but not their
interaction) adjusted for overdispersion.
PM data available only every 6th day
(except for daily PM10 in Chicago), vs.
every day for gases. Power likely differs
across pollutants, but number of sites and
monitoring days not presented.  Two
pollutant models forced to have same lag
for both pollutants. Autocorrelations or
intercorrelations of pollutant coefficients
not presented or discussed.
For >64 adults, CO, NO2 and O3 (in
summer) most consistently associated with
the HA's. PM effects more variable,
especially in Phoenix. Both positive and
negative significant associations for PM
and other pollutants at different lags
suggest possible unaddressed negative
autocorrelation. In LA, PM associated
with admissions in single pollutant
models, but not in two pollutant models.
The forcing of simultaneous pollutants to
have the same lag (rather than maximum
lag), which likely maximizes
intercorrelations between pollutant
coefficients, may have biased the two
pollutant coefficients, but information not
presented..  Analysis in 3 age groups in
LA yielded similar results. Author
concluded that "the gases, other than
ozone, were more strongly associated with
COPD admissions than PM, and that there
was considerable heterogeneity in the
effects of individual pollutants in different
geographic areas".
                                                                                                                            Most Significant Positive ER
                                                                                                                            Single Pollutant Models:
                                                                                                                            COPD HA's (>64vrs.) (50 Mg/m3 PM10):
                                                                                                                            Chicago: Lag 0 ER =2% (CI: -0.2, 4.3)
                                                                                                                            LA:     Lag 2 ER = 6. 1%(CI: 1.1,11.3)
                                                                                                                            Phoenix: Lag 0 ER = 6.9% (CI: -4.1, 19.3)
                                                                                                                            LA COPD HA's
(50
         PM10, 25
                                                                                                                                                  3 PM2.5 or PM1
(0-19 yrs.): PM10 lg2=10.7%(CI: 4.4, 17.3)
(0-19 yrs.): PM25 lgO=4.3%(CI: -0.1, 8.9)
(0-19 yrs.): PM10.25 lg2=17.1%(CI: 8.9, 25.8)
(20-64 yrs.): PM10  lg2=6.5%(CI: 1.7, 11.5)
(20-64 yrs.): PM25 lg2=5.6%(CI: 1.9, 9.4)
(20-64 yrs.): PM10.2.5 lg2=9%(CI: 3, 15.3)

(> 64 yrs): PM10 Ig2 = 6.1% (1.1, 11.3)
(> 64 yrs): PM2.5 Ig2 = 5.1% (0.9, 9.4)

(>64 yrs.): PM10.25 Ig3=5.1% (CI: -0.4, 10.9)

(>64 yr) 2 Poll. Models (CO = co-poll.)

PM10: Lag 2 ER= 0.6% (CI: -5.1,6.7)
PM25: Lag 2 ER =  2.0% (-2.9, 7.1)
O

-------
to
o
o
to
            TABLE 8B-2 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                       Study Description:
                                                 Results and Comments
                                               PM Index, Lag, Excess Risk %,
                                             (95% CI = LCI, UCL) Co-Pollutants
oo
td
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
         United States (cont'd)

         Sheppard et al. (1999)
         Seattle, WA, Pop. = NR
         1987-1994
         PM10 mean = 31.5 ,ug/rn3
         PM10IQR= 19-39Aig/m3
         PM25 mean =16.7 /j.g/m3
         PM25.10 mean = 16.2 ,ug/rn3
         PM2^.10IQR=9-21,ug/m3
 Friedman etal. (2001)
 Atlanta, GA
 Summer 1996/control vs. Olympics
 PM10 decrease for 36.7 ,ug/m3 to
 30.8,ug/m3
         Zanobetti and Schwartz (2001)
         Cook County, Illinois
         1988-1994
         PM10:  33 Mg/ni3 median
         Janssen et al. (2002)
         14 U.S. cities
         1985-1994
         see Samet et al. (2000a,b)
Daily asthma hospital admissions (HA's)
for residents aged <65 (mean=2.7/day)
regressed on PM10, PM2 5, PM2 5.10, SO2, O3,
and CO in a Poisson regression model with
control for time trends, seasonal variations,
and temperature-related weather effects.
Appendicitis HA's analyzed as a control.
Except O3 in winter, missing pollutant
measures estimated in a multiple imputation
model. Pollutants varied in number of sites
available for analysis, CO the most (4) vs. 2
forPM.

Asthma events in children aged 1 to
16 years were related to pollutant levels
contrasting those during the Summer
Olympics games during a 17 day period to
control periods before and afer the
Olympics.
                                   Respiratory admissions for lung disease in
                                   persons with or without diabetes as a
                                   co-morbidity related to PMIO measures.
                                   Regression coefficients of the relation
                                   between PM10 and hospital admissions for
                                   respiratory disease from Samet et al.
                                   (2000a,b) and prevalence of air
                                   conditioning (AC).
Asthma HA's significantly associated with
PM10, PM25, and PM10_25 mass lagged 1
day, as well as CO.  Authors found PM
and CO to be jointly associated with
asthma admissions.  Highest increase in
risk in spring and fall. Results conflict
with hypothesis that wood smoke (highest
in early study years  and winter) would be
most toxic.  Associations of CO with
respiratory HA's taken by authors to be an
index of incomplete combustion, rather
than direct CO biological effect.

Asthma events were reduced during the
Olympic period. A  significant reduction
in asthma events was associated with
ozone concentration. The high correlation
between ozone and PM limit the ability to
determine which pollutants may have
accounted for the reduction in asthma
events.

Weak evidence that diabetes modified the
risks of PM10 induced respiratory hospital
admissions while diabetes modified the
risk of PM10 induced COPD admissions in
older people. Found a significant
interaction with hospital admissions for
heart disease and PM with more than
twice the risk in diabetics as in persons
without diabetes.

Regression coefficients of the relation
between ambient PM10 and hospital
admissions for COPD decreased with
increasing percentage of homes with
central AC.
                                                                                                                   Asthma Admissions (ages 0-64)
                                                                                                                   PM10 (lag=lday); 50 Mg/m3
                                                                                                                   ER= 13.7%(CI: 5.5%, 22.6)
                                                                                                                   PM25 (lag=lday); 25 Mg/m3
                                                                                                                   ER=8.7%(CI: 3.3%, 14.3)
                                                                                                                   PM25.10 (lag=lday); 25 Mg/m3
                                                                                                                   ER= 11.1%(CI:2.8%,20.1)
                                                                                                                            3 day cumulative exposure PM10
                                                                                                                            per 10 Mg/m3
                                                                                                                            1.0(0.80-2.48)
                                                                                COPD
                                                                                PM10
                                                                                10Aig/m3
                                                                                with diabetes
                                                                                2.29 (-0.76-5.44)
                                                                                without diabetes
                                                                                1.50(0.42-2.60)

-------
to
o
o
to
            TABLE 8B-2 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                       Study Description:
                                                                                              Results and Comments
                                               PM Index, Lag, Excess Risk %,
                                             (95% CI = LCI, UCL) Co-Pollutants
oo
td
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Canada

          Burnett etal. (1997b)
          Toronto, Canada (1992-1994),
          Pop. = 4 mill.
          PM25 mean =16.8 Mg/ni3
          PM2^5 IQR = 8-23 Mg/m3
          PM10.2 5 mean =11.6 Mg/m3
 PM1
               5 IQR = 7-14
PM10 mean = 28.4 ,ug/m3
PM10IQR= 16-38 Mg/m3
CoH mean = 0.8 (per 103 lin. ft.)
CoHIQR = 0.5-1. l(per 103 lin ft)
SO4 mean = 57.1 nmole/m3
SO4 IQR =14-71 nmole/m3
H+ mean = 5 nmole/m3
H+ IQR = 0-6 nmole/m3

Burnett etal. (1999)
Metro-Toronto, Canada
1980-1994

Pollutant: mean, median, IQR:
FPest (Mg/m3):  18,16,10
CPest (Mg/m3):  12,10,8
PM10 est (Mg/m3):  30,27,15
                                            Hospital admissions (HA's) for respiratory
                                            diseases (tracheobronchitis, chronic
                                            obstructive long disease, asthma,
                                            pneumonia) analyzed using Poisson
                                            regression (adjusting for long-term temporal
                                            trends, seasonal variations, effects of short-
                                            term epidemics, day-of-week, ambient
                                            temperature and dew point).  Daily particle
                                            measures: PM25, coarse particulate
                                            mass(PM10_25), PM10, SO4, H+, and gaseous
                                            pollutants (O3, NO2, SO2, and CO)
                                            evaluated.
                                            Daily hospitalizations for asthma (493,
                                            mean 1 I/day), obstructive lung disease
                                            (490-492, 496, mean 5/day), respiratory
                                            infection (464, 466, 480-487, 494, mean
                                            13/day) analyzed separately in relation to
                                            environmental covariates. Same geographic
                                            area as in Burnett et al., 1997b. Three size-
                                            classified PM metrics were estimated, not
                                            measured, based on a regression on TSP,
                                            SO4, and COH in a subset of every 6th-day
                                            data. Generalized additive models. Non-
                                            parametric LOESS prefilter applied to both
                                            pollution and hospitalization data. Day of
                                            week controls. Tested 1-3 day averages of
                                            air pollution ending on lags 0-2. Covariates:
                                            O3, NO2, SO2, CO, temperature, dewpoint
                                            temperature, relative humidity.
Positive air pollution-HA associations
found, with ozone being pollutant least
sensitive to adjustment for co-pollutants.
However, even after the simultaneous
inclusion of O3 in the model, the
association with the respiratory hospital
admissions were still significant for PM10,
PM25, PM10.25, CoH,, S04, and H+.
                                                                            In univariate regressions, all three PM
                                                                            metrics were associated with increases in
                                                                            respiratory outcome. In multi-pollutant
                                                                            models, there were no significant PM
                                                                            associations with any respiratory outcome
                                                                            (results not shown).  Use of estimated PM
                                                                            metrics limits the interpretation of
                                                                            pollutant-specific results reported.
                                                                            However, results suggest that a linear
                                                                            combination of TSP, SO4, and COH does
                                                                            not have a strong independent association
                                                                            with cardiovascular admissions when a
                                                                            full range of gaseous pollutants are also
                                                                            modeled.
Respiratory HA's all ages(no co-pollutant)
PM10 (50 Mg/m3, 4d avg. lag 0)
   ER = 10.6% (CI: 4.5-17.1)
PM25 (25 Mg/m3, 4d avg. lag  1)
   ER = 8.5%(CI: 3.4,13.8)
PM10.25 (25 Mg/m3, 5d avg. lag 0)
   ER=12.5%(CI: 5.2,20.0)
Respiratory HA's all ages(O, co-pollutant)
PM10 (50 Mg/m3, 4d avg. lag 0)
   ER = 9.6% (CI: 3.5, 15.9)
PM25 (25 Mg/m3, 4d avg., lag 1)
   ER = 6.2% (1.0, 11.8)
PM10.25 (25 Mg/m3, 5d avg. lag 0)
   ER = 10.8% (CI: 3.7, 18.1)
                                        Percent excess risk (95% CI) per 50 ,
                                        PM10; 25 Mg/m3 PM25 and PM10.25:
                                        Asthma
                                        PM25 (0-1-2 d):  6.4(2.5,10.6)
                                        PM10 (0-1 d): 8.9 (3.7, 14.4)
                                        PM10.2.5 (2-3-4 d): 11.1(5.8,16.6)

                                        COPD
                                        PM25: 4.8 (-0.2, 10.0)
                                        PM10: 6.9(1.3,12.8)
                                        PM10.25(2-3-4d):  12.8(4.9,21.3)

                                        Resp. Infection:
                                        PM25: 10.8(7.2, 14.5)
                                        PM10: 14.2(9.3,19.3)
                                        PM10.25 (0-1-2 d):  9.3 (4.6, 14.2)
O

-------
to
o
o
to
            TABLE 8B-2 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                       Study Description:
                                                 Results and Comments
                                               PM Index, Lag, Excess Risk %,
                                             (95% CI = LCI, UCL) Co-Pollutants
oo
td
to
         Canada (cont'd)

         Burnett etal. (1997c)
         16 Canadian CitiesO 81-91)
         Population=12.6 MM
         CoHmean=0.64(per 103 lin. ft)
         CoH IQR=0.3-0.8(per 103 lin ft)
 Burnett etal. (2001)
 Toronto, Canada
 1980-1994
 PM25: 18Mg/m3
 PM10.2.5:  16.2 Mg/m3
 (both estimated values)
Air pollution data were compared to
respiratory hospital admissions
(mean=l .46/million people/day) for
16 cities across Canada. Used a random
effects regression model, controlling for
long-wave trends, day of week, weather,
and city-specific effects.

Respiratory admissions in children aged
<2 years relates to mean pollution levels.
O3, NO2, SO2, and CO
(ICD-9:  493 asthma; 466 acute bronchitis;
464.4 croup or pneumonia, 480-486).
Time-series analysis adjusted with LOESS.
The 1 day lag of 03 was positively
associated with respiratory admissions in
the April to December period, but not in
the winter months. Daily maximum 1-hr.
CoH from 11 cities and CO also positively
associated with HA's, even after
controlling for O3.

Summertime urban air pollution,
especially ozone,  increases the risk that
children less than 2 years of age will be
hospitalized for respiratory disease.
                                                                                                                   Respiratory HA's all ages (with O,,CO)
                                                                                                                   CoHIQR = 0.5,lagO:
                                                                                                                   CoH ER = 3.1% (CI: 1.0-4.6%)
                                                                                                                            PM25lagO
                                                                                                                            15.8%(t=3.29)
                                                                                                                            PM25lagO
                                                                                                                            with O3 1.4% (0.24)

                                                                                                                            PM10.2.5 lag 1
                                                                                                                            18.3%(t=3.29)
                                                                                                                            with O3 4.5% (0.72)
fe
H
6
o
o
H
O
O
H
W
O
O
         Europe

         Atkinson et al. (1999b)
         London (92 - 94)
         Population = 7.2 MM
         PM10Mean = 28.5
         10th-90thIQR = 15.8-46.5
         BS mean = 12.7 Mg/m3
         10th-90thIQR = 5.5-21.6 M
                                   All-age respiratory (mean=150.6/day), all-
                                   age asthma (38.7/day), COPD plus asthma
                                   in adults >64 yr. (22.9/day), and lower
                                   respiratory (64.1/day) in adults >64 yr
                                   (16.7/day) hospital admissions in London
                                   hospitals considered. Counts for ages 0-14,
                                   15-64, and >64 yr also examined.  Poisson
                                   regression used, controlling for season, day-
                                   of-week, meteorology, autocorrelation,
                                   overdispersion, and influenza epidemics.
                                         Positive associations found between
                                         respiratory-related emergency hospital
                                         admissions and PM10 and SO2, but not for
                                         O3orBS. When SO2 and PM10 included
                                         simultaneously, size and significance of
                                         each was reduced. Authors concluded that
                                         SO2 and PM10 are both indicators of the
                                         same pollutant mix in this city. SO2 and
                                         PM10 analyses by temperature tertile
                                         suggest that warm season effects
                                         dominate. Overall, results consistent with
                                         earlier analyses for London, and
                                         comparable with those for North America
                                         and Europe.
                                       PM10 (50 Mg/ni3), no co-pollutant.
                                       All Respiratory Admissions:
                                       All age (lag Id) ER = 4.9% (CI: 1.8, 8.1)
                                       0-14 y (lag Id) ER = 8.1% (CI: 3.5, 12.9)
                                       15-64y (lag 2d) ER = 6.9% (CI: 2.1, 12.9)
                                       65+ y (lag 3d) ER = 4.9% (CI: 0.8, 9.3)
                                       Asthma Admissions:
                                       All age (lag 3d) ER = 3.4% (CI: -1.8, 8.9)
                                       0-14y(lag3d)ER = 5.4%(CI: -1.2, 12.5)
                                       15-64 y(lag 3d) ER= 9.4% (CI: 1.1, 18.5)
                                       65+y.(lag Od) ER = 12% (CI: -1.8,27.7)
                                       COPD & Asthma Admissions (65+yrs.)
                                       (lag 3d) ER = 8.6% (CI: 2.6, 15)
                                       Lower Respiratory Admissions (65+ yrs.)
                                       (lag 3d) ER = 7.6% (CI: 0.9, 14.8)

-------
to
o
o
to
            TABLE 8B-2 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS  STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                       Study Description:
                                                 Results and Comments
                                               PM Index, Lag, Excess Risk %,
                                             (95% CI = LCI, UCL) Co-Pollutants
         Europe (cont'd)
oo
td
to
oo
fe
H
6
o
o
H
O
O
H
W
O
O
         Wordleyetal. (1997)
         Study Period: 4/92 -3/94
         Birmingham, UK
         Population = NR
         PM10 daily values:
         Mean = 25.6 /j.g/m3
         range = 2.8,  130.9 ^g/m3
         PM10 3 day running, mean:
         Mean = 25.5 /-ig/m3
         range = 7.3,  104.7 Mg/m3
 Prescott et al. (1998)
 Edinburgh (10/92-6/95)
 Population = 0.45 MM
 PM10 mean. =20.7 /j,g/m3
 PM10 min/max=5/72 /-ig/m
 PM10 90°% -10°% = 20
         McGregor et al. (1999)
         Birmingham, UK.
         Population = NR
         MeanPM,0=30.0ug/m3
                                   Relation between PM10 and total HA's for
                                   respiratory (mean = 21.8/d), asthma
                                   (mn.=6.2/d), bronchitis (mn.=2.4/d),
                                   pneumonia (mn.=3.4/d), and COPD
                                   (mn.=3.2/d) analyzed, using linear
                                   regression after adjusting for day of week,
                                   month, linear trend, RH, and T (but not T-
                                   RH interaction). RR's compared for
                                   various thresholds vs. mean risk of HA.
Poisson log linear regression models  used
to investigate relation of daily HA's with
NO2, O3, CO, and PM10. Adjustments made
for seasonal and weekday variation, daily T
(using 8 dummy variables), and wind speed.
Separate analyses for age<65 yr. (mean resp
HA = 3.4/day) and age >64 yr. (mean resp
HA = 8.7/day), and for subjects with
multiple HA's.
                                   A synoptic climatological approach used to
                                   investigate linkages between air mass types
                                   (weather situations), PM10, and all
                                   respiratory hospital admissions (mean=
                                   19.2/day) for the Birmingham area.
PM10 positively associated with all HA's
for respiratory, asthma, bronchitis,
pneumonia, and COPD.  Pneumonia, all
respiratory, and asthma HA's also
significantly positively associated with the
mean of PM10 over the past three days,
which gave 10 to 20% greater RR's per 10
Mg/m3, as expected given smaller day to
day deviations. Other air pollutants
examined but not presented, as "these did
not have a significant association with
health outcomes independent from that of
PM10".

The two strongest findings were for
cardiovascular HA's of people aged >64,
which showed a positive association with
PM10 as a mean of the 3 previous days.
PM10 was consistently positively
associated with Respiratory HA's in both
age groups, with the greatest effect size in
those >64, especially among those with
>4 HA's during '81-'95. Weak
significances likely contributed to by low
population size.

Study results show distinct differential
responses of respiratory admission rates to
the six winter air mass types. Two of
three types of air masses associated with
above- average admission rates also favor
high PM10 levels.  This is suggestive of
possible linkage between weather, air
quality, and health.
50
                                                                                         in PM1
                                                                                All Respiratory HA's (all ages)
                                                                                (lagOd) ER = 12.6% (CI: 5.7, 20)
                                                                                Asthma HA's (all ages)
                                                                                (Iag2d) ER = 17.6% (CI: 3, 34.4)
                                                                                Bronchitis HA's (all ages)
                                                                                (lagOd) ER= 32.6% (CI: 4.4, 68.3)
                                                                                Pneumonia HA's (all ages)
                                                                                (lag3d)ER=31.9%(CI: 15,51.4)
                                                                                COPD HA's (all ages)
                                                                                (lagld)ER= 11.5%(CI: -3,28.2)
Single Pollutant Models
PM10 = 50 Mg/m3, mean of lags 1-3

Respiratory HA's (age<65)
ER= 1.25 (-12.8, 17.5)
Respiratory HA's (age>64)
ER= 5.33 (-9.3, 22.3)
Respiratory HA's (age>64, >4 HA's)
ER= 7.93 (-19.0,43.7)
                                                                                NR

-------
to
o
o
to
           TABLE 8B-2 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                      Study Description:
                                                 Results and Comments
                                                                                                                                   PM Index, Lag, Excess Risk %,
                                                                                                                                 (95% CI = LCI, UCL) Co-Pollutants
oo
td
to
VO
fe
H
6
o
o
H
O
O
H
W
O
         Europe (cont'd)

         Hagen et al. (2000)
         Drammen, Sweden( 11794-12/97)
         Population = 110,000
         PM10 mean = 16.8 Mg/m3
         PM10IQR= 9.8-20.9 Mg/m3
 Dab etal. (1996)
 Paris, France (87 - 92)
 Population =6.1 MM
 PM13 mean = 50.8 ,ug/rn3
 PM13 S^S111 range = 19.0-137.3
         BS S^S111 Range =11. 0-123. 3
         Anderson etal. (1997)
         Amsterdam(77 - 89)
         Barcelona ( 86- 92)
         London (87-91)
         Milan ( 80- 89)
         Paris ( 87 - 92)
         Rotterdam ( 77 - 89)
         Populations .= 0.7(A), 1.7(B),
         7.2(L),1.5(M),6.5(P),0.6(R)MM
         BS Means = 6, 41, 13, -, 26, 22
         TSP Means = 41,155, -, 105, -,41
Examined PM10, SO2, NO2, VOC's, and O3
associations with respiratory hospital
admissions from one hospital (mean =
2.2/day).  Used Poisson GAM controlling
for temperature and RH (but not their
interaction), long-wave and seasonality,
day-of-week, holidays, and influenza
epidemics.

Daily mortality and general admissions to
Paris public hospitals for respiratory causes
were considered (means/day:  all
resp.=79/d, asthma=14/d, COPD=12/d).
Time series analysis used linear regression
model followed by a Poisson regression.
Epidemics of influenza A and B,
temperature, RH, holidays, day of week,
trend, long-wave variability, and nurses'
strike variables included. No two pollutant
models considered.

All-age daily hospital admissions (HA's)
for COPD considered in 6 APHEA cities;
Mean/day = l.l(A), 11(B), 20(L), 5(M),
11(P), 1.1 (R). Poisson regression
controlling for day of week, holidays,
seasonal and other cycles, influenza
epidemics, temperature, RH, and
autocorrelation. Overall multi-city
estimates made using inverse variance wts.,
allowing for inter-city variance.
                                                                           As a single pollutant, the PM10 effect was
                                                                           of same order of magnitude as reported in
                                                                           other studies. The PM10 association
                                                                           decreased when other pollutants were
                                                                           added to the model. However, the VOC's
                                                                           showed the strongest associations.
                                                                                    For the all respiratory causes category, the
                                                                                    authors found "the strongest association
                                                                                    was observed with PM13" for both hospital
                                                                                    admissions and mortality, indicating a
                                                                                    coherence of association across outcomes.
                                                                                    Asthma was  significantly correlated with
                                                                                    NO2 levels, but not PM13.
                                                                           Ozone gave the most consistent
                                                                           associations across models. Multi-city
                                                                           meta-estimates also indicated associations
                                                                           for BS and TSP.  The warm/cold season
                                                                           RR differences were important only for
                                                                           ozone, having a much stronger effect in
                                                                           the warm season. COPD effect sizes
                                                                           found were much smaller than in U.S.
                                                                           studies, possibly due to inclusion of non-
                                                                           emergency admissions or use of less
                                                                           health-relevant PM indices.
Respiratory Hospital Admissions(all ages)
For IQR=50 Mg/m3
-Single Pollutant Model:
PM10 (lag 0) ER = 18.3% (CI: -4.2, 46)
-Two Pollutant Model (with O3):
PM10 (lag 0) ER = 18.3% (CI: -4.2, 45.4)
-Two Pollutant Model (with Benzene):
PM10 (lag 0) ER = 6.5% (CL-14 ,31.8)

For PM,, = 50 Mg/m3; BS = 25 ug/m3;
Respiratory HA's (all ages):
PM13 Lag 0 ER = 2.2% (CI: 0.2, 4.3)
BSLagOER=1.0%(0.2, 1.8)
COPD HA's (all ages):
PM13 Lag 2 ER = ~2.3% (CI: -6.7, 2.2)
BS Lag 2 ER = -1.1% (-2.9, 0.6)
Asthma HA's  (all ages):
PM13 Lg 2 ER = -1.3% (CI: -4.6, 2.2)
BS Lg 0 ER = 1.2% (-0.5, 2.9)


BS (25 Mg/m3) Id lag, no co-pollutant:
All Age COPD Hospital Admissions
ER= 1.7% (0.5, 2.97)

TSP (100 Mg/m3) Id lag, no co-pollutant:
All Age COPD Hospital Admissions
ER = 4.45%(CI: -0.53,9.67)
O

-------
to
o
o
to
            TABLE 8B-2 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	
 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                       Study Description:
                                                 Results and Comments
                                               PM Index, Lag, Excess Risk %,
                                             (95% CI = LCI, UCL) Co-Pollutants
oo
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
         Europe (cont'd)

         Diaz etal. (1999)
         Madrid (94 - 96)
         Population = NR
         TSP mean * 40 A/
 Spixetal. (1998)
 London (L) (87 -91)
    Pop. =7.2 Million (MM)
    BSMean= 13 A/g/m3
 Amsterdam (A) (77 - 89)
    Pop. =0.7 MM
    BS Mean = 6 ^g/m3
    TSP mean = 41 ,ug/m3
 Rotterdam (R) (77 - 89)
    Pop. =0.6MM
    BS Mean = 22 Mg/m3
    TSP mean = 41 ,ug/m3
 Paris (P) (87 - 92),
    Pop.= 6.14MM
    BS Mean = 26 Mg/m3
 Milano (M) (80 - 89)
    Pop. = 1.5 MM
    TSP Mean =120
         Vigottietal. (1996)
         Study Period.: 80-89
         Milan, IT
         Population =1.5 MM
         TSP mean = 139.0,ug/m3
         TSP IQR= 82.0, 175.7 Mg
ARIMA modeling used to analyze
emergency respiratory and circulatory
admissions (means/day=7.8,7.6) from one
teaching hospital. Annual, weekly, and 3
day periodicities controlled, but no time
trend included, and temperature crudely fit
with v-shaped linear relationship.


Respiratory (ICD9 460-519) HA's in age
groups 15-64 yr and 65 + yrs. related to
SO2, PM (BS or TSP), O3, and NO2 in the
APHEA study cities using standardized
Poisson models with confounder controls
for day of week, holidays, seasonal and
other cycles, temperature, RH, and
autocorrelation. PM lag considered ranged
from 0-3 day, but varied from city to city.
Quantitative pooling conducted by
calculating the weighted means of local
regression coefficients using a fixed-effects
model when no heterogeneity could be
detected; otherwise, a random-effects model
employed.
                                   Association between adult respiratory HA's
                                   (15-64 yr mean =11.3/day, and 65 + yr
                                   mean =8.8/day) and air pollution evaluated,
                                   using the APHEA protocol. Poisson
                                   regression used with control for weather
                                   and long term trend, year, influenza
                                   epidemics, and season
Although TSP correlated at zero lag with
admissions in winter and year-round, TSP
was never significant in ARIMA models;
so effect estimates not reported for TSP.
Also, found biologically implausible u-
shaped relationship for O3, possibly
indicating unaddressed temperature
effects.

Pollutant associations noted to be stronger
in areas where more than one monitoring
station was used for assessment of daily
exposure.  The most consistent finding
was an increase of daily HA's for
respiratory diseases (adults and elderly)
with O3. The SO2 daily mean was
available in all cities, but SO2 was not
associated consistently with adverse
effects. Some significant PM associations
were seen, although no conclusion related
to an overall particle effect could be
drawn. The effect of BS was significantly
stronger with high NO2 levels on the same
day, but NO2 itself was not associated with
HA's. Authors concluded that "there was
a tendency toward an association of
respiratory admissions with BS, but the
very limited number of cities prevented
final conclusions."

Increased risk of respiratory HA was
associated with both SO2 and TSP.  The
relative risks were similar for both
pollutants. There was no modification of
the TSP effect by SO2 level. There was a
suggestion of a higher TSP effect on
hospital admissions in the cool months.
                                                                                                                   N/A
Respiratory Admissions (BS = 25 ug/
BS (L, A, R, P)
15-64 yrs:  1.4% (0.3, 2.5)
  65+yrs:  1.0% (-0.2, 2.2)
TSP(A,R,M)(100Mg/m3)
15-64 yrs:  2.0 (-2.1, 6.3)
  65+yrs:  3.2 (-1.2, 7.9)
Respiratory HA's
BS (L, A, R, P): Warm (25 Mg/m3)
15-64 yrs:  -0.5% (-5.2, 4.4)
  65+yrs:  3.4% (-0.1, 7.1)
BS (L, A, R, P): Cold (25 Mg/m3)
15-64 yrs:  2,0% (0.8, 3.2)
  65+yrs: 0% (-2.2, 2.3)
TSP (A, R, M): Warm (100 Mg/m3)
15-64 yrs:  6.1% (0.1, 12.5)
  65+yrs:  2.0% (-3.9, 8.3)
TSP (A, R, M): Cold (100 Mg/m3)
15-64 yrs:  -5.9% (-14.2, 3.2)
  65+yrs:  4.0% (-0.9, 9.2)

Young Adult (15-64 yrs.) Resp. HA's
100 Mg/ni3 increase in TSP
Lag 2 ER = 5% (CI: 0, 10)

Older Adult (65+ yrs.) Resp. HA's
100 Mg/ni3 increase in TSP
Lag 1 ER = 5% (CI: -1,10)

-------
to
o
o
to
            TABLE 8B-2 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                       Study Description:
         Results and Comments
       PM Index, Lag, Excess Risk %,
     (95% CI = LCI, UCL) Co-Pollutants
oo
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
         Europe (cont'd)

         Anderson et al. (1998)
         London (87 - 92)
         Population = 7.2 MM
         BS daily mean = 14.6 ,ug/rn3
         BS25-75thIQR = 24-38
         Kontosetal. (1999)
         Piraeus, Athens GR (87 - 92)
         Population = NR
         BS mean =46.5 Mg/m3
         BS max =200 ,ug/m3
         Ponce de Leon et al. (1996)
         London (4/87-2/92)
         Population = 7.3 million
         BS mean. =14.6 Mg/ni3
         BS S^-gS111 %=6 -
                                   Poisson regression used to estimate the RR
                                   of London daily asthma hospital admissions
                                   associated with changes in O3, SO2, NO2
                                   and particles (BS) for all ages and for 0-14
                                   yr. (mean=19.5/d), 15-64 yr. (mean=13.1/d)
                                   and 65 + yr. (mean =2.6/d). Analysis
                                   controlled for time trends, seasonal factors,
                                   calendar effects, influenza epidemics, RH,
                                   temperature, and auto-correlation.
                                   Interactions with co-pollutants and
                                   aeroallergens tested via 2 pollutant models
                                   and models with pollen counts (grass, oak
                                   and birch).

                                   Relation of respiratory HA's for children
                                   (0-14 yrs.) (mean = 4.3/day) to BS, SO2,
                                   NO2, and O3 evaluated, using a
                                   nonparametric stochastic dynamical system
                                   approach and frequency domain analyses.
                                   Long wave and effects of weather
                                   considered, but non-linearity and
                                   interactions of T and RH relation with HA's
                                   not addressed.

                                   Poisson regression analysis of daily counts
                                   of HA's (means/day:  all ages=125.7; Ages
                                   0-14=45.4; Ages 15-64=33.6; Ages
                                   65+=46.7).  Effects of trend, season and
                                   other cyclical factors, day of the week,
                                   holidays, influenza epidemic, temperature,
                                   humidity, and
                                   autocorrelation addressed.  However,
                                   temperature modeled as linear, with no RH
                                   interaction.  Pollution variables were BS,
                                   SO2, O3, andNO2, lagged 0-3 days.
Daily hospital admissions for asthma
found to have associations with O3, SO2,
NO2, and particles (BS), but there was
lack of consistency across the age groups
in the specific pollutant. BS association
was strongest in the 65 + group, especially
in winter. Pollens not consistently
associated with asthma HA's, sometimes
being positive, sometimes negative.  Air
pollution associations with HA's not
explained by airborne pollens in
simultaneous regressions, and there was
no consistent pollen-pollutant interaction.

Pollution found to explain significant
portion of the HA variance.  Of pollutants
considered, BS was consistently among
most strongly explanatory pollutants
across various reported analyses.
O3 associated with increase in daily HA's,
especially in the "warm" season.
However, u-shape of the O3 dose-response
suggests that linear temperature control
was not adequate.  Few significant
associations with other pollutants, but
these tended to be positive (especially in
cold season, Oct-March, and for older
individuals for BS).
Asthma Admissions. BS=25 Mg/m3
BS Lag = 0-3 day average concentration
All age ER= 5.98% (0.4, 11.9)
<15yr.  ER = 2.2% (-4.6, 9.5)
15-64yrER= 1.2% (-5.3, 8.1)
65+ yr. ER = 22.8% (6.1, 42.5)

BS=50 Mg/m3, 2d lag & co-pollutant:
Older Adult (>64 yrs.) Asthma Visits:
BS alone:  ER = 14.6% (2.7, 27.8)
&O3:      ER = 20.0% (3.0, 39.8)
&N02:   ER= 7.4% (-8.7, 26.5)
SO2:      ER= 11.8% (-2.2, 27.8)

NR
Respiratory HA's (all ages)
Single Pollutant Models
For Oct-Mar. BS = 25 ,ug/m3
Lag 1ER = 0.2% (-1.9, 2.3)
For Apr-Sep. BS = 25 ,ug/m3
Lag 1ER =-2.7% (-6.0, 0.8)

Respiratory HA's (>65)
Single Pollutant Models
For Oct-Mar. BS = 25 Mg/m3
Lag 2 ER= 1.2% (-2.1,4.5)
For Apr-Sep. BS = 25 Mg/m3
Lag 2 ER = 4.5% (-1.0, 10.4)

-------
to
o
o
to
           TABLE 8B-2 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                      Study Description:
        Results and Comments
       PM Index, Lag, Excess Risk %,
     (95% CI = LCI, UCL) Co-Pollutants
oo
td
OJ
to
          Europe (cont'd)

          Schouten et al. (1996)
          Amsterdam/Rotterdam (77 - 89)
          Amsterdam Pop. = 0.69 Million
          Rotterdam Pop. = 0.58 Million
          Amsterdam, NE
          BSmean. =11 Mg/ni3
          BS S^-gS^/o =1-37 ,ug/m3
          Rotterdam, NE
          BS mean. =26 Mg/m3
          BS5th-95tll%=6-6lMg/m3
                                  Daily emergency HA's for respiratory
                                  diseases (ICD 460-519), COPD (490-492,
                                  494, 496), and asthma (493). The mean
                                  HA/d (range) for these were: 6.70 (0-23),
                                  1.74 (0-9) and 1.13 (0-7) respectively in
                                  Amsterdam and 4.79 (0-19), 1.57 (0-9),
                                  and 0.53 (0-5) in Rotterdam. HA
                                  associations with BS, O3, NO2, and SO2
                                  analyzed, using autoregressive Poisson
                                  regression allowing for overdispersion and
                                  controlling for season, day of week,
                                  meteorological factors, and influenza
                                  epidemics.
BS did not show any consistent effects in
Amsterdam; but in Rotterdam BS was
positively related to HA's. Most
consistent BS associations in adults >64
yrs. in winter. Positive O3 association in
summer in people aged >64 in Amsterdam
and Rotterdam. SO2 and NO2 did not
show any clear effects. Results not
changed in pollutant interaction analyses.
The authors concluded short-term air
pollution-emergency HA's association is
not always consistent at these individual
cities' relatively low counts of daily HA's
and low levels of air pollution. Analyses
for all ages of all the Netherlands gave a
strong BS-HA association in winter.
Single Pollutant Models
For BS=25 Mg/m3, 2 day lag
For all of the Netherlands:
Respiratory HA's (all ages)
Winter:
ER = 2.0% (-1.5, 5.7)
Summer:
ER = 2.4% (0.6, 4.3)
fe
H
6
o

o
H
O
O
H
W
O
o

-------
to
o
o
to
           TABLE 8B-2 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                                      Study Description:
                                                Results and Comments
                                                                                                                                 PM Index, Lag, Excess Risk %,
                                                                                                                               (95% CI = LCI, UCL) Co-Pollutants
oo
td
fe
H
6
o
o
H
O
O
H
W
O
O
 Europe (cont'd)

 Sunyeretal. (1997)
 Barcelona (86 - 92)
   Population = NR
   BS Median:  40 Mg/m3
   BS Range:  11-258 (B
 Helsinki (86 - 92)
   Population = NR
   BS Median:  -
   BS Range:  -
 Paris (86 - 92)
   Population = NR
   BS Median:  28 Mg/m3
   BS Range:  4-186Mg/m3
 London (86 - 92)
   Population = NR
   BS Median:  IS^g/m3
   BS Range:  3-95 ,ug/m3
         Teniasetal(1998)
         Study Period.: 94-95
         Valencia, Spain
         Hosp. Cachment Pop. =200,000
         BS mean = 57.7 Mg/m3
         BSIQR = 25.6-47.7 Mg/m3
Evaluated relations of BS, SO2, NO2, and
O3 to daily counts of asthma HA's and ED
visits in adults [ages 15-64 years: mean/day
= 3.9 (B); 0.7 (H); 13.1 (H); 7.3 (P)] and
children [ages < 15 years: mean/day = 0.9
(H); 19.8 (L); 4.6 (P)]. Asthma
(ICD9=493) studied in each city, but the
outcome examined differed across cities:
ED visits in Barcelona; emergency hospital
asthma admissions in London and Helsinki,
and total asthma admissions in Paris.
Estimates from all cities obtained for entire
period and also by warm or cold seasons,
using Poisson time-series regression,
controlling for temperature and RH, viral
epidemics, day  of week effects, and
seasonal and secular trends. Combined
associations were estimated using meta-
analysis.
                                   Associations between adult (14+ yrs.)
                                   emergency asthma ED visits to one city
                                   hospital (mean =1.0/day) and BS, NO2, O3,
                                   SO2 analyzed, using Poisson auto-
                                   regressive modeling, controlling for
                                   potential confounding weather and time
                                   (e.g., seasonal) and trends using the
                                   APHEA protocol.
                                                                                   Daily admissions for asthma in adults
                                                                                   increased significantly with increasing
                                                                                   ambient levels of NO2, and positively (but
                                                                                   non-significantly) withBS.  The
                                                                                   association between asthma admissions
                                                                                   and pollution varied across cities, likely
                                                                                   due to differing asthma outcomes
                                                                                   considered. In children, daily admissions
                                                                                   increased significantly with SO2 and
                                                                                   positively (but non-significantly) with BS
                                                                                   and NO2, though the latter only in cold
                                                                                   seasons. No association observed in
                                                                                   children for O3.  Authors concluded that
                                                                                   "In addition to particles, NO2 and SO2 (by
                                                                                   themselves or as a constituent of a
                                                                                   pollution mixture) may be important in
                                                                                   asthma exacerbations".
                                        Association with asthma was positive and
                                        more consistent for NO2 and O3 than for
                                        BS or SO2.  Suggests that secondary
                                        oxidative-environment pollutants may be
                                        more asthma relevant than primary
                                        reduction-environment pollutants (e.g.,
                                        carbonaceous particles). NO2 had greatest
                                        effect on BS in co-pollutant models, but
                                        BS became significant once 1993 was
                                        added, showing power to be a limitation of
                                        this study.
ER per 25 Mg/m3 BS (24 h Average)
Asthma Admissions/Visits:
<15 yrs.:
  London ER=1.5% (IgOd)
  Paris ER= 1.5%(lg2d)
  Total ER=1.5%(-1.1,4.1)
15-64 yrs:
  Barcelona ER = 1.8% (Ig 3d)
  London ER = 1.7% (Ig Od)
  Paris ER= 0.6% (IgOd)
  Total ER= 1.0% (-0.8, 2.9)
Two Pollutant (per 25 ug/m3 BS)
Asthma Admissions (24 h Avg)
<15yrs, (BS&NO2):
  London ER = 0.6% (Ig Od)
  ParisER = 2.9%(lg2d)
  Total ER= 1.8% (-0.6, 4.3)
<15 yrs, (BS & SO2):
  London ER = -1.1 % (Ig Od)
  ParisER=-1.4%(lg2d)
  Total ER=-1.3 (-5.0, 2.5)
15-64 yrs, (BS&NO2):
  Barcelona ER = 1.5% (Ig Od)
  London ER = -4.7% (Ig Od)
  ParisER=-0.7%(lgld)
  Total ER = -0.5% (-5.1, 4.4)

Adult Asthma HA's, BS = 25 Mg/m3
For 1993-1995:
Lag OER= 10.6% (0.9, 21.1)
For 1994-1995:
Lag OER = 6.4% (-4.8, 18.8)

-------
to
o
o
to
           TABLE 8B-2 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                            Study Description:
Results and Comments
PM Index, Lag, Excess Risk %,
(95% CI = LCI, UCL) Co-Pollutants
oo
td
fe
H
6
o
o
H
O
O
H
W
O
O
          Europe (cont'd)

          Anderson et al. (2001)
          West Midland, England
          (October 1994-December 1996)
          Population = 2.3 million
          PM10 mean = 23.3 Mg/m3
          PM25 mean = 14.5 /j.g/m3
          PM10.2.5 = 9.0 Mg/m3
          (by subtraction)
                                   Respiratory hospital admissions (mean =
                                   66/day) related to PM10, PM25, PM10.25, BS,
                                   SO4", NO2, O3, SO2, CO. Regression with
                                   quasilikelihood approach controlling for
                                   seasonal patterns, temp, humidity, influenza
                                   episodes, day week. Adjusted for residual
                                   serial correlation and over-dispersion.
Respiratory admissions (all ages) not
associated with any pollutant. Analyses
by age revealed some associations to PM10
and PM2 5 and respiratory admissions in
the 0-14 age group.  There was a striking
seasonal interaction in the cool season
versus the warm season. PM10_25 effects
cannot be excluded. Two pollutant
models examined particulate measures.
PM2 5 effects reduced by inclusion of
black smoke.
Respiratory HA - lag 0+1 days
PM,n IncrementlO-90% (11.4-38.3 Mg
All ages:  1.5 (-0.7 to 3.6)
Ages 0-14: 3.9 (0.6 to 7.4)
Ages 15-64:  0.1 (-4.0 to 4.4)
Ages>65: -1.1 (-4.3 to 2.1)
PM, ,(6.0-25.8)
All ages:  1.2 (-0.9 to 3.4)
Ages 0-14: 3.4 (-0.1 to 7.0)
Ages 15-64:  -2.1 (-6.4 to 2.4)
Ages>65: -1.3 (-4.7 to 2.2)
PM10.25(4.1tol5.2)
All ages:  0.2 (-2.5 to 3.0)
Ages 0-14: 4.4 (-0.3 to 9.4)
Ages 15-64:  -4.9 (-9.9 to 0.4)
Ages>65: -1.9 (-6.0 to 2.5)

COPD (ICD-9 490-492, 494-496)
EMiou
Age>65: -1.8 (-6.9 to 3.5)
PM,.
Age>65: -3.9 (-9.0 to 1.6)
EMio-2,.
Age>65: -1.7 (-8.9 to 5.3)

Asthma (ICD- 9-493) (mean lag 0+1)
PM10
Ages 0-14: 8.3 (1.7 to 15.3)
Ages 15-64:  -2.3 (-10.0 to 6.1)
EM.2.5.
Ages 0-14: 6.0 (-0.9 to 13.4)
Ages 15-64:  -8.4 (-16.4 to 0.3)
PM10.25
Ages 0-14: 7.1 (-2.1 to 17.2)
Ages 15-64:  -10.7 (-19.9 to-0.5)

-------
to
o
o
to
            TABLE 8B-2 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	
 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                            Study Description:
Results and Comments
                                                                                                                             PM Index, Lag, Excess Risk %,
                                                                                                                             (95% CI = LCI, UCL) Co-Pollutants
oo
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
 Europe (cont'd)

 Atkinson et al. (2001) Eight city
 study: Median/range
 Barcelona 1/94 - 12/96
 PM10  53.3,ug/m3(17.1 - 131.7)
 Birmingham 3/92-12/94
 PM10  21.5Mg/m3(6.5 - 115)
 London 1/92 -  12/94
 PM10  24.9 ,ug/m3 (7.2 - 80.4)
 Milan -No PM10
 Netherlands 1/92 - 9/95
 PM10 33.4 Mg/m3(l 1.3- 130.8)
 Paris 1/92 - 9/96
 PM1020.lMg/m3(5.8- 80.9)
 Rome - No PM10
 Stockholm 3/94 - 12/96
 PM1013.6A/g/m3(4.3-43.3)

 Thompson et al. (2001) Belfast,
 Northern Ireland 1/1/93 - 12/31/95.
 PM10 Mg/m3 mean (SD)
 May-October24.9 (13.7)
 November-April 31.9 (24.3)
         Fusco et al. (2001) Rome, Italy
         1995-1997
         PM - suspended particles measured
                                            As part of the APHEA 2 project,
                                            association between PM10 and daily counts
                                            of emergency hospital admissions for
                                            Asthma (0-14 and 15-64 yrs), COPD and
                                            all-respiratory disease (65+ yrs) controlling
                                            for environmental factors and temporal
                                            patterns were studied.
                                            The rates of acute asthma admission to
                                            children's emergency was studied in
                                            relation to day-to-day fluctuation of PM10
                                            and other pollutants using Poisson
                                            regression.
                                   Daily counts of hospital admissions for total
                                   respiratory conditions, acute respiratory
                                   infection including pneumonia, COPD, and
                                   asthma was analyzed in relation to PM
                                   measures and gaseous pollutants using
                                   generalized additive models controlling for
                                   mean temperature, influenza, epidemics,
                                   and other factors.
This study reports that PM was associated
with daily admissions for respiratory
disease in a selection of European cities.
Average daily ozone levels explained a
large proportion of the between-city
variability in the size of the particle effect
estimates in the over 65 yr age group.  In
children, the particle effects were
confounded with NO2 on a day-to-day
basis.
A weak, but significant association
between PM10 concentration and asthma
emergency-department admissions was
seen. After adjusting for multiple
pollutants only the benzene level was
independently associated with asthma
emergency department admission.
Benzene was highly correlated to PM10,
SO2 and NO2 levels.

No effect was found for PM.  Total
respiratory admissions were significantly
associated with same-day level of NO2 and
CO.  There was no indication that the
effects of air pollution were present at lags
>2 days. Among children, total
respiratory and asthma admissions were
strongly associated with NO2 and  CO.
Multipollutant model analysis yielded
weaker and more unstable results.
                                                                                                                             For 10 Mg/rn3 increase
                                                                                                                             Asthma Admission Age 0-14 yrs:
                                                                                                                             PM10 for cities ranged from -0.9% (-2.1, 0.4) to
                                                                                                                             2.8% (0.8, 4.8) with an overall effect estimate
                                                                                                                             of 1.2% (0.2-2.3)

                                                                                                                             Asthma Admission Age 15-64 yrs:
                                                                                                                             Overall PM 1.1% (0.3-  1.8)

                                                                                                                             Admission of COPD and Asthma Age 65+
                                                                                                                             years:
                                                                                                                             Overall PM 1.0% (0.4 -  1.5)

                                                                                                                             Admission All Respiratory Disease Age 65+
                                                                                                                             years:
                                                                                                                             Overall PM 0.9% (0.6 -1.3)

-------
to
o
o
to
            TABLE 8B-2 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	
 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                            Study Description:
Results and Comments
                                                                                                                             PM Index, Lag, Excess Risk %,
                                                                                                                             (95% CI = LCI, UCL) Co-Pollutants
oo
td
 fe
 H
 6
 o

 o
 H
O
 O
 H
 W
 O
         Europe (cont'd)

         Hrubaetal. (2001)
         (Central Slovakia) (1996)
         TSP 87 Mg/m3
         Latin America

         Bragaetal. (1999)
         Sao Paulo, Brazil (92 - 93)
         Population = NR
         PM10 mean = 66.3 ,ug/rn3
         PM10 Std. Deviation = 26.1
         PM,0 Min./Max. = 26.7/165.4
         Gouveia and Fletcher (2000)
         Study Period. 92-94
         Sao Paulo, Brazil
         Population =9.5 MM x 66%
         PM10 mean = 64.9 Mg/m3
         PM10IQR = 42.9-75.5 Mg/m3
         PM1010/90th%=98.lMg/m3
         PM1095th%=131.6Mg/m3
                                   Logistic regression modeled TSP exposure
                                   and hospital admission for asthma,
                                   bronchitis, or pneumonia in children, ages
                                   7-11 years, N=667.
Pediatric (<13 yrs.) hospital admissions
(mean=67.6/day) to public hospitals serving
40% of the population were regressed
(using both Poisson and maximum
likelihood methods) on air pollutants,
controlling for month of the year, day-of-
week, weather, and the daily number of
non-respiratory admissions
(mean=120.7/day). Air pollutants
considered included PM10, O3, SO2, CO,
andNO2.

Daily public hospital respiratory disease
admissions for children (mean resp. < 5y =
56.1/d; mean pneumonia <5y =40.8/d;
mean asthma <5 y = 8.5/d; mean
pneum.
-------
to
o
o
to
           TABLE 8B-2 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                            Study Description:
Results and Comments
PM Index, Lag, Excess Risk %,
(95% CI = LCI, UCL) Co-Pollutants
oo
td
fe
H
6
o
o
H
O
O
H
W
O
O
         Latin America (cont'd)

         Rosas etal. (1998)
         SW Mexico City (1991)
         Population = NR
         PM10 mean. =77 /j.g/m3
         PM10 min/max= 25/183
         Morgan etal. (1998)
         Sydney, AU (90 - 94)
         Population = NR
         PM25 24 h mean = 9.6 A(g/m3
         PM2^5 10th-90th% = 3.6-18 Mg/
         PM25 max-1 h mean = 22.8 /-i
         PM2^5 10th-90th% = 7.5-44.4 M
                                   Log-regression analysis of relations
                                   between emergency hospital admissions for
                                   asthma for children <15 yrs
                                   (mean=2.5/day), adults (mean=3.0/day),
                                   and adults >59 yrs (mean=0.65/day) and lag
                                   0-2 d pollen, fungal spores, air pollutants
                                   (O3, NO2, SO2, and PM10) and weather
                                   factors. Long wave controlled only by
                                   separating the year into two seasons: "dry"
                                   and "wet". Day-of-week not included in
                                   models.
                                   A Poisson analysis, controlled for
                                   overdispersion and autocorrelation via
                                   GEE, of asthma (means: 0-14
                                   yrs.=15.5/day; 15-64=9/day)), COPD (mean
                                   65+yrs =9.7/day), and heart disease HA's.
                                   PM25 estimated from nephelometry.
                                   Season and weather controlled using
                                   dummy variables.
Few statistical associations were found
between asthma admissions and air
pollutants. Grass pollen was associated
with child and adult admissions, and
fungal spores with child admissions.
Authors conclude that aeroallergens may
be more strongly associated with asthma
than air pollutants, and may act as
confounding factors in epidemiologic
studies.  Results are limited by low power
and the lack of long-wave auto-correlation
controls in the models.
Childhood asthma was primarily
associated with NO2, while COPD was
associated with both NO2 and PM.  1 -hr.
max PM25 more consistently positively
related to respiratory HA's than 24-h avg
PM25. Adding all other pollutants
lowered PM effect sizes, although
pollutant inter-correlations makes many
pollutant model interpretations difficult.
No association found between asthma and
O3 or PM. The authors cited the error
introduced by estimating PM2 5 and the
low PM levels as  possible reasons for the
weak PM-respiratory HA associations.
NR
Asthma HA's
Single Pollutant Model:
For24hrPM25 = 25Mg/m3
1-14 yrs.(lagl) ER = -1.5% (CI: -7.8, 5.3)
15-64 yrs.(lagO) ER = 2.3% (CI: -4, 9)
ForlhPM25=25Mg/m3
1-14 yrs.(lagl)ER = + 0.5% (CI: -1.9, 3.0)
15-64 yrs.(lagO) ER = 1.5%(CI: -0.9,4)
Multiple Pollutant Model:
                                                                                                                           1-14 yrs.(lagl) ER = -0.6% (CI: -7.4, 6.7)
                                                                                                                           COPD (65+yrs.)
                                                                                                                           Single Pollutant Model:
                                                                                                                           (lag 0) ER =4.2% (CI: -1.5, 10.3)
                                                                                                                           (lag 0) ER = 2% (CI: -0.3, 4.4)
                                                                                                                           Multiple Pollutant Model:
                                                                                                                           (lag 0) ER = 1.5% (CI: -0.9, 4)

-------
to
o
o
to
           TABLE 8B-2 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                            Study Description:
                                        Results and Comments
                                       PM Index, Lag, Excess Risk %,
                                       (95% CI = LCI, UCL) Co-Pollutants
oo
td
OJ
oo
fe
H
6
o
o
H
O
O
H
W
O
          Tanakaetal. (1998)
          StdyPd.: 1/92-12/93
          Kushiro, Japan
          Pop. = 102 adult asthmatics
          PM10 mean = 24.0 Mg/m3
          PM10IQR = NR
 Wong etal. (1999)
 Study Period.: 94-95
 Hong Kong
 Population = NR
 PM10 mean = 50.1 /-ig/
 PM10 median = 45.0 Mg
 PM10 IQR =30.7, 65.5
                               g/m3
                                   Associations of HA's for asthma (in 44
                                   non-atopic and 58 atopic patients) with
                                   weather or air pollutants (NO, NO2,
                                   SO2,PM10, O3, and acid fog) evaluated.
                                   Odds ratios (OR) and 95% CFs calculated
                                   between high and low days for each
                                   environmental variable. Poisson regression
                                   was performed for the same dichototomized
                                   variables.
Poisson regression analyses were applied to
assess association of daily NO2, SO2, O3,
and PM10 with emergency HA's for all
respiratory (median = 13I/day) and COPD
(median = 101/day) causes. Effects by age
groups (0-4, 5-64, and 65+ yrs.) also
evaluated. Using the APHEA protocol,
models accounted for time trend, season
and other cyclical factors, T, RH,
autocorrelation and overdispersion. PM10
measured by TEOM, which likely
underestimates mass.
                                        Only the presence of acid fog had a
                                        significant OR >1.0 for both atopies and
                                        non-atopies. PM10 associated with a
                                        reduction in risk (OR<1.0) for both
                                        atopies and non-atopies. Poisson
                                        regression gave a non-significant effect by
                                        PM10 on asthma HA's.  However, no long-   Poisson Coefficient for PM10 > 30
                                       For same-day (lag=0) PM10
                                       Adult Asthma HA's
                                       OR for <30 vs. >30 ,ug/m3PM10:
                                       Non-atopic OR = 0.77 (CI: 0.61, 0.98)
                                       Atopic OR = 0.87 (CI: 0.75, 1.02)
wave or serial auto-correlation controls
applied, so the opposing seasonalities of
PM vs. HA's indicated in time series data
plots are likely confounding these results.

Positive associations were found for HA's
for all respiratory diseases and COPD with
all four pollutants.  PM10 results for lags
0-3 cumulative. Admissions for asthma,
pneumonia, and influenza were associated
with NO2, O3, and PM10. Those aged > or
= 65 years were at higher risk, except for
PM10. No significant respiratory HA
interactions with PM10 effect were found
for high NO2, high O3, or cold season.
                                                                                                                           Non-atopic B = -0.01 (SE = 0.15)
                                                                                                                           Atopic B = -0.002 (SE = 0.09)
           g/m3 (Lags = 0-3 days)
Respiratory HA's
Allage:  ER = 8.3%(CI: 5.1, 11.5)
0-4yrs.:  ER = 9.9% (CI: 5.4, 14.5)
5-64yrs.: ER = 8.8% (CI: 4.3, 13.4)
65+yrs.: ER= 9.3%(CI: 5.1, 13.7)
Asthma HA's (all ages)
ER= 7.7% (1.0, 14.9)
COPD HA's (all ages)
ER= 10.0% (5.6, 14.3)
Pneumonia and Influenza HA's (all ages)
ER= 13.1% (7.2, 19.4)	
O

-------
              Appendix 8B.3: PM-Respiratory Visits Studies
April 2002                         8B-39      DRAFT-DO NOT QUOTE OR CITE

-------
to
o
o
to
to
o
o
                 TABLE 8B-3. ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
Study Description:
Results and Comments
                                                                                                                              PM Index, Lag, Excess Risk %,
                                                                                                                              (95% CI = LCI, UCL) Co-Pollutants
oo
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          United States

          Choudhury et al. (1997)
          Anchorage, Alaska (90 - 92)
          Population = 240,000
          PM10 mean = 41.5 ,ug/rn3
          PM10(SD) = 40.87
          PM10 maximum=565 /j.g/w?
 Lipsettetal. (1997)
 Santa Clara County, CA
 Population = NR
 (Winters 88 - 92)
 PM10 mean = 61.2 /-ig/m3
 PM10 Min/Max = 9/165 /j,g/m3
         Norrisetal. (1999)
         Seattle, WA (9/95-12/96)
         Pop.  Of Children <18= 107,816
         PM10 mean. =21.7 Mg/m3
         PM10IQR=11.6Mg/m3
         asp mean = 0.4 nrl/10~4
         («12.0Mg/m3PM25)
         (=9.5Mg/m3PM25)
Using insurance claims data for state
employees and dependents living in
Anchorage, Alaska, number of daily
medical visits determined for asthma (mean
= 2.42/day), bronchitis, and upper
respiratory infections. Used linear
regression, including a time-trend variable,
crude season indicator variables (i.e.,
spring, summer, fall, winter), and a variable
for the month following a volcanic eruption
in 1992.

Asthma emergency department (ER) visits
from 3 acute care hospitals (mean=7.6/day)
related to CoH, NO2, PM10, and O3 using
Poisson model with long-wave, day of
week, holiday, and weather controls
(analysis stratified by minimum T). Every
other day PM10 estimated from CoH.
Residential wood combustion (RWC)
reportedly a major source of winter PM.
Gastro-enteritis (G-E) ER admissions also
analyzed as a control disease.
                                    The association between air pollution and
                                    childhood (<18 yrs.) ED visits for asthma
                                    from the inner city area with high asthma
                                    hospitalization rates (0.8/day, 23/day/10K
                                    persons) were compared with those from
                                    lower hospital utilization areas(l. I/day,
                                    8/day/10K persons). Daily ED counts were
                                    regressed against PM10, light scattering
                                    (asp), CO, SO2, andNO2 using a
                                    semiparametric Poisson regression model
                                    evaluated for over-dispersion and auto-
                                    correlation.
                                                                             Positive association observed between
                                                                             asthma visits and PM10. Strongest
                                                                             association with concurrent-day PM10
                                                                             levels. No co-pollutants considered.
                                                                             Temperature and RH did not predict visits,
                                                                             but did interact with the PM10 association.
                                                                             Morbidity relative risk higher with respect
                                                                             to PM10 pollution during warmer days.
Consistent relationships found between
asthma ER visits and PM10, with greatest
effect at lower temperatures.  Sensitivity
analyses supported these findings.  NO2
also associated, but in simultaneous
regressions only PM10 stayed associated.
ER visits for gastroenteritis not
significantly associated with air pollution.
Results demonstrate an association
between wintertime ambient PM10 and
asthma exacerbations in an area where
RWC is a principal PM source.

Associations found between ED visits for
asthma in children and fine PM and
CO.  CO and PM10 highly correlated with
each other (r=.74) and K, an indicator of
woodsmoke pollution. There was no
stronger association between ED visits for
asthma and air pollution in the higher
hospital utilization area than in the lower
utilization area in terms of RR's.
However, considering baseline risks/1 OK
population indicates a higher PM
attributable risk (AR) in the inner city.
                                        Asthma Medical Visits (all ages):
                                        For mean = 50 Mg/m3 PM10 (single poll.)
                                        Lag = 0 days
                                        ER = 20.9%(CI: 11.8,30.8)
                                                                                                                              Asthma ED Visits (all ages)
                                                                                                                              PM10 = 50 Mg/m3 (2 day lag):
                                                                                                                              At20°F,ER=34.7%(CI: 16,56.5)
                                                                                                                              At30°F,ER = 22%(CI: 11,34.2)
                                                                                                                              At 41 ° F, ER = 9.1% (CI: 2.7, 15.9)
                                                                                 Children's (<18 yrs.) Asthma ED Visits
                                                                                 Single Pollutant Models:
                                                                                 24h PM10 =50 Mg/m3
                                                                                 Lagl ER= 75.9% (25.1, 147.4)
                                                                                 For25Mg/m3PM25
                                                                                 Lagl ER = 44.5% (CI: 21.7, 71.4)

                                                                                 Multiple Pollutant Models:
                                                                                 24h PM10 =50 Mg/m3
                                                                                 Lagl ER= 75.9%(CI: 16.3, 166)
                                                                                 For25Mg/m3PM25
                                                                                 Lagl ER= 51.2% (CI: 23.4, 85.2)

-------
to
o
o
to
           TABLE 8B-3 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                            Study Description:
                                         Results and Comments
                                       PM Index, Lag, Excess Risk %,
                                       (95% CI = LCI, UCL) Co-Pollutants
oo
td
 United States (cont'd)

 Norris et al. (2000)
 Spokane, WA (1/95 - 3/97)
 Population = 300,000
 PM10 mean. = 27.9 /j,g/m3
 PM10 Mm/Max =4.7/186.4 Mg/m3
 PM10IQR = 21.4Mg/m3

 Seattle, WA (9/95 - 12/96)
 Pop. Of Children <18 = 107,816
 PM10 mean. = 21.5 ,ug/ni3
 PM10 Min/Max = 8/69.3 Mg/m3
 PM10IQR=11.7Mg/m3

 Tolbert et al. (2000b)
 Atlanta, GA (92 - 94 Summers)
 Population = 80% of children in
 total population of 3 million
 PM10 mn. (SE) = 38.9 (15.5) Mg/
 PM10 Range = 9, 105 ,ug/m3
Associations investigated between an
atmospheric stagnation index (# of hours
below median wind speed), a "surrogate
index of pollution", and asthma ED visits
for persons <65 yr. (mean=3.2/d) in
Spokane and for children <18 yr.
(mean=l .8/d) in Seattle.  Poisson GAM
model applied, controlling for day of week,
long-wave effects, and temperature and dew
point (as non-linear smooths). Factor
Analysis (FA) applied to identify PM
components associated with asthma HA's.

Pediatric (<17 yrs. of age) ED visits (mean
= 467/day) related to air pollution (PM10,
O3, NO,,, pollen and mold) using GEE and
logistic regression and Bayesian models.
Autocorrelation, day of week, long-term
trend terms, and linear temperature controls
included.
Stagnation persistence index was strongly
associated with ED visits for asthma in
both cities. Factor analysis indicated that
products of incomplete combustion
(especially wood-smoke related K, OC,
EC, and CO) are the air pollutants driving
this association. Multi-pollutant models
run with "stagnation" as the "co-pollutant"
indicated importance of general air
pollution over any single air pollutant
index, but not of the importance of various
pollutants relative to each other.

Both PM10 and O3 positively associated
with asthma ED visits using all three
modeling approaches.  In models with
both O3 and PM10, both pollutants become
non-significant because of high
collinearity of the variables (r=0.75).
                                                                                                                            Asthma ED Visits
                                                                                                                            Single Pollutant Models

                                                                                                                            Persons<65 years (Spokane)
                                                                                                                            For PM10IQR = 50 ,ug/m3
                                                                                                                            Lag 3 ER = 2.4% (CI: -10.9, 17.6)

                                                                                                                            Persons<18 years (Seattle)
                                                                                                                            For PM10 IQR = 50 Mg/m3
                                                                                                                            Lag 3 ER = 56.2% (95 CI: 10.4,121.1)
Pediatric (<17 yrs. of age) ED Visits
PM10 = 50 Mg/m3
Lag 1 day ER = 13.2% (CI: 1.2, 26.7)
With O3 8.2 (-7.1, 26.1)
fe
H
6
o
o
H
O
O
H
W
O
o

-------
 to
 o
 o
 to
            TABLE 8B-3 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                            Study Description:
                                         Results and Comments
PM Index, Lag, Excess Risk %,
(95% CI = LCI, UCL) Co-Pollutants
 oo
 td
 -U
 to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 United States (cont'd)

 Tolbert et al. (2000a)
 Atlanta
 Period 1:  1/1/93-7/31/98
 Mean, median, SD:
 PM10 Cug/m3): 30.1,28.0,12.4

 Period 2:  8/1/98-8/31/99
 Mean, median, SD:
 PM10 (,ug/m3): 29.1,27.6,12.0
 PM25 Og/m3):  19.4, 17.5, 9.35
 CP (>g/m3):  9.39, 8.95, 4.52 10-
 100 nmPM counts
 (count/cm3):  15,200, 10,900,
 26,600
 10-100 nm PM surface area
 (unrVcm3): 62.5,43.4,116
 PM25 soluble metals (,ug/m3):
 0.0327, 0.0226, 0.0306
 PM25 Sulfates Og/m3): 5.59, 4.67,
 3.6
 PM2 5 Acidity Cwg/m3):  0.0181,
 0.0112,0.0219
 PM2 5 organic PMCwg/m3): 6.30,
 5.90,3.16
 PM25 elemental carbon (,ug/m3):
 2.25,1.88,1.74
Preliminary analysis of daily emergency
department (ED) visits for asthma (493),
wheezing (786.09) COPD (491, 492, 4966)
LRI 466.1, 480, 481, 482, 483, 484, 485,
486), all resp disease (460-466, 477, 480-
486, 491, 492, 493, 496, 786.09) for
persons > 16 yr in the period before (Period
1) and during (Period 2) the Atlanta
superstation study. ED data analyzed here
from just 18 of 33 participating hospitals;
numbers of participating hospitals increased
during period 1. Mean daily ED visits for
dysrhythmias and all DVD in period 1 were
6.5 and 28.4, respectively. Covariates:
NO2, O3, SO2, CO temperature, dewpoint,
and, in period 2 only, VOCs. PM measured
by both TEOM and Federal Reference
Method; unclear which used in analyses.
For epidemiologic analyses, the two time
periods were analyzed separately.  Poisson
regression analyses were conducted with
cubic splines for time, temperature and
dewpoint.  Day-of-week and hospital
entry/exit indicators also included.
Pollutants treated a-priori as three-day
moving averages of lags 0, 1, and 2. Only
single-pollutant results reported.
                                                                                     In period 1, observed significant COPD
                                                                                     association with 3-day average PM10.
                                                                                     COPD was also positively associated with
                                                                                     NO2, O3, CO and SO2. No statistically
                                                                                     significant association observed between
                                                                                     asthma and PM10 in period 1. However,
                                                                                     asthma positively associated with ozone
                                                                                     (p=0.03). In period 2, i.e., the first year of
                                                                                     operation of the superstation, no
                                                                                     statistically significant associations
                                                                                     observed with PM10 or PM2 5. These
                                                                                     preliminary results should be interpreted
                                                                                     with caution given the incomplete and
                                                                                     variable nature of the databases analyzed.
Period 1:
PM10 (0-2 d):
  asthma:
    5.6% (-8.6, 22.1)
  COPD:
    19.9% (0.1, 43.7)

Period 2: (all 0-2 day lag)
PM10:   asthma
        18.8% (-8.7, 54.4)
        COPD
        -3.5% (29.9-  33.0)
PM25:   asthma
        2.3% (-14.8, 22.7)
        COPD
        12.4% (-7.9, 37.2)
PM10_25:  asthma
        21.1% (-18.2,  79.3)
        COPD
        -23.0% (50.7-20.1)
O

-------
to
o
o
to
            TABLE 8B-3 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                            Study Description:
                                         Results and Comments
                                        PM Index, Lag, Excess Risk %,
                                        (95% CI = LCI, UCL) Co-Pollutants
          United States (cont'd)

          Yang etal( 1997)
          Study Period: 92 - 94
          Reno-Sparks, Nevada
          Population = 298,000
          PM10 mean = 33.6 Mg/m3
          PM10 range = 2.2, 157.3
                                   Association between asthma ER visits
                                   (mean = 1.75/d, SD=1.53/d) and PM10, CO
                                   and O3 assessed using linear WLS and
                                   ARIMA regression, including adjustments
                                   for day-of-week, season, and temperature
                                   (but not RH or T-RH interaction). Season
                                   adjusted only crudely, using month dummy
                                   variable.
                                         Only O3 showed significant associations
                                         with asthma ER visits. However, the
                                         crude season adjustment and linear model
                                         (rather than Poisson) may have adversely
                                         affected results.  Also, Beta-gauge PM10
                                         mass index used, rather than direct
                                         gravimetric mass measurements.
                                        NR
oo
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 Canada

 Delfmoetal. (1997)
 Montreal, Canada
 Population= 3 million
 6-9/92, 6-9/93
 1993 Means (SD):
 PM10=21.7Mg/m3(10.2)
 PM25=12.2,ug/m3(7.1)
 SO4==  34.8nmol/m3(33.1)
 H+=   4nmol/m3(5.2)
         Delfmoetal. (1998)
         Montreal, Canada
         6-8/89,6-8/90
         MeanPM10= 18.6 ,ug/m3
         (SD=9.3, 90th% = 30.0 Mg
Association of daily respiratory emergency
department (ED) visits (mean = 98/day
from 25 of 31 acute care hospitals) with O3,
PM10, PM25, SO4~, and H+ assessed using
linear regression with controls for temporal
trends, auto-correlation, and weather.  Five
age sub-groups considered.
                                   Examined the relationship of daily ED
                                   visits for respiratory illnesses by age
                                   (mean/day: <2yr.=8.9; 2-34yr.=20.1; 35-
                                   64yr.=22.6; >64yr.=20.3) with O3 and
                                   estimated PM25.  Seasonal and day-of-week
                                   trends, auto-correlation, relative humidity
                                   and temperature were addressed in linear
                                   time series regressions.
No associations with ED visits in '92, but
33% of the PM data missing then. In'93,
only H+ associated for children <2, despite
very low H+ levels. H+ effect stable in
multiple pollutant models and after
excluding highest values.  No associations
for ED visits in persons aged 2-64 yrs.
For patients >64 yr, O3, PM10, PM2 5, and
SO4~ positively associated with visits
(p < 0.02), but PM effects smaller than for
03.

There was an association between PM2 5
and respiratory ED visits for older adults
(>64), but this was confounded by both
temperature and O3.  The fact that PM2 5
was estimated, rather than measured, may
have weakened its relationship with ED
visits, relative to O3.
                                                                                                                            Respiratory ED Visits
                                                                                                                             Adults >64: (pollutant lags = 1 day)
                                                                                                                             50 Mg/m3 PM10ER = 36.6% (10.0, 63.2)
                                                                                                                             25 Mg/m3 PM2.5 ER = 23.9% (4.9, 42.8)
                                                                                 Older Adults(>64 yr) Respiratory ED Visits
                                                                                 Estimated PM25 = 25 /-ig/m3

                                                                                 Single Pollutant:
                                                                                 (lag 1 PM25) ER = 13.2 (-0.2, 26.6)

                                                                                 With Ozone (lag 1PM2 5):
                                                                                 Est. PM25 (lagl) ER = 0.8% (CI: -14.4, 15.8)
o

-------
 to
 o
 o
 to
            TABLE 8B-3 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                             Study Description:
                                         Results and Comments
                                        PM Index, Lag, Excess Risk %,
                                        (95% CI = LCI, UCL) Co-Pollutants
 oo
 td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Canada (cont'd)

          Stiebetal. (1996)
          New Brunswick, Canada
          Population = 75,000
          May-Sept. 84 - 92

          SO42'Mean =5.5 Mg/m3
          Range: 1-23, 95th% =14 ,ug/m3
          TSP Mean = 36.7 Mg/m3
          Range:5-108, 95°% =70 Mg/m3
 Stieb et al. (2000)
 Saint John, Canada
 7/1/92-3/31/96
 mean and S.D.:
 PM10(Mg/m3):  14.0,9.0
 PM25(Aig/m3):  8.5,5.9
 H+(nmol/m3):  25.7,36.8
 Sulfate (nmol/m3):  31.1,29.7
 COHmean(103lnft): 0.2,0.2
 COHmax(103lnft): 0.6,0.5
Asthma ED visits (mean=l .6/day) related to
daily O3 and other air pollutants (SO2, NO2,
SO42", and TSP). PM measured only every
6th day. Weather variables included
temperature, humidex, dewpoint,  and RH.
ED visit frequencies were filtered to remove
day of week and long wave trends. Filtered
values were regressed on pollution and
weather variables for the same day and the
3 previous days.

Study of daily emergency department (ED)
visits for asthma (mean 3.5/day), COPD
(mean 1.3/day), resp infections (mean
6.2/day), and all respiratory conditions
(mean 10.9/day) for persons of all ages.
Covariates included CO, H2S, NO2, O3,
SO2, total reduced sulfur (TRS), a large
number of weather variables, and 12 molds
and pollens.  Stats:  generalized additive
models with LOESS prefiltering of both ED
and pollutant variables, with variable
window lengths. Also controlled for day of
week and LOESS-smoothed functions of
weather. Single-day, and five day average,
pollution lags tested out to lag 10. The
strongest lag, either positive or negative,
was chosen for final models.  Both single
and multi-pollutant models reported. Full-
year and May-Sep models reported.
Positive, statistically significant (p < 0.05)
association observed between O3 and
asthma ED visits 2 days later; strength of
the association greater in nonlinear
models. Ozone effect not significantly
influenced by addition of other pollutants.
However, given limited number of
sampling days for sulfate and TSP, it was
concluded that "a particulate effect could
not be ruled out".

In single-pollutant models, significant
positive associations were observed
between all respiratory ED visits and
PM10, PM25, H2S, O3, and SO2.
Significant negative associations were
observed with H+, and COH max.  PM
results were similar when data were
restricted to May-Sep. In multi-pollutant
models, no PM metrics significantly
associated with all cardiac ED visits in full
year analyses, whereas both O3 and SO2
were.  In the May-Sep subset, significant
negative association found for sulfate. No
quantitative results presented for non-
significant variables in these multi-
pollutant regressions.
                                                                                                                    Emergency Department Visits (all ages)
                                                                                                                    Single Pollutant Model
                                                                                                                    100 Mg/m3 TSP = 10.7% (-66.4, 87.8)
                                                                                                                             PM2.5, (lag 3) 15.1 (-0.2, 32.8)

                                                                                                                             PM10, (lag 3) 32.5 (10.2, 59.3)
o

-------
to
o
o
to
            TABLE 8B-3 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                            Study Description:
Results and Comments
PM Index, Lag, Excess Risk %,
(95% CI = LCI, UCL) Co-Pollutants
oo
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
         Europe

         Atkinson et al. (1999a)
         London (92 - 94)
         Population = NR
         PM10Mean = 28.5,ug/m3
         10th-90thIQR = 15.8-46.5 ,ug/m3
         BSmean=12.7Mg/m3
         10th-90thIQR = 5.5-21.6 Mg/m3
         Hajatetal. (1999)
         London, England (92 - 94)
         Population = 282,000
         PM10 mean = 28.2 /j,g/m3
         PM1010'-90th%=16.3-46.4 Aig/
         BSmean=10.lMg/m3
         BS 10'-90th%=4.5-15.9Mg/m3
                                   All-age Respiratory (mean=90/day),
                                   Asthma (25.9/day), and Other Respiratory
                                   (64.1/day) ED visits from 12 London
                                   hospitals considered, but associated
                                   population size not reported.  Counts for
                                   ages 0-14, 15-64, and>64 also examined.
                                   Poisson regression used, controlling for
                                   season, day of week, meteorology,
                                   autocorrelation, overdispersion, and
                                   influenza epidemics.
                                    Examined associations of PM10, BS, NO2,
                                    O3, SO2, and CO, with primary care general
                                    practitioner asthma and "other LRD"
                                    consultations.  Asthma consultation means
                                    per day = 35.3 (all ages); 14.(0-14 yrs,);
                                    17.7 (15-64 yrs.); 3.6 (>64 yrs.).  LRD
                                    means = 155 (all ages);  39.7(0-14 yrs,);
                                    73.8 (15-64 yrs.); 41.1 (>64 yrs.). Time-
                                    series analyses of daily numbers of
                                    consultations performed, controlling for
                                    time trends, season factors, day of week,
                                    influenza, weather, pollen levels, and serial
                                    correlation.
PM10 positively associated, but not BS, for
all-age/all-respiratory category. PM10
results driven by significant children and
young adult associations, while older adult
visits had negative (but non-significant)
PM10-ED visit relationship. PM10
positively associated for all ages, children,
and young adults for asthma ED visits.
However, PM10-asthma relationship
couldn't be separated from SO2 in multi-
pollutant regressions. Older adult ED
visits most strongly associated with CO.
No O3-ED visits relationships found (but
no warm season analyses attempted).
Positive associations, weakly significant
and consistent across lags, observed
between asthma consultations and NO2
and CO in children, and with PM10 in
adults, and between other LRD
consultations and SO2 in children.
Authors concluded that there are
associations between air pollution and
daily concentrations for asthma and other
lower respiratory disease in London. In
adults, the authors concluded that the only
consistent association was with PM10.
Across all of the various age, cause, and
season categories considered, PM10 was
the pollutant most coherent in giving
positive pollutant RR estimates for both
asthma and other LRD (11 of 12
categories positive) in single pollutant
models considered.
PM10 (50 Mg/m3) No co-pollutant:
All Respiratory ED visits
All age(lag ld)ER = 4.9% (CI: 1.3, 8.6)
<15yrs(lag 2d)ER = 6.4% (CI: 1, 12.2)
15-64yr(lagld)ER= 8.6% (CI: 3.4, 14)
Asthma ED visits
All age (lag Id) ER = 8.9% (CI: 3, 15.2)
<15yrs (lag 2d) ER = 12.3% (CI: 3.4, 22)
15-64yr(lg ld)ER= 13% (CI: 4.6,22.1)

PM10 (50 Mg/m3) 2d lag & co-pollutant:
Children's (<15 yrs.) Asthma ED Visits:
PM alone: ER = 12.3% (CI: 3.4, 22)
&N02:      ER= 7.8% (CI: -1.2, 17.6)
&O3:    ER =10.5% (CI: 1.6,20.1)
&S02:   ER = 8.1%(CI: -1.1,18.2)
&CO:   ER = 12.1%(CI: 3.2,21.7)

Asthma Doctor's Visits:
50 Mg/m3 PM10
-Year-round, Single Pollutant:
All ages (Ig 2): ER = 5.4% (CI: -0.6, 11.7)
0-14 yrs.(lg 1): ER = 6.4% (-1.5, 14.6)
15-64 yrs.(lg 0): ER = 9.2% (CI: 2.8, 15.9)
>64yrs.(lg 2): ER= 11.7% (-1.8, 26.9)
-Year-round, 2 Pollutant, Children (0, 14):
(PM10 lag = 1 day)  PM10 ER's:
W/N02: ER= 0.8% (CI: -8.7, 11.4)
W/O3: ER= 5.5% (-2.1,  13.8)
W/S02: ER= 3.2% (CI:  -6.4, 13.7)
Other Lower Resp. Pis. Doctor's Visits:
50 Mg/m3 PM10
-Year-round, Single Pollutant:
All ages (Ig 2): ER = 3.5% (CI: 0,  7.1)
0-14yrs.(lg 1): ER = 4.2%(CI: -1.2, 9.9)
15-64 yrs.(lg 2): ER= 3.7% (CI: 0.0, 7.6)
>64yrs.(lg 2): ER = 6.2% (CI: 0.5, 12.9)

-------
to
o
o
to
            TABLE 8B-3 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY HOSPITAL
	ADMISSIONS  STUDIES	

 Reference/Citation
 Location, Duration
 PM Index/Concentrations
                                            Study Description:
                                         Results and Comments
                                        PM Index, Lag, Excess Risk %,
                                        (95% CI = LCI, UCL) Co-Pollutants
oo
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
         Europe (cont'd)

         Hajatetal. (2001)
         London (1992-1994)
         44,406-49,596 registered patients
         <1 to 14 years
         PM10mean28.5(13.9)
 Medina etal. (1997)
 Greater Paris 91 -95
 Populations 6.5 MM
 Mean PM13 = 25 Mg/m
 PM13 min/max = 6/95 /-
         BS min/max = 3/130 ,ug/m3
         Damiaetal. (1999)
         Valencia, Spain (3/94-3/95)
         Population = NR
         BS mean = 101 Mg/m3
         BS range = 34-213 ,ug/m3
                                    Daily physician consultations (mean daily
                                    4.8 for children; 15.3 for adults) for allergic
                                    rhinitis (ICD-9, 477), SO2, O3, NO2, CO,
                                    PM10, and pollen using generalized additive
                                    models with nonparametric smoother.
Evaluated short-term relationships between
PM13 and BS concentrations and doctors'
house calls (mean=8/day; 20% of city total)
in Greater Paris. Poisson regression used,
with non-parametric smoothing functions
controlling for time trend, seasonal patterns,
pollen counts, influenza epidemics,  day-of-
week, holidays, and weather.
                                   Associations of BS and SO2 with weekly
                                   total ED admissions for asthma patients
                                   aged > 12 yrs (mean = 10/week) at one
                                   hospital over one year assessed, using linear
                                   stepwise regression. Season-specific
                                   analyses done for each of 4 seasons, but no
                                   other long-wave controls. Linear T, RH,
                                   BP, rain, and wind speed included as crude
                                   weather controls in ANOVA models.
SO2 and O3 show strong associations with
the number of consultations for allergic
rhinitis. Estimates largest for a lag of 3 or
4 days prior to consultations, with
cumulative measures stronger than single
day lags. Stronger effects were found for
children than adults. The two-pollutant
analysis of the children's model showed
that PM10 and NO2 associations
disappeared once either SO2 or O3 was
incorporated into the model.

A relationship between all age (0-64 yrs.)
asthma house calls and PM13, BS, SO2,
NO2, and O3 air pollution, especially for
children aged 0-14 (mean = 2/day).
In two-pollutant models including BS
with, successively, SO2, NO2, and O3, only
BS and O3 effects remained stable.  These
results also indicate that air pollutant
associations noted for hospital ED visits
are also applicable to a wider population
that visits their doctor.

Both BS and SO2 correlated with ED
admissions for asthma (SO2: r=0.32; BS:
i=0.35), but only BS significant in
stepwise multiple regression. No linear
relationship found with weather variables.
Stratified ANOVA found strongest BS-ED
association in the autumn and during
above average temperatures.  Uncontrolled
autocorrelation (e.g., within-season) and
weather effects likely remain in models.
                                                                                 PM10 - Increment (10-90%)
                                                                                 (15.8-46.5)
                                                                                 Age <1-14 years
                                                                                 lag 3:  10.4 (2.0 to 19.4)
                                                                                 Cum 0-3:  17.4 (6.8 to 29.0)

                                                                                 Ages 15-64 years
                                                                                 lag 2:  7.1(2.6 to 11.7)
                                                                                 Cum 0-6: 20.2 (14.1 to 26.6)
                                                                                                                              Doctor's Asthma House Visits:
                                                                                                                              50 Mg/m3 PM13
                                                                                                                              Year-round, Single Pollutant:
                                                                                                                              All ages (Ig 2): ER = 12.7% (CI: 4.1, 21.9)
                                                                                                                              0-14 yrs.(lg 0-3): ER = 41.5% (CI: 20, 66.8)
                                                                                                                              15-64 yrs.(lg 2): ER = 6.3% (CI: -4.6, 18.5)
                                                                                 Asthma ED Visits (all ages):
                                                                                 BS = 40 Mg/m3 (single pollutant)
                                                                                 BS as a lag 0 weekly average:
                                                                                 ER = 41.5% (CI = 39.1, 43.9)
O

-------
                TABLE 8B-3 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY MEDICAL VISITS
 to
 o
 o
 to
Reference/Citation
Location, Duration
PM Index/Concentrations
Study Description:
Results and Comments
PM Index, Lag, Excess Risk %,
(95% CI = LCI, UCL) Co-Pollutants
 oo
 td
          Europe (cont'd)

          Pantazopoulou et al. (1995)
          Athens, GR( 1988)
          Population = NR
          Winter (1/88-3/88,9/88-12/88)
          BS mean. =75 /-ig/m3
          BS 5th-95th%=26 - 161 Mg/m3
          Summer (3/22/88-3/88,9/21/88)
          BS mean. =55 /-ig/m3
          BS5th-95th%=19-90Mg/m3
          Garty etal. (1998)
          PM10 mean ~ 45 Mg/ni3
          Tel Aviv, Israel (1993)
                                   Examined effects of air pollution on daily
                                   emergency outpatient visits and admissions
                                   for cardiac and respiratory causes. Air
                                   pollutants included: BS, CO, andNO2.
                                   Multiple linear regression models used,
                                   controlling for linear effects of temperature
                                   and RH, day of week, holidays, and dummy
                                   variables for month to crudely control for
                                   season, separately for winter and summer.
                                   Seven day running mean of asthma ED
                                   visits by children (1-18 yrs.) to a pediatric
                                   hospital modeled in relation to PM10 in Tel
                                   Aviv, Israel.
                                         Daily number of emergency visits related
                                         positively with each air pollutant, but only
                                         reached nominal level of statistical
                                         significance for NO2 in winter.  However,
                                         the very limited time for each within-
                                         season analysis (6 mo.) undoubtably
                                         limited the power of this analysis to detect
                                         significant effects.  Also, possible lagged
                                         pollution effects were apparently not
                                         investigated, which may have reduced
                                         effect estimates.

                                         No PM10 associations  found with ED
                                         visits. The ER visits-pollutant correlation
                                         increased significantly when the
                                         September peak was excluded.  Use of a
                                         week-long average and associated
                                         uncontrolled long-wave fluctuations (with
                                         resultant autocorrelation) likely prevented
                                         meaningful analyses of short -term PM
                                         associations with ED visits.
                                        Single Pollutant Models
                                        For Winter (BS = 25 Mg/m3
                                        Outpatient Hospital Visits
                                        ER= 1.1% (-0.7, 2.3)
                                        Respiratory HA's
                                        ER = 4.3%(0.2, 8.3)
                                        For Summer, BS = 25 Mg/
                                        Outpatient Hospital Visits
                                        ER= 0.6% (-4.7, 6.0))
                                        Respiratory HA's
                                        ER= 5.5% (-3.6, 14.7)

                                        N/A
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
O

-------
                TABLE 8B-3 (cont'd). ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY MEDICAL VISITS
to
o
o
to
Reference/Citation
Location, Duration
PM Index/Concentrations
Study Description:
Results and Comments
PM Index, Lag, Excess Risk %,
(95% CI = LCI, UCL) Co-Pollutants
oo
td
-U
oo
          Latin Amercia

          Ilabacaetal. (1999)
          Santiago, Chile
          February 1995-August 1996
          PM10:  warm: 80.3 Mg/ni3
                cold:  123.9,ug/m3
          PM25: warm: 34.3 Mg/rn3
                cold:  71.3 Mg/m3
                                 Number of daily respiratory emergency
                                 visits (REVs) related to PM by Poisson
                                 model smooths for longer- and short-term
                                 trends. SO2,NO2,O3.
                                        Stronger coefficients for models including
                                        PM2 5 than for models including PM10 or
                                        PM10_2 5. Copollutant effects were
                                        significantly associated with REVs.  For
                                        respiratory patients, the median number of
                                        days between the onset of the first
                                        symptoms and REV was two to three days.
                                        For the majority of patients (70%) this
                                        corresponded to the lag observed in this
                                        study indicating that the timing of the
                                        pollutant effect is consistent with the
                                        temporal pattern of REV in this
                                        population.
                                       REV, lag 2
                                       Cold
                                       PM25,lag2
                                       OR: 1.027 (1.01 to 1.04) for a45
                                       increment

                                       PM10, lag 2
                                       OR: 1.02 (1.01 to 1.04) for a 76 M
                                       increment

                                       PM25,lag2
                                       OR: 1.01 (1.00* to 1.03) for a 32
                                       increment

                                       Pneumonia, lag 2
                                       PM10:  1.05 (1.00* to 1.10)
                                       64 ,ug/rn3 increment
                                       PM25:  1.04 (1.00* to 1.09)
                                       45 Mg/nr3 increment
                                       PM10.25:  10.5 (1.00* to 1.10)
                                       32 ,ug/rn3 increment

                                       'decimals < 1.00
fe
H
6
o
o
H
O
O
H
W
O
O

-------
                TABLE 8B-3 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY MEDICAL VISITS
 to
 o
 o
 to
Reference/Citation
Location, Duration
PM Index/Concentrations
Study Description:
Results and Comments
PM Index, Lag, Excess Risk %,
(95% CI = LCI, UCL) Co-Pollutants
 oo
 td
 -U
 VO
 fe
 H
 6
 o

 o
 H
O
 O
 H
 W
 O
          Latin Amercia (cont'd)

          Lin etal. (1999)
          Sao Paulo, BR (91-93)
          Population=NR
          PM10 mean =65 Mg/m3
          PM10 SD=27 Mg/m3
          PM10 range=15-193 Mg/
Ostroetal. (1999b)
Santiago, CI (7/92—12/93)
<2 yrs. Population « 20,800
3-14 yrs. Population ~ 128,000
PM10 mean. =108.6 Mg/m3
PM10 Min/Max=18.5/380 Mg/m3
PM10IQR = 70.3 - 135.5 Mg/m3
                                  Respiratory ED visits by children (0-12
                                  yrs.) To a major pediatric hospital
                                  (mean=56/day) related to PM10, SO2, NO2,
                                  CO, and O3 using Gaussian linear
                                  regression modeling, Poisson modeling, and
                                  a polynomial distributed lag model. Lower
                                  respiratory (mean = 8/day) and upper
                                  respiratory (mean = 9/day) all evaluated.
                                  Analyses considered effects of season, day
                                  of week, and extreme weather (using T, RH
                                  dummy variables).
Analysis of daily visits to primary health
care clinics for upper (URS) or lower
respiratory symptoms (LRS) for children 2-
14 yr (mean LRS=111. I/day) and < age 2
(mean LRS=104.3/day). Daily PM10 and O3
and meteorological variables considered.
The multiple regression GAM included
controls for seasonality (LOESS smooth),
temperature, day of week, and month.
PM10 was found to be "the pollutant that
exhibited the most robust and stable
association with all categories of
respiratory disease".  O3 was the only
other pollutant that remained associated
when other pollutants all simultaneously
added to the model. However, some
pollutant coefficients went negative in
multiple pollutant regressions, suggesting
coefficient intercorrelations in the multiple
pollutant models.  More than 20%
increase in ED visits found on the most
polluted days, "indicating that air
pollution is a substantial pediatric health
concern".
Analyses indicated an association between   Lower Resp. Symptoms Clinic Visits
                                                                                 50 Mg/nf PM10 (0-5-day lag mean)
                                                                                 Respiratory ED Visits (<13 yrs.)
                                                                                 Single pollutant model:
                                                                                 PM10ER=21.7%(CI: 18.2,25.2)
                                                                                 All pollutant models: PM10 ER=28.8%
                                                                                 (CI: 21.4,36.7)
                                                                                 Lower Respiratory ED Visits (<13 yrs.)
                                                                                 Single pollutant model:
                                                                                 PM10 ER=22.8% (CI: 12.7, 33.9)
                                                                                 All pollutant models: PM10 ER=46.9%
                                                                                 (CI: 27.9,68.8)
PM10 and medical visits for LRS in
children ages 2-14 and in children under
age 2 yr.  PM10 was not related to non-
respiratory visits (mean =208/day).
Results unchanged by eliminating high
PM10 (>235 ,ug/m3) or coldest days
(<8°C).  Adding O3 to the model had little
effect on PM,n-LRS associations.
PM10 = 50 Mg/m3
   Single Pollutant Models:
-Children<2 years
Lag3ER = 2.5%(CI:0.2,4.8)
-Children 2-14 years
Lag 3 ER = 3.7% (CI: 0.8, 6.7%)
   Two Pollutant Models (with O3):
-Children<2 years
Lag 3 ER = 2.2% (CI: 0, 4.4)
-Children 2-14 years
Lag 3 ER = 3.7% (CI: 0.9, 6.5)
O

-------
                TABLE 8B-3 (cont'd).  ACUTE PARTICIPATE MATTER EXPOSURE AND RESPIRATORY MEDICAL VISITS
to
o
o
to
Reference/Citation
Location, Duration
PM Index/Concentrations
Study Description:
Results and Comments
                                                                                                                              PM Index, Lag, Excess Risk %,
                                                                                                                              (95% CI = LCI, UCL) Co-Pollutants
         Australia
         Smith etal. (1996)
         StdyPd.:  12/92-1/93,12/93-1/94
         West Sydney, AU
         Population = 907,000
         -Period 1 (12/92-1/93)
         Bscalt median = 0.25 10~4/m
Bsc
                  = 0.18-0.39 10-4/m
         Bscatt95th% = 0.86 10-4/m
         -Period 2 (12/93-1/94)
         Bscalt median = 0.19 10~4/m
Bsc
                  = 0.1-0.38 10-4/m
         Bscatt 95th5% = 3.26 10-4/mPM10
         median =18 Mg/m3
         PM10 IQR =11.5-28.8 Mg/m3
         PM1095th% = 92.5Mg/m3
Study evaluated whether asthma visits to
emergency departments (ED) in western
Sydney (mean-10/day) increased as result
of bushfire-generated PM ( Bscalt from
nephelometry) in Jan., 1994 (period 2). Air
pollution data included nephelometry
(Bscatt), PMio, SO2, andNO2. Data analyzed
using two methods: (1) calculation of the
difference in proportion of all asthma ED
visits between the time periods, and;  (2)
Poisson regression analyses. Control
variables included T, RH, BP, WS, and
rainfall.
oo
td
No difference found in the proportion of
all asthma ED visits during a week of
bushfire-generated air pollution, compared
with the same week 12 months before,
after adjusting for baseline changes over
the 12-month period.  The max. Bscatt
reading was not a significant predictor of
the daily asthma ED visits in Poisson
regressions.  However, no long-wave
controls applied, other than indep. vars.,
and the power to detect differences was
weak (90% for a 50% difference). Thus,
the lack of a  difference may be due to low
statistical strength or to lower toxicity of
particles from burning vegetation at
ambient conditions vs. fossil fuel
combustion.
                                                                                                                    ED Asthma Visits (all ages)
                                                                                                                    Percent change between bushfire and non
                                                                                                                    bushfire weeks:
                                                                                                                    PM10 = 50 Mg/m3
                                                                                                                    ER = 2.1%(CI: -0.2,4.5)
         Asia
         Ye etal. (2001)
         Tokyo, Japan
         Summer months
         July-August, 1980-1995
         PM10 46.0 mean
                                   Hospital emergency transports for
                                   respiratory disease for >65 years of age
                                   were related to pollutant levels NO2, O3,
                                   PM10, S02, and CO.
                                         For chronic bronchitis PM10  with a lag
                                         time of 2 days was the most statistically
                                         significant model covariate.
                                        Asthma (ICD-9-493)
                                        Coefficienct estimate (SE)
                                        0.003 (0.001)
 fe
 H
 6
 o

 o
 H
O
 O
 H
 W
 O
          Chew etal. (1999)
          Singapore (90 - 94)
          Population = NR
          TSPmean = 51.2,ug/m3
          TSPSD = 20.3,ug/m3
          TSP range = 13-184,ug/m3
                                   Child (3-13 yrs.) ED visits (mean =
                                   12.8/day) and HA's (mean = 12.2/day) for
                                   asthma related to levels of SO2, NO2, TSP,
                                   and O3 using linear regression with
                                   weather, day-of-week controls. Auto-
                                   correlation effects controlled by including
                                   prior day response variable as a regression
                                   variable.  Separate analyses done for
                                   adolescents (13-21 yrs.) (mean ED=12.2,
                                   meanHA=3.0/day).
                                         Positive associations found between TSP,
                                         SO2, and NO2, and daily HA and ED visits
                                         for asthma in children, but only with ED
                                         visits among adolescents. Lack of power
                                         (low counts) for adolescents' HA's
                                         appears to have been a factor in the  lack of
                                         associations. When ED visits stratified by
                                         year, SO2 and TSP remained associated in
                                         every year, but not NO2. Analyses for
                                         control diseases (appendicitis and urinary
                                         tract infections) found no associations.
                                        TSP(100 Mg/m3)No co-pollutant:

                                        Child (3-13 vrs.Wsthma ED visits
                                        Lag ldER = 541% (CI: 198.4, 1276.S
O

-------
              Appendix 8B.4: Pulmonary Function Studies
April 2002                        8B-51      DRAFT-DO NOT QUOTE OR CITE

-------
to
o
o
to
 TABLE 8B-4.  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON PULMONARY FUNCTION
	TESTS IN STUDIES OF ASTHMATICS	

                                                                                                                              Effect measures standardized to 50 ,ug/m3
                                                                                                                              PM10 (25 ,ug/m3 PM2.5). Negative coefficients
                                                                                                                              for lung function and ORs greater than 1 for
                                                                                                                              other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
Results and Comments
Effects of co-pollutants
OO
td
(^
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          UnitedStates

          Thurstonetal. (1997)
          Summers 1991-1993.
          O3, H+, sulfate
          Canada

          Vedaletal. (1998)
          Port Alberni, BC
          PM10 measurements were made using a Sierra-Anderson
          dichotomous sampler. PM10 ranged from 1 to
          159,ug/m3.
          Europe

          Gielenetal. (1997)
          Amsterdam, NL
          Mean PM10 level: 30.5 ,ug/m3 (16, 60.3).
          Mean maximum 8 hr O,. 67 i/g/m3.
          Hiltermannetal. (1998)
          Leiden, NL
          July-Oct, 1995
          O3, NO2, SO2, BS, and PM10 ranged from 16.4 to
          97.9 Mg/m3)
                                            Three 5-day summer camps conducted in 1991,
                                            1992, 1993. Study measured symptoms and
                                            change in lung function (morning to evening).
                                            Poisson regression for symptoms.
                                            Study of 206 children aged 6 to 13 years living
                                            in Port Alberni, British Columbia. 75 children
                                            had physician-diagnosed asthma, 57 had an
                                            exercised induced fall in FEV1, 18 children
                                            with airway obstruction, and 56 children
                                            without any symptoms.  Respiratory symptom
                                            data obtained from diaries. An autoregressive
                                            model was fitted to the data, using GEE
                                            methods.  Covariates included temp., humidity,
                                            and precipitation.
                                            Study evaluated 61 children aged 7 to 13 years
                                            living in Amsterdam, The Netherlands.  77
                                            percent of the children were taking asthma
                                            medication and the others were being
                                            hospitalized for respiratory problems. Peak
                                            flow measurements were taken twice daily.
                                            Associations of air pollution were evaluated
                                            using time series analyses. The analyses
                                            adjusted for pollen counts, time trend, and day
                                            of week.

                                            270 adult asthmatic patients from an out-patient
                                            clinic in Leiden, The Netherlands were studied
                                            from July 3 to October 6,  1995.  Peak flow
                                            measured twice daily. An autoregressive model
                                            was fitted to the data. Covariates included
                                            temp, and day of week. Individual responses
                                            not modeled.
                                            The O3-APEFR relationship was seen as
                                            the strongest.
                                            In general, PM10 was associated with
                                            changes in both peak flow and respiratory
                                            symptoms. Ozone, SO2, and sulfate levels
                                            were low because of low vehicle
                                            admissions.
                                            The strongest relationships were found
                                            with ozone, although some significant
                                            relationships found with PM10.
                                            No relationship between ozone or PM10
                                            and PET was found
                                       Lag 0, PM10 average PEE- -0.27 (-0.54,
                                       -0.01) per 10 ,ug/m3 increment
                                       Lag 0, PM10:
                                        Evening PEE = -0.08 (-2.49, 2.42)
                                       Lag 1, PM10:
                                        Morning PEE = 1.38 (-0.58, 3.35)
                                       Lag 2, PM10:
                                        Morning PEE = 0.34 (-1.78, 2.46)
                                        Evening PEE = -1.46 (-3.23, 0.32)
                                       Lag 0, PM10:
                                        Average PEE = -0.80 (-3.84, 2.04)
                                       7 day ave., PM10:
                                        Average PEE = -1.10 (-5.22, 3.02)

-------
to
o
o
to
          TABLE 8B-4 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON PULMONARY
         	FUNCTION TESTS IN STUDIES OF ASTHMATICS	

                                                                                                                                      Effect measures standardized to 50 ,ug/m3
                                                                                                                                      PM10 (25 ,ug/m3 PM2.5). Negative coefficients
                                                                                                                                      for lung function and ORs greater than 1 for
                                                                                                                                      other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
                                                                                                          Results and Comments
                                                                                                          Effects of co-pollutants
          Europe (cont'd)
oo
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Peters etal. (1996)
          Erfurt and Weimar, Germany
          SO2, TSP, PM10, sulfate fraction, and PSA.
          Mean PM10 level was 112 ,ug/m3.
          PM was measured by a Marple-Harvard impactor.
Peters etal. (1997b)
Erfurt, Germany
PM fractions measured over range of sizes from
ultrafme to fine, including PM10.
Particles measured using size cuts of 0.01 to 0.1, 0.1 to
0.5, and 0.5 to 2.5 //m.
Mean PM10 level:  55 ,ug/m3 (max 71).  Mean SO2:
100,ug/m3(max383).
PM was measured using a Harvard impactor. Particle
size distributions were estimated using a conduction
particle counter.
          Peters etal. (1997c)
          Sckolov, Czech Republic
          Winter 1991-1992
          PM10, SO2, TSP, sulfate, and particle strong acid.
          Median PM10 level: 47 ^g/m3 (29, 73).
          Median SO2: 46 ^g/m3 (22, 88).
          PM was measured using a Harvard impactor.  Particle
          size distributions were estimated using a conduction
          particle counter.
Panel of 155 asthmatic children in the cities of
Erfurt and Weimar, E. Germany studied.  Each
panelist's mean PEE over the entire period
subtracted from the PEE value to obtain a
deviation. Mean deviation for all panelists on
given day was analyzed using an autoregressive
moving average. Regression analyses done
separately for adults and children in each city
and winter; then combined results calculated.

Study of 27 non-smoking adult asthmatics
living in Erfurt, Germany during winter season
of 1991-1992. Morning and  evening peak flow
readings recorded. An auto-regressive model
was used to analyze deviations in individual
peak flow values, including terms for time
trend, temp., humidity, and wind speed and
direction.
                                                                                               Five day average SO2 was associated with
                                                                                               decreased PEE. Changes in PEE were not
                                                                                               associated with PM levels.
                                                   89 children with asthma in Sokolov, Czech
                                                   Republic studied.  Subjects kept diaries and
                                                   measured peak flow for seven months during
                                                   winter of 1991-2.  The analysis used linear
                                                   regression for PET. First order autocorrelations
                                                   were observed and corrected for using
                                                   polynomial distributed lag (PDL) structures.
                                                                                                          Strongest effects on peak flow found with
                                                                                                          ultrafme particles. The two smallest
                                                                                                          fractions, 0.01 to 0.1 and 0.1 to 0.5 were
                                                                                                          associated with a decrease of PEF.
                                            Five day mean SO2, sulfates, and particle
                                            strong acidity were also associated with
                                            decreases in PM PFT as well as PM10.
Lag 0, PM10:
 Evening PEF =-0.38 (-1.83, 1.08)
Lag 1, PM10:
 Morning PEF = -1.30 (-2.36, 0.24)
5 Day Mean, PM10:
 Morning PEF =-1.51 (-3.20,0.19)
 Evening PEF = -2.31 (-4.54, -0.08)
LagO, PM25:
 Evening PEF = -0.75 (-1.66, 0.17)
Lagl, PM25:
 Morning PEF =-0.71 (-1.30,0.12)
5 Day Mean, PM25:
  Morning PEF =-1.19 (-1.81, 0.57)
  Evening PEF = -1.79 (-2.64,  -0.95)

Lag 0, PM10:
 Morning PEF = -0.71 (-2.14, 0.70)
 Evening PEF = -0.92 (-1.96, 0.12)
5 Day mean PM10:
 Evening PEF = -1.72 (-3.64, 0.19)
 Morning PEF = -0.94 (-2.76, 0.91
o

-------
to
o
o
to
 TABLE  8B-4 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON PULMONARY
	FUNCTION TESTS IN STUDIES OF ASTHMATICS	
                                                                                                                             Effect measures standardized to 50 ,ug/m3
                                                                                                                             PM10 (25 fj.g/m1 PM2.5).  Negative coefficients
                                                                                                                             for lung function and ORs greater than 1 for
                                                                                                                             other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
Results and Comments
Effects of co-pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          Europe (cont'd)

          Timonen and Pekkanen (1997)
          Kupio, Finland
          PM10, BS, NO2, and SO2.
          The intequartile range on PM10 was 8 to '<
          Penttinen et al. (2001) studied adult asthmatics for
          6 months in Helsinki, Findland. PM was measured
          using a single-stage Harvard impactor. Particle number
          concentrations were measured using an Electric Aerosol
          Spectrometer. NO2PM10 ranged from 3.8 to 73.7 //g/m3.
          PM25 ranged from 2.4 to 38.3 ,ug/m3.
          Pekkanen etal. (1997)
          Kuopio, Finland
          PM fractions measured over range of sizes from
          ultrafme to fine, including PM10.
          Mean PM10 level: 18 ^g/m3 (10, 23).
          Mean NO2 level:  28,ug/m3.
          Segalaetal. (1998)
          Paris, France
          Nov. 1992-May 1993.
          BS, SO2, NO2, PM13 (instead of PM10), measured.
          Mean PM13 level: 34.2 ,ug/m3 (range 8.8, 95).
          Mean SO2 level: 21.7 ^g/m3 (range 4.4, 83.8).
          Mean NO2 level:  56.9 ^g/m3 (range 23.8, 121.9).
          PM was measured by p-radiometry.

          Gauvinetal. (1999)
          Grenoble, France
          Summer 1996, Winter 1997
          Mean (SD) ^g/m3
          PM10 Summer 23 (6.7)
          PM10 Winter 38(17.3)
          Sunday  15.55(5.12)
          Weekday 24.03 (7.2)
                                         Studied 74 asthmatic children (7 to 12 yr) in
                                         Kuoio, Finland.  Daily mean PEF deviation
                                         calculated for each child. Values were
                                         analyzed, then using linear first-order
                                         autoregressive model. PM was measured using
                                         single stage Harvard Impactors.

                                         57 asthmatics were followed with daily PEF
                                         measurements and symptom and medications
                                         diaries from November 1996 to April 1997.
                                         PEF deviations from averages were used as
                                         dependent variables. Independent variables
                                         included PM[, PM2 5, PM10, particle counts, CO,
                                         NO, and

                                         Studied 39 asthmatic children aged 7-12 years
                                         living in Kuopio, Finland. Changes in peak flow
                                         measurements were analyzed using a linear
                                         first-order autoregressive model.  PM was
                                         measured using single stage Harvard impactors.
                                         Study of 43 mildly asthmatic children aged 7-15
                                         years living in Paris, France from Nov. 15, 1992
                                         to May 9, 1993. Peak flow measured three
                                         times a day.  Covariates in the model included
                                         temperature and humidity. An autoregressive
                                         model was fitted to the data using GEE
                                         methods.
                                         Two panels:  mild adult asthmatics, ages 20-60
                                         years, (summer-18 asthmatics, 20 control
                                         subjects; winter-19 asthmatics, 21 control
                                         subjects) were examined daily for FEV! and
                                         PEF.  Bronchial reactivity was compared
                                         Sunday vs. weekday. Temperature and RH
                                         controlled.
                                            Lagged concentrations of NO2 related to
                                            declines in morning PEF as well as PM10
                                            and BS.
                                            The strongest relationships were found
                                            between PEF deviations and PM particles
                                            below 0.1 ,um.  No associations were
                                            found between particulate pollution and
                                            respiratory symptoms.
                                            Changes in peak flow found to be related
                                            to all measures of PM, after adjusting for
                                            minimum temperature.  PNO.032-0.10
                                            (I/cm3) and PN1.0-3.2 (I/cm3) were most
                                            strongly associated with morning PEF
                                            deviations.
                                            Effects found related to PM10 were less
                                            than those found related to the other
                                            pollutants. The strongest effects were
                                            found with SO,.
                                            Respiratory function decreased among
                                            asthmatic subjects a few days (lag
                                            2/4 days) after daily PM10 levels had
                                            increased.  Bronchial reactivity was not
                                            significantly different between the
                                            weekdays and weekends.  No copollutant
                                            analysis conducted.
                                       AM PEF = -.115 (-.448, .218) PM25 lag one
                                       day
                                       AM PEF = -.001 (-.334, .332) PM25 lag two
                                       days
                                       Lag 0, PM10:
                                        Evening PEF = -0.35 (-1.14, 0.96)
                                       Lag 1, PM10:
                                        Morning PEF = -2.70 (-6.65, 1.23)
                                       Lag 2, PM10:
                                        Morning PEF = -4.35 (-8.02, -0.67)
                                        Evening PEF = -1.10 (-4.70, 2.50)

                                       Small sized particles had relationships similar
                                       to those of PM10 for morning and evening
                                       PEF.

                                       Lag 4, PM13:
                                        Morning PEF = -0.62 (-1.52, 0.28)
                                       For a 10 ,ug/m3 increase in PM10
                                       Summer
                                       FEVj
                                         -1.25%(-0.58to-1.92)
                                       PEF
                                         -0.87% (-0.1 to-1.63)

-------
to
o
o
to
 TABLE 8B-4 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON PULMONARY
	FUNCTION TESTS IN STUDIES OF ASTHMATICS	

                                                                                                                          Effect measures standardized to 50 ,ug/m3
                                                                                                                          PM10 (25 ,ug/m3 PM2.5). Negative coefficients
                                                                                                                          for lung function and ORs greater than 1 for
                                                                                                                          other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
                                                  Type of study, sample size, health outcomes
                                                  measured, analysis design, covariates included,
                                                  analysis problems, etc.
Results and Comments
Effects of co-pollutants
OO
td
 fe
 H
 6
 o

 o
 H
O
 O
 H
 W
 O
 O
          Europe (cont'd)

          Agocs etal. (1997)
          Budapest, Hungary
          SO2 and TSP were measured. TSP was measured by
          beta reactive absorption methods.
          Australia

          Rutherford etal. (1999)
          Brisbane, Australia
          PM10, TSP, and particle diameter.
          PM10 ranged form 11.4to 158.6 ,ug/m3. Particle sizing
          was done by a Coulter Multisizer.
Latin America

Romieu etal. (1996)
Mexico City, Mexico
During study period, maximum daily 1-h O3 ranged
from 40 to 370 ppb (mean 190 ppb, SD = 80 ppb).
24 h ave, PM10 levels ranged from 29 to 363 ,ug/m3
(mean 166.8 Mg/m3, SD 72.8 ,ug/m3).
For 53 percent of study days, PM10 levels exceeded
150 /^g/m3. PM10 was measured by a Harvard impactor.
          Romieu etal. (1997)
          Mexico City, Mexico
          During study period, maximum daily 1-h ozone ranged
          from 40 to 390 ppb (mean 196 ppb SD = 78 ppb)
          PM10 daily average ranged from 12 to 126 i/g/m3.
          PM10 was measured by a Harvard impactor.
                                        Panel of 60 asthmatic children studied for two
                                        months in Budapest, Hungary.  Mixed model
                                        used relating TSP to morning and evening
                                        PEER measurements, adjusting for SO2, time
                                        trend, day of week, temp., humidity
                                        Study examined effects of 11 dust events on
                                        peak flow and symptoms of people with asthma
                                        in Brisbane, Australia. PEE data for each
                                        individual averaged for a period of 7 days prior
                                        to the identified event. This mean was
                                        compared to the average for several days of PEE
                                        after the event, and the difference was tested
                                        using a paired t-test.
                                                            Study of 71 children with mild asthma aged 5-7
                                                            years living in the northern area of Mexico City.
                                                            Morning and evening peak flow measurements
                                                            recorded by parents. Peak flow measurements
                                                            were standardized for each person and a model
                                                            was fitted using GEE methods.  Model included
                                                            terms for minimum temperature.
                                        Study of 65 children with mild asthma aged 5-
                                        13 yr in southwest Mexico City. Morning and
                                        evening peak flow measurements made by
                                        parents. Peak flow measurements standardized
                                        for each person and model was fitted using GEE
                                        methods.  Model included terms for minimum
                                        temperature.
                                                                                             The paired t-tests were stat. significant for
                                                                                             some days, but not others. No general
                                                                                             conclusions could be drawn.
                                                                                   Ozone strongly related to changes in
                                                                                   morning PEE as well as PM10.
                                                                                             Strongest relationships were found
                                                                                             between ozone (lag 0 or 1) and both
                                                                                             morning and evening PET.
                                                                                                                                   No significant TSP-PEFR relationships found.
                                      Lag 0, PM10:
                                       Evening PEE = -4.80 (-8.00, -1.70)
                                      Lag 2, PM10:
                                       Evening PEE = -3.65 (-7.20, 0.03)
                                      LagO, PM25:
                                       Evening PEE = -4.27 (-7.12, -0.85)
                                      Lag 2, PM25:
                                       Evening PEE = -2.55 (-7.84, 2.74)
                                      Lag 1, PM10
                                       Morning PEE = -4.70 (-7.65, -1.7)
                                      Lag 2, PM10
                                       Morning PEE = -4.90 (-8.4, -1.5)

                                      Lag 0, PM10:
                                       Evening PEE = -1.32 (-6.82, 4.17)
                                      Lag 2, PM10:
                                       Evening PEE = -0.04 (-4.29, 4.21)
                                       Morning PEE = 2.47 (-1.75, 6.75)
                                      Lag 0, PM10:
                                       Morning PEE = 0.65 (-3.97, 5.32)

-------
            Appendix 8B.5: Short-Term PM Exposure Effects
                 On Symptoms in Asthmatic Individuals
April 2002                         8B-56     DRAFT-DO NOT QUOTE OR CITE

-------
to
o
o
to
          Reference citation, location, duration,
          pollutants measured, summary of values
 TABLE 8B-5. SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON
	SYMPTOMS IN STUDIES  OF ASTHMATICS	

                                                                                                                Effect measures standardized to 50 ,ug/m3
                                                                                                                PM10 (25 ,ug/m3 PM2.5). Negative coefficients
                                                                                                                for lung function and ORs greater than 1 for
                                                                                                                other endpoints suggest PM effects
Type of study, sample size, health outcomes
measured, analysis design, covariates
included, analysis problems, etc.
Results and Comments
Effects of co-pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
           United States

           Delfmoetal. (1996)
           San Diego, CA
           Sept-Oct 1993
           Ozone and PM2 5 measured. PM was measured by a
           Harvard impactor. PM2 5 ranged from 6 to 66 ^g/m3 with
           a mean of 25.

           Delfmoetal. (1997)
           San Diego County, CA
           PM10 and ozone PM was measured using a tapered-
           element oscillating microbalance. PM10 ranged from 6 to
           51 ,ug/m3 with a mean of 26.
          Delfmoetal. (1998)
          So. California community
          Aug. - Oct. 1995
          Highest 24-hour PM10 mean: 54//g/m3.
          PM10 and ozone PM was measured using a tapered-
          element oscillating microbalance. PM10 ranged from 6 to
          51 ,ug/m3 with a mean of 26.
          Yu et al. (2000) study of a panel of 133 children aged
          5-12 years in Seattle, WA. PM was measured by
          gravimetric and nephelometry methods. PM[ 0 ranged
          from 2 to 62 ,ug/m3 with a mean of 10.4. PM10 9 to
          86 ^g/m3 mean 24.7.

          Ostro et al. (2001) studied exacerbation of asthma in
          African-American children in Los Angeles. PM was
          measured by a beta-attenuated Andersen monitor.  PM1(
          ranged from 21 to 119 ,ug/m3 with a mean of 51.8.
                             Study of 12 asthmatic children with history of
                             bronchodilator use. A random effects model
                             was fitted for ordinal symptoms scores and
                             bronchodilator use in relation to 24-hr PM,,.
                             A panel of 9 adults and 13 children were
                             followed during late spring 1994 in semi-rural
                             area of San Diego County at the inversion
                             zone elevation of around  1,200 feet. A
                             random effects model was fitted to ordinal
                             symptom scores, bronchodilator use, and PEE
                             in relation to 24-hour PM10. Temp., relative
                             humidity, fugal spores, day of week and O3
                             evaluated

                             Relationship of asthma symptoms to O3 and
                             PM10 examined in a So. California
                             community with high O3  and low PM.  Panel
                             of 25 asthmatics ages 9-17 followed daily,
                             Aug. - Oct.,  1995.  Longitudinal regression
                             analyses utilized GEE model controlling for
                             autocorrelation, day of week, outdoor fangi
                             and weather.

                             Daily diary records were  collected from
                             November 1993 through  August 1995 during
                             screening for the CAMP  study. A repeated
                             measures logistic regression analysis was used
                             applied using GEE methods

                             138 children aged 8 to 13 years who had
                             physician diagnosed asthma were included.
                             A daily diary was used to record symptoms
                             and medication use. GEE methods were used
                             to estimate the effects of air pollution on
                             symptoms controlling for meteorological and
                             temporal variables.
                                           Pollen not associated with asthma symptom
                                           scores.  12-hr personal O3 but not ambient
                                           O3 related to symptoms.
                                           Although PM10 never exceeded 51 i/g/m3,
                                           bronchodilator use was significantly
                                           associated with PM10(0.76 [0.027, 0.27])
                                           puffs per 50 //g/m3.  Fungal spores were
                                           associated with all respiratory outcomes.
                                           Asthma symptoms scores significantly
                                           associated with both outdoor O3 and PM10
                                           in single pollutant and co-regressions. 1-hr
                                           and 8-hr maxi PM10 had larger effects than
                                           24-hr mean.
                                           One day lag CO and PM10 levels and the
                                           same day PM10 and S)2 levels had the
                                           strongest effects on asthma symptoms after
                                           controlling for subject specific variables
                                           and time-dependent confounders.

                                           Symptoms were generally related to PM10
                                           and NO2, but not to ozone. Reported
                                           associations were for pollutant variables
                                           lagged 3 days. Results for other lag times
                                           were not reported.
                                         No significant relationships with PM10.
                                         24-h-1.47 (0.90-2.39)
                                         8-h-2.17 (1.33-3.58)
                                         1-h-1.78 (1.25-2.53)
                                         OR symptom = 1.18 (1.05, 1.33) (PM10 same
                                         day)
                                         OR symptom = 1.17 (1.04, 1.33) (PM10 one
                                         day lag)
                                         24-h
                                         OR wheeze = 1.02 (0.99, 106) )PM10 lag 3
                                         days)
                                         OR cough = 1.06(1.02, 1.09) (PM10 lag 3
                                         days)
                                         OR shortness of breath = 1.08 (1.02, 1.13)
                                         (PM10 lag 3 days)
                                         1-h
                                         OR cough = 1.05 (1.02, 1.18) lag 3 days

-------
to
o
o
to
            TABLE 8B-5 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON SYMPTOMS
           	IN STUDIES OFASTHMATICS	

                                                                                                                                      Effect measures standardized to 50 ,ug/m3
                                                                                                                                      PM10 (25 ,ug/m3 PM2.5). Negative coefficients
                                                                                                                                      for lung function and ORs greater than 1 for
                                                                                                                                      other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates
included, analysis problems, etc.
Results and Comments
Effects of co-pollutants
OO
td
(^
OO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          United States (cont'd)

          Thurstonetal. (1997)
          Summers 1991-1993.
          O3, H+, sulfate, pollen, daily max temp, measured.
Canada

Vedaletal. (1998)
PM10 measured by Sierra-Anderson dichotomous sampler
PM10 range: -1 to 159 ,ug/m3
Port Alberrni
British, Columbia

Europe

Gielenetal. (1997)
Amsterdam, NL
PM10 and ozone.
PM10 was measured using a Sierra-Anderson
dichotomous sampler.  PM10 ranged from 15 to 60 ,ug/m3.
          Hiltermannetal. (1998)
          Leiden, NL
          July-Oct 1995.
          Ozone, PM10, NO2, SO2, BS
          PM10 ranged from 16 to 98 ,ug/m3  with a mean of 40.
          Hiltermannetal. (1997)
          The Netherlands
          Ozone and PM10
          PM10 averaged 40 ,ug/m3,
                                                    Three 5-day summer camps conducted in
                                                    1991, 1992, 1993. Study measured
                                                    symptoms and change in lung function
                                                    (morning to evening). Poisson regression for
                                                    symptoms.
                                                              206 children aged 6 to 13 years, 75 with
                                                              physician's diagnosis of asthma.  Respiratory
                                                              symptom data from diaries, GEE model.
                                                              Temp., humidity.
Study of 61 children aged 7 to 13 years living
in Amsterdam, NL. 77 percent were taking
asthma medication and the others were being
hospitalized for respiratory problems.
Respiratory symptoms recorded by parents in
diary. Associations of air pollution evaluated
using time series analyses, adjusted for pollen
counts, time trend, and day of week.

Study of 270 adult asthmatic patients from an
out-patient clinic in Leiden, NL from July 3,
to October 6, 1995. Respiratory symptom
data obtained from diaries.  An autoregressive
model was fitted to the data. Covariates
included temperature and day of week.

Sixty outpatient asthmatics examined for
nasal inflammatory parameters in The
Netherlands from July 3 to October 6, 1995.
Associations of log transformed inflammatory
parameters to 24-h PM10 analyzed, using a
linear regression model.  Mugwort-pollen and
O, were evaluated.
                                          Ozone related to respiratory symptoms
                                          No relationship between symptoms and
                                          other pollutants.
                                          PM10 associated with respiratory symptoms.
                                                                                                        Strongest relationships found with O3,
                                                                                                        although some significant relationships
                                                                                                        found with PM10.
                                                                                              PM10, O3, and NO2 were associated with
                                                                                              changes in respiratory symptoms.
                                                                                              Inflammatory parameters in nasal lavage of
                                                                                              patients with intermittent to severe
                                                                                              persistent asthma were associated with
                                                                                              ambient O3 and allergen exposure, but not
                                                                                              with PM10 exposure.
                                        LagO
                                        Cough OR = 1.08(1.00, 1.16)perlO//g/m3
                                        PM,n increments
                                        Lag 0, Symptoms:
                                         Cough OR = 2.19 (0.77, 6.20)
                                         Branch. Dial. OR = 0.94 (0.59, 1.50)
                                        Lag 2, Symptoms:
                                         Cough OR = 2.19 (0.47, 10.24)
                                         Branch. Dial. OR = 2.90 (1.80, 4.66)
                                        Lag 0, Symptoms:
                                         Cough OR = 0.93 (0.83, 1.04)
                                         Short, breath OR = 1.17 (1.03, 1.34)
                                        7 day average, Symptoms:
                                         Cough OR = 0.94 (0.82, 1.08)
                                         Short, breath OR = 1.01 (0.86, 1.20)

-------
to
o
o
to
             TABLE 8B-5 (cont'd). SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON SYMPTOMS
           	IN STUDIES OF ASTHMATICS	

                                                                                                                                        Effect measures standardized to 50 //g/m3
                                                                                                                                        PM10 (25 ,ug/m3 PM2.5). Negative coefficients
                                                                                                                                        for lung function and ORs greater than 1 for
                                                                                                                                        other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates
included, analysis problems, etc.
Results and Comments
Effects of co-pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          Europe (cont'd)

          Peters etal. (1997b)
          Erfurt, Germany
          PM fractions measured over range of sizes from ultrafme
          to fine, including PM10.
          Mean PM10 level:  55 ,ug/m3 (max 71).
          Mean SO2: 100 ,ug/m3 (max 383).
          PM was measured using a Harvard impactor.
Peters etal. (1997c)
Sokolov, Czech Republic
Winter 1991-1992
PM10, SO2, TSP, sulfate, and particle strong acid.
Median PM10: 47 ^g/m3 (29, 73).
Median SO2: 46 ^g/m3 (22, 88).
PM was measured using a Harvard impactor.  Particle
size distributions were estimated using a conduction
particle counter.

Peters etal. (1997c)
Sokolov, Czech Republic
PM10 one central site.  SO4 reported.
MeanPM10:  55 Mg/m3, max 177^g/m3.
SO4 - fine: mean 8.8 |/g/m3, max 23.8 ,ug/m3. PM was
measured using a Harvard impactor. Particle size
distributions were estimated using a conduction particle
counter.

Neukirch etal. (1998)
Paris,  France
SO2, NO2, PM13 and BS.
PM was measured by radiometry.
PM13 ranged from 9 to 95 ,ug/m3 with a mean of 34.
                                                     Study of 27 non-smoking adult asthmatics
                                                     living in Erfurt, Germany during winter
                                                     season 1991-1992.  Diary used to record
                                                     presence of cough.  Symptom information
                                                     analyzed using multiple logistic regression
                                                     analysis.
Study of 89 children with asthma in Sokolov,
Czech Republic. Subjects kept diaries and
measured peak flow for seven months during
winter of 1991-2. Logistic regression for
binary outcomes used. First order
autocorrelations were observed and corrected
for using polynomial distributed lag
structures.
                                                               Role of medication use evaluated in panel
                                                               study of 82 children, mean ages 9.8 yr., with
                                                               mild asthma in Sokolov, Czech Republic
                                                               Nov. 1991 - Feb 1992. Linear and logistic
                                                               regression evaluated PM10, SO2, temp, RH
                                                               relationships to respiratory symptoms.
                                                               Panel of 40 nonsmoking adult asthmatics in
                                                               Paris studied. GEE models used to associate
                                                               health outcomes with air pollutants. Models
                                                               allowed for time-dependent covariates,
                                                               adjusting for time trends, day of week, temp.
                                                               and humidity.
                                           Weak associations found with 5 day mean
                                           sulfates and respiratory symptoms.
Significant relationships found between
TSP and sulfate with both phlegm and
runny nose.
                                           Medicated children, as opposed to those not
                                           using asthma medication, increased their
                                           beta-agonist use in direct association with
                                           increases in 5-day mean of SO4 particles
                                           <2.5 ,um, but medication did not prevent
                                           decrease in PEF and increase in prevalence
                                           of cough attributable to PM air pollution.


                                           Significant relationships found for
                                           incidence  of respiratory symptoms and
                                           three or more day lags of SO2, and NO2.
                                           Only selected results were given.
                                                                                                                                                   Lag 0, PM10:
                                                                                                                                                    Cough OR = 1.32(1.16, 1.50)
                                                                                                                                                    Feeling ill OR = 1.20(1.01, 1.44)
                                                                                                                                                   5 Day Mean, PM10:
                                                                                                                                                    Cough OR = 1.30(1.09, 1.55)
                                                                                                                                                    Feeling ill OR = 1.47 (1.16, 1.86)
                                                                                                                                                   LagO, PM25:
                                                                                                                                                    Cough OR = 1.19(1.07, 1.33)
                                                                                                                                                    Feeling ill OR = 1.24(1.09, 1.41)
                                                                                                                                                   5 Day Mean, PM25:
                                                                                                                                                    Cough OR = 1.02(0.91, 1.15)
                                                                                                                                                    Feeling ill OR = 1.21 (1.06, 1.38)

                                                                                                                                                   Lag 0, Symptoms:
                                                                                                                                                    Cough OR = 1.01 (0.97, 1.07)
                                                                                                                                                    Phlegm OR =1.13 (1.04, 1.23)
                                                                                                                                                   5 Day Mean, Symptoms:
                                                                                                                                                    Cough OR = 1.10(1.04, 1.17)
                                                                                                                                                    Phlegm OR =1.17 (1.09, 1.27)
                                         Cough 1.16 (1.00, 1.34) 6.5 ^g/m3 increase
                                         5-day mean SO4
                                         5-d Mean SO4/increase of 6.5 ,ug/m3
                                         Beta-Agonist Use       1.46 (1.08, 1.98)
                                         Theophylline Use       0.99 (0.77, 1.26)
                                         No PM10 analysis
                                         Significant relationships found between
                                         incidence of respiratory symptoms and three
                                         or more day lags of PM13.

-------
to
o
o
to
            TABLE 8B-5 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON SYMPTOMS
           	IN STUDIES OF ASTHMATICS	
                                                                                                                                       Effect measures standardized to 50 ,ug/m3
                                                                                                                                       PM10 (25 ,ug/m3 PM2.5).  Negative coefficients
                                                                                                                                       for lung function and ORs greater than 1 for
                                                                                                                                       other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates
included, analysis problems, etc.
Results and Comments
Effects of co-pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          Europe (cont'd)

          Segalaetal. (1998)
          Paris, France
          SO2, NO2, PM13 (instead of PM10), and BS.
          PM was measured by p-radiometry.
          Giintzeletal. (1996)
          Switzerland
          SO2, NO2, TSP
Taggartetal. (1996)
Northern England
SO2, NO2 and BS.
          Latin America

          Romieuetal. (1997)
          Mexico City, Mexico
          During study period, max daily 1-h O3 range:  40 to 390
          ppb (mean 196 ppb SD = 78 ppb)
          PM10 daily average range: 12 to 126 ,ug/m3.
          PM was measured by a Harvard impactor.
          Romieuetal. (1996)
          During study period, max daily range:  40 to 370 ppb
          (mean 190 ppb, SD = 80 ppb).
          24 h ave. PM10 levels range: 29 to 363 ,ug/m3 (mean
          166.8 Mg/m3, SD 72.8 ,ug/m3).
          PM10 levels exceeded 150 ,ug/m3 for 53% of study days.
          24-h ave. PM25 levels range 23-177 i/g/m3 (mean
          85.7,ug/m3)
          PM was measured by a Harvard impactor.	
Study of 43 mildly asthmatic children aged
7-15 yr in Paris.  Patients followed Nov. 15,
1992 to May 9, 1993. Respiratory symptoms
recorded daily in diary.  An autoregressive
model fitted to data using GEE methods.
Covariates included temp, and humidity.

An asthma reporting system was used in
connection with pollutant monitoring in
Switzerland from fall of 1988 to fall 1990.
A Box-Jenkins ARIMA time series model was
used to relate asthma to TSP, O3, SO2, and
NO2 after adjusting for temperature.

Panel of 38 adult asthmatics studied July 17
to Sept. 22, 1993 in northern England. Used
generalized linear model to relate pollutants to
bronchial hyper-responsiveness, adjusting for
temperature.
                                                    Study of 65 children with mild asthma aged
                                                    5-13 yr living in southwest Mexico City.
                                                    Respiratory symptoms recorded by the
                                                    parents in daily diary. An autoregressive
                                                    logistic regression model used to analyze
                                                    presence of respiratory symptoms.
                                                    Study of 71 children with mild asthma aged
                                                    5-7 yr living in northern Mexico City.
                                                    Respiratory symptoms recorded by parents in
                                                    daily diary. An autoregressive logistic
                                                    regression model was used to analyze the
                                                    presence of respiratory symptoms.
                                                                                              Effects found related to PM13 were less than
                                                                                              those found related to the other pollutants.
                                                                                              No significant relationships found.
Small effects seen in relation to NO2 and
BS.
                                          Strongest relationships found between O3
                                          and respiratory symptoms.
                                          Cough and LRI were associated with
                                          increased O3 and PM10 levels.
                                        Lag 2, Symptoms:
                                          Short. Breath OR = 1.22 (0.83, 1.81)
                                          Resp. Infect. OR = 1.66 (0.84, 3.30)
                                        Lag 0, Symptoms:
                                         Cough OR = 1.05(0.92, 1.18)
                                         Phlegm OR =1.05 (0.83, 1.36)
                                         Diff Breath OR = 1.13 (0.95,  1.33)
                                        Lag 2, Symptoms:
                                         Cough OR = 1.00(0.92, 1.10)
                                         Phlegm OR = 1.00(0.86, 1.16)
                                         Diff. Breath OR = 1.2(1.1, 1.36)

                                        PM10 (lag 0) increase of 50 //g/m3 related to:
                                        LRI= 1.21  (1.10, 1.42)
                                        Cough = 1.27(1.16, 1.42)
                                        Phlegm  = 1.21 (1.00, 1.48)
                                        PM25 (lag 0) increase of 25 ,ug/m3 related to:
                                        LRI= 1.18(1.05, 1.36)
                                        Cough = 1.21(1.05, 1.39)
                                        Phlegm  = 1.21 (1.03, 1.42)

-------
            Appendix 8B.6: Short-Term PM Exposure Effects
               On Pulmonary Function in Nonasthmatics
April 2002                         8B-61      DRAFT-DO NOT QUOTE OR CITE

-------
to
o
o
to
 TABLE 8B-6.  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON PULMONARY FUNCTION
	TESTS IN STUDIES OF NONASTHMATICS	

                                                                                                                                 Effect measures standardized to 50 ,ug/m3
                                                                                                                                 PM10 (25 ,ug/m3 PM2.5).  Negative coefficients
                                                                                                                                 for lung function and ORs greater than 1 for
                                                                                                                                 other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
                                                      Type of study, sample size, health outcomes
                                                      measured, analysis design, covariates included,
                                                      analysis problems, etc.
Results and Comments
Effects of co-pollutants
OO
td
ON
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           UnitedStates

           Hoeketal. (1998)
           (summary paper)
          Lee and Shy (1999)
          North Carolina
          Mean 24 h PM10 cone, over two years: 25.1 ,ug/m3.
          Korricketal. (1998)
          Mt. Washington, NH
          O3 levels measured at 2 sites near top of the mountain.
          PM2 5 measured near base of the mountain.
          PM was measured by a Harvard impactor.

          Naeheretal. (1999)
          Virginia
          PM10, PM25, sulfate fraction, H+, and ozone
Neas etal. (1996)
State College, PA
PM21. mean 23.5; max 85.8 ,ug/m3.
                                               Results summarized from several other studies
                                               reported in the literature.  These included:
                                               asymptomatic children in the Utah Valley (Pope
                                               et al., 1991), children in Bennekom, NL (Roemer
                                               et al., 1993), children in Uniontown, PA (Neas et
                                               al., 1995), and children in State College, PA
                                               (Neas et al., 1996). Analyses done using a first-
                                               order autoregressive model with adjustments for
                                               time trend and ambient temp.

                                               Study of the respiratory health status of residents
                                               whose households lived in six communities near
                                               an incinerator in southwestern North Carolina.
                                               Daily PEER measured in the afternoon was
                                               regressed against 24 hour PM10 level lagged by
                                               one day.  Results were adjusted for gender, age,
                                               height, and hypersensitivity.

                                               Study of the effects of air pollution on adult
                                               hikers on Mt. Washington, NH. Linear and non-
                                               linear regressions used to evaluate effects of
                                               pollution on lung function.
                                               Daily change in PEE studied in 473 non-smoking
                                               women in Virginia during summers 1995-1996.
                                               Separate regression models run, using normalized
                                               morning and evening PEE for each individual.

                                               Study of 108 children in State College, PA, during
                                               summer of 1991 for daily variations in symptoms
                                               and PEFRs in relation to PM2 [ An autoregressive
                                               linear regression model was used. The regression
                                               was weighted by reciprocal number of children of
                                               each reporting period. Fungus spore cone., temp.,
                                               O, and SO2 were examined.
                                                                                                    Other pollutants not considered.
                                                                                                    PM10 was not related to variations in
                                                                                                    respiratory health as measured by
                                                                                                    PEER.
                                                                                                    PM2 5 had no effect on the O3
                                                                                                    regression coefficient.
                                                                                                    Ozone was only pollutant related to
                                                                                                    evening PEE.
                                                                                                               Spore concentration associated with
                                                                                                               deficient in morning PERF.
                                   Significant decreases in peak flow found to be
                                   related to PM10 increases.
                                   Morning PEE decrements were associated with
                                   PM10, PM2 5, and H+. Estimated effect from
                                   PM2 5 and PM10 was similar. No PM effects
                                   found for evening PEE.

                                   PM2 j (25 //g/m3) related to RR of:
                                   PM PFER (lag 0) = -0.05 (-1.73, 0.63)
                                   PM PEER (lag 1)= -0.64 (-1.73, 0.44)
O

-------
to
o
o
to
   TABLE 8B-6 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON PULMONARY FUNCTION
  	TESTS IN  STUDIES OF NONASTHMATICS	
                                                                                                                                         Effect measures standardized to 50 //g/m3
                                                                                                                                         PM10 (25 fj.g/m1 PM2.5). Negative coefficients
                                                                                                                                         for lung function and ORs greater than 1 for
                                                                                                                                         other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
                                                                                                                Results and Comments
                                                                                                                Effects of co-pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
           United States (cont'd)

           Neasetal. (1999)
           Philadelphia, PA
           Median PM10 level: 31.6 in SW camps,
           27.8 in NE camps (IQR ranges of about 18).
           Median PM2 5 level: 22.2 in the SW camps,
           20.7 in NE camps (IQR ranges about 16.2 and 12.9,
           respectively).
           Particle-strong acidity, fine sulfate particle, and O3 also
           measured.
Schwartz and Neas (2000)
Eastern U.S.
PM2 5 and CM (PM10.2 5) measured.
Summary levels not given.
          Linn etal. (1996)
          So. California
          NO2 ozone, and PM5 measured.
          PM5 was measured using a Marple low volume sampler PM5
          ranged from 1-145 //g/m3 with a mean of 24.
          Europe

          Boezen etal. (1999)
          Netherlands
          PM10, BS, SO2, and NO2 measured, but methods were not
          given.  PM10 ranged from 4.8 to 145//g/m3 with site means
          ranging from 26 to 54 //g/m3.
                                                       Panel study of 156 normal children attending
                                                       YMCA and YWCA summer camps in greater
                                                       Philadelphia area in 1993. Children followed for
                                                       at most 54 days. Morning and evening deviations
                                                       of each child's PEE were analyzed using a mixed-
                                                       effects model adjusting for autocorrelation.
                                                       Covariates included time trend and temp. Lags
                                                       not used in the analysis.
Analyses for 1844 school children in grades 2-5
from six urban areas in eastern U.S. and from
separate studies from Uniontown and State
College, PA. Lower resp. symptoms, cough and
PEE used as endpoints. The authors replicated
models used in the original analyses.  CM and
were used individually and jointly in the analyses.
Sulfate fractions also used in the analyses.
Details of models not given.

Study of 269 school children in Southern
California twice daily for one week in fall, winter
and spring for two years. A repeated measures
analysis of covariance was used to fit an
autoregressive model, adjusting for year, season,
day of week, and temperature.
                                                       Data collected from children during three winters
                                                       (1992-1995) in rural and urban areas of The
                                                       Netherlands. Study attempted to investigate
                                                       whether children with bronchial
                                                       hyperresponsiveness and high serum Ige levels
                                                       were more susceptible to air pollution.  Prevalence
                                                       of a 10 percent PEE decrease  was related to
                                                       pollutants for children with bronchial
                                                       hyperresponsiveness and high serum Ige levels.
                                              Analyses that included sulfate fraction
                                              and O3 separately also found
                                              relationship to decreased flow. No
                                              analyses reported for multiple
                                              pollutant models.
                                                                                                                Sulfate fraction was highly correlated
                                                                                                                with PM25 (0.94), and, not
                                                                                                                surprisingly, gave similar answers.
                                                                                                     Morning FVC was significantly
                                                                                                     decreased as a function of PM5 and
                                                                                                     NO,
                                              No consistent pattern of effects
                                              observed with any of the pollutants
                                              for 0, 1, and 2 day lags.
Lag 0, PM10:
 Morning PEE = -8.16 (-14.81, -1.55)
 Evening PEE = -1.44 (-7.33, 4.44)
5 day ave, PM10
 Morning PEE = 2.64 (-6.56, 11.83)
 Evening PEE = 1.47 (-7.31, 10.22)
LagO,PM25
  Morning PEE = -3.28 (-6.64, 0.07)
  Evening PEE = -0.91 (-4.04, 2.21)
5 day ave., PM2 5
 Morning PEE = 3.18 (-2.64, 9.02)
 Evening PEE = 0.95 (-4.69, 6.57)

Uniontown Lag 0,PM25 :
 Evening PEE = -1.52 (-2.80, -0.24)
State College Lag 0, PM25:
 Evening PEE = -0.93 (-1.88, 0.01)

Results presented for CM showed no effect.
Results for PM10 were not given.

-------
to
o
o
to
 TABLE  8B-6 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON PULMONARY FUNCTION
	TESTS IN STUDIES OF NONASTHMATICS	

                                                                                                                                       Effect measures standardized to 50 ,ug/m3
                                                                                                                                       PM10 (25 ,ug/m3 PM2.5).  Negative coefficients
                                                                                                                                       for lung function and ORs greater than 1 for
                                                                                                                                       other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
                                                                                                                Results and Comments
                                                                                                                Effects of co-pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          Europe (cont'd)

          Frischeretal. (1999)
          Austria
          PM10 measured gravimettrically for 14-d periods.
          Annual mean PM10 levels range:  13.6 - 22.9 ,ug/m3.
          O3 range: 39.1 ppb - 18.5 pbs between sites.
          Grievinketal. (1999)
          Netherlands
          PM10 and BS.
          PM10 ranged from 12 to 123 //g/m3 with a mean of 44.
          Kiinzli et al. (2000)
          Roemer et al. (2000)
          PM10 means for 17 panels ranged 11.2 to 98.8 ,ug/m3.
          SO2, NO2, and elemental content of PM also measured.
          Measurement methods were not described.
          Scarlett etal. (1996)
          PM10, O3, and NO2 measured.
                                                     At nine sites in Austria during 1994, 1995, and
                                                     1996, a longitudinal study designed to evaluate O3
                                                     was conducted. During 1994 - 1996, children
                                                     were measured for FVC, FEVj and MEF50 six
                                                     times, twice a year in spring and fall. 1060
                                                     children provided valid function tests. Mean age
                                                     7.8±0.7yr.  GEE models used. PM10, SO2, NO2,
                                                     and temp, evaluated.

                                                     A panel of adults with chronic respiratory
                                                     symptoms studied over two winters in The
                                                     Netherlands starting in 1993/1994.  Logistic
                                                     regression analysis was used to model the
                                                     prevalence of large PEF decrements.
                                                     Individual linear regression analysis of PEF on
                                                     PM was calculated and adjusted for time trends,
                                                     influenza incidence, and meteorological variables.

                                                     Ackermann-Liebrich et al. (1997) data reanalyzed.
                                                     Authors showed that a small change in FVC
                                                     (-3.14 percent) can result in a 60% increase in
                                                     number of subjects with FVC less than 80 percent
                                                     of predicted.

                                                     Combined results from 1208 children divided
                                                     among 17 panels studied.  Separate results
                                                     reported by endpoints included symptoms as
                                                     reported in a dairy and PEF. Individual panels
                                                     were analyzed using multiple linear regression
                                                     analysis on deviations from mean PEF adjusting
                                                     for auto-correlation. Parameter estimates were
                                                     combined using a fixed-effects model where
                                                     heterogeneity was not present and a random-
                                                     effects model where it was present.

                                                     In study of 154 school children, pulmonary
                                                     function was measured daily for 31  days.
                                                     Separate autoregressive models for each child
                                                     were pooled, adjusting for pollen, machine,
                                                     operator, time of day, and time trend.
                                              Small but consistent lung function
                                              decrements in cohort of school
                                              children associated with ambient O3
                                              exposure.
                                              Subjects with low levels of serum
                                              P-carotene more often had large PEF
                                              decrements when PM10 levels were
                                              higher, compared with subjects with
                                              high serum P-carotene.
                                              Results suggested serum P-carotene
                                              may attenuate the PM effects on
                                              decreased PEF.

                                              The results were for two hypothetical
                                              communities, A and B.
                                              Daily concentrations of most elements
                                              were not associated with the health
                                              effects.
                                              PM10 was related to changes in FEV
                                              and FVC
                                                                                                                                                   PM10 showed little variation in exposure
                                                                                                                                                   between study site.  For PM10, positive effect
                                                                                                                                                   seen for winter exposure but was completely
                                                                                                                                                   confounded by temperature.

                                                                                                                                                   PM10 Summertime
                                                                                                                                                             P = 0.003 SE 0.012 p=0.77
                                                                                                                                                   PM10 analyses not focus of this paper.

-------
to
o
o
to
   TABLE 8B-6 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON PULMONARY FUNCTION
  	TESTS IN STUDIES OF NONASTHMATICS	

                                                                                                                                         Effect measures standardized to 50 ,ug/m3
                                                                                                                                         PM10 (25 ,ug/m3 PM2.5).  Negative coefficients
                                                                                                                                         for lung function and ORs greater than 1 for
                                                                                                                                         other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
Results and Comments
Effects of co-pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          Europe (cont'd)

          van der Zee etal. (1999)
          Netherlands
          PM10 averages ranged 20 to 48 ,ug/m3.
          BS, sulfate fraction, SO2, and NO2 also measured.
van der Zee et al. (2000)
Netherlands
PM10 averages ranged 24 to 53 /2g/m3.
BS, sulfate fraction, SO2, and NO2 also measured.
PM10 was measured using a Sierra Anderson 241
dichotomous sampler.
          Tiittanen etal. (1999)
          Kupio, Finland
          Median PM10 level: 28 (25th, 75th percentiles = 12, 43).
          Median PM25 level:  15 (25th, 75th percentiles = 9, 23).
          Black carbon, CO, SO2, NO2, and O3 also measured.
          PM was measured using single stage Harvard samplers.
Panel study of 795 children aged 7 to 11 years,
with and without chronic respiratory symptoms
living in urban and nonurban areas in the
Netherlands.  Peak flow measured for three
winters starting in 1992/1993.  Peak flow
dichotomized at 10 and 20% decrements below
the individual median. Number of subjects was
used as a weight.  Minimum temperature day of
week, and time trend variables were used as
covariates. Lags of 0,  1 and 2 days were used, as
well as  5 day moving average.

Panel study of 489 adults aged 50-70 yr, with and
without chronic respiratory symptoms,  living in
urban and nonurban areas in the Netherlands.
Resp. symptoms and peak flow measured for three
winters starting in 1992/1993.  Symptom
variables analyzed as a panel instead of using
individual responses. The analysis was treated as
a time series, adjusting for first order
autocorrelation. Peak flow dichotomized at
10 and  20% decrements below the individual
median. The number of subjects used as a weight.
Minimum temp., day of week,  and time trend
variables used as covariates. Lags of 0, 1 and
2 days used, as well as 5 day moving average.

Six-week panel study of 49 children with chronic
respiratory disease followed in the spring of 1995
in Kuopio, Finland.  Morning and evening
deviations of each child's PEF analyzed, using a
general linear model estimated by PROC MIXED.
Covariates included a time trend, day of week,
temp., and humidity. Lags of 0 through 3 days
were used, as well as a 4-day moving average.
Various fine particles were examined.
                                                                                                     In children with symptoms,
                                                                                                     significant associations found
                                                                                                     between PM10, BS and sulfate fraction
                                                                                                     and the health endpoints. No multiple
                                                                                                     pollutant models analyses reported.
BS tended to have the most consistent
relationship across endpoints. Sulfate
fraction also related to increased
respiratory effects.  No analyses
reported for multiple pollutant
models. Relationship found between
PM10 and the presence of 20%
decrements in symptomatic subjects
from urban areas.
                                                                                                     Ozone strengthened the observed
                                                                                                     associations. Introducing either NO2
                                                                                                     or SO2 in the model did not change
                                                                                                     the results markedly.  Effects varied
                                                                                                     by lag.  Separating effects by size was
                                                                                                     difficult.
                                    Lag 0, PM10, Urban areas
                                     Evening PEF OR = 1.15 (1.02, 1.29)
                                    Lag 2, PM10, Urban areas
                                     Evening PEF OR = 1.07 (0.96, 1.19)
                                    5 day ave, PM10, Urban areas
                                     Evening PEF = 1.13 (0.96, 1.32)
Lag 0, PM10, Urban areas
 Morning large decrements
 OR= 1.44(1.02,2.03)
 Lag 2, PM10, Urban areas
 Morning large decrements
 OR= 1.14(0.83,  1.58)
5 day ave, PM10, Urban areas
 Morning large decrements
 OR= 1.16(0.64,2.10)

Results should be viewed with caution because
of problems in analysis.
                                    Lag 0, PM10:
                                     Morning PEF = 1.21 (-0.43, 2.85)
                                     Evening PEF = 0.72 (-0.63, 1.26)
                                    4 day ave, PM10
                                     Morning PEF= -1.26(-5.86, 3.33)
                                     Evening PEF = 2.33 (-2.62, 7.28)
                                    LagO,PM25
                                     Morning PEF = 1.11 (-0.64, 2.86)
                                     Evening PEF = 0.70 (-0.81, 2.20)
                                    4 day ave., PM25
                                     Morning PEF = -1.93 (-7.00, 3.15)
                                     Evening PEF = 1.52 (-3.91, 6.94)

-------
to
o
o
to
   TABLE 8B-6 (cont'd). SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON PULMONARY FUNCTION
  	TESTS IN STUDIES OF NONASTHMATICS	

                                                                                                                                        Effect measures standardized to 50 ,ug/m3
                                                                                                                                        PM10 (25 ,ug/m3 PM2.5). Negative coefficients
                                                                                                                                        for lung function and ORs greater than 1 for
                                                                                                                                        other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
Results and Comments
Effects of co-pollutants
          Europe (cont'd)
oo
td
ON
Oi
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Ward et al. (2000)
          West Midlands, UK
          Daily measurements of PM10, PM2 5, SO2, CO, O3, and
          oxides of nitrogen.
          Details on PM monitoring were inomplete.
Osunsanya et al. (2001) studied 44 patients aged > 50 with
COPD in Aberdeen, UK. PM was measured using tapered
element oscillating microbalance. Particle sizes were
measured a TSI model 3934 scanning particle sizer. PM10
ranged from 6 to 34 /2g/m3 with a median of 13.
          Cuijpers et al. (1994)
          Maastricht, NL
          SO2, NO2, BS, ozone, and H+ measured. PM measurements
          were made with a modified Sierra Anderson sampler.  PM10
          ranged from 23 to 54 //g/m3.

          Latin America

          Gold etal. (1999)
          Mexico City, Mexico
          Mean 24 h O3 levels: 52ppb.
          MeanPM25:  30 |/g/m3.
          MeanPM10:  49,ug/m3.
Panel study of 9 yr old children in West Midlands,
UK for two 8-week periods representing winter
and summer conditions.  Individual PEE values
converted to z-values.  Mean of the z-values
analyzed in a linear regression model, including
terms for time trend, day of week, meteorological
variables, and pollen count. Lags up to four days
also used.

Symptom scores, bronchodilator use, and PEE
were recorded daily for three months. GEE
methods were used to analyze the dichotomous
outcome measures. PEE was converted to a
dichotomous measure by defining a 10 percent
decrement as the outcome of interest.
                                                       Summer episodes in Maastricht, The Netherlands
                                                       studied. Paired t tests used for pulmonary
                                                       function tests.
                                                       Peak flow studied in a panel of 40 school-aged
                                                       children living in southwest Mexico City. Daily
                                                       deviations from morning and afternoon PEFs
                                                       calculated for each subject. Changes in PEE
                                                       regressed on individual pollutants allowing for
                                                       autocorrelation and including terms for daily
                                                       temp., season,  and time trend.
                                                                                                     Results on effects of pollution on lung
                                                                                                     function to be published elsewhere.
No associations were found between
actual PEE and PM10 or ultrafme
particles. A change of PM10 from 10 to
20 |/g/m3 was associated with a 14
percent decrease in the rate of high
scores of shortness of breath.  A
similar change in PM10 was associated
with a rate of high scores of cough.

Small decreases in lung function
found related to pollutants.
                                              O3 significantly contributed to
                                              observed decreases in lung function,
                                              but there was an independent PM
                                              effect.
                                                                                                                                                   The endpoint was measured in terms of scores
                                                                                                                                                   rather than L/min.
                                                                                  Quantitative results not given.
                                    Both PM2 5 and PM10 significantly related to
                                    decreases in morning and afternoon peak flow.
                                    Effects of the two pollutants similar in
                                    magnitude when compared on percent change
                                    basis.
O

-------
to
o
o
to
 TABLE 8B-6 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON PULMONARY FUNCTION
	TESTS IN STUDIES OF NONASTHMATICS	

                                                                                                                                      Effect measures standardized to 50 ,ug/m3
                                                                                                                                      PM10 (25 ,ug/m3 PM2.5). Negative coefficients
                                                                                                                                      for lung function and ORs greater than 1 for
                                                                                                                                      other endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
                                                                                                               Results and Comments
                                                                                                               Effects of co-pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          New Zealand

          Harreetal. (1997)
          Christchurch, NZ
          SO2, NO2, PM10, and CO measured.
          Details on monitoring methods and pollutant ranges were
          not given.
          Jalaludin et al. (2000) studied PEE in 148 children 6
          primary schools in Sydney, Australia. PM was measured by
          tapered element oscillating microbalance. Mean PM10 was
          22.8+- 13.9 ,ug/m3.
          Asia

          Chen etal. (1999)
          Taiwan
          Beta-gauge PM10 ranged 44.5 to 189.0 ,ug/m3 for peak
          concentrations.
          Tan et al. (2000) Southeast Asian smoke-haze event
          9/29-10/27 1997
          PM10 mean daily was 125.4 ± 44.9 |/g/m3
          ultra range of 47 to 216 ,ug/m3 in Singapore
                                                     Study of 40 subjects aged over 55 years with
                                                     COPD living in Christchurch, New Zealand
                                                     conducted during winter of 1994.  Subjects
                                                     recorded their peak flow measurements.  A log-
                                                     linear regression model with adjustment for first
                                                     order auto-correlation was used to  analyze peak
                                                     flow data  and a Poisson regression model was
                                                     used to analyze symptom data.

                                                     148 children in grades 3-5 were followed  for 11
                                                     months, recording PEE twice daily. The
                                                     normalized change in PEE was analyzed using
                                                     GEE methods.  PEE was related to SO3, PM10,
                                                     NO2, as well as meteorological variables.
                                                     In 3 Taiwan communities in 1995, PM10 by B-
                                                     gauge measured at selected primary schools in
                                                     each community.  Spirometry tests (FVC, FEVLO,
                                                     FEF25.75%, PEE) obtained in period May 1995 to
                                                     Jan. 1996 using ATS protocol in study pop. aged
                                                     8 to 13 yr. 895 children were analyzed.  Study
                                                     was designed to investigate short-term effect of
                                                     ambient air pollution in cross-sectional survey.
                                                     Multivariate linear model analysis used in both
                                                     one pollutant and multipollutant models, with 1-,
                                                     2-, and 7-day lags. SO2, CO, O3, NO2 and PM10
                                                     examined, as were meteorol. variables.

                                                     Examined the association between acute air
                                                     pollution caused by biomass burning and
                                                     peripheral UBC counts in human serial
                                                     measurement made during the event were
                                                     compared with a period after the haze cleared
                                                     (Nov. 21 -Dec. 5, 1997)
                                              Few significant associations found
                                              between the health endpoints and the
                                              pollutants.
                                              Daily mean deviations in PEE were
                                              related to ozone, but no relationships
                                              were found with PM10 or NO2.
                                              Multiple pollutant models gave
                                              similar results to those given by the
                                              single pollutant models.
                                              In the one-pollutant model, daytime
                                              peak O3 cone, with a 1-day lag
                                              significantly affected both FVC and
                                              FEVj. NO2, SO2, CO affected FVC.
                                              PM10 showed nonsignificant
                                              decrement. No significant result
                                              demonstrated in the model for the
                                              exposure with 7 days lag.  In the
                                              multi-pollutant model, only peak  O3
                                              cone, with 1-day lag showed sig.
                                              effect on FVC and FEVL0.


                                              Indices of atmospheric pollution were
                                              significantly associated in the
                                              elevated band neutrophil counts
                                              expressed as a percentage of total
                                              polymonphonuclear leukocytes
                                              (PMN). No statistically significant
                                              difference in FEU! and FUC were
                                              observed during and after haze
                                              exposure.
                                                                                                                                                   Lag 0, PM10:
                                                                                                                                                    PEF=-0.£
                                                                                                                                                                (-2.33, 0.61)
                                                                                                                                                   Change from AM to PM PEE = 0.045 (-.205,
                                                                                                                                                   2.95) lag one day
                                                                                                                                                   One pollutant model daytime average

                                                                                                                                                   PM10-2daylag
                                                                                                                                                             FVC-0.37 se 0.39

-------
            Appendix 8B.7: Short-Term PM Exposure Effects
                    On Symptoms in Nonasthmatics
April 2002                        8B-68      DRAFT-DO NOT QUOTE OR CITE

-------
to
o
o
to
          Reference citation, location, duration,
          pollutants measured, summary of values
                   TABLE 8B-7.  SHORT-TERM PARTICIPATE  MATTER EXPOSURE EFFECTS  ON SYMPTOMS
                                                              IN STUDIES OF NONASTHMATICS

                                                                                                                                         Effect measures standardized to 50 //g/m3
                                                                                                                                         PM10 (25 fj.g/m1 PM2.5). Negative coefficients
                                                                                                                                         for lung function and ORs greater than 1 for
                                                                                                                                         other endpoints suggest PM effects
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
                                                                                                                Results and Comments
                                                                                                                Effects of co-pollutants
OO
td
ON
VO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
           United States

           Schwartz and Neas (2000)
           Eastern U.S.
           PM25 and CM (PM10_25 by substation)..
           Summary levels not given
Zhang et al. (2000)
Vinton, Virginia
24- h PM10, PM25, sulfate and strong acid measured in
1995.
           Canada

           Long etal. (1998)
           Winnepeg, CN
           PM10, TSP, and VOC measured.
           Methods for PM monitoring not given.  Ranges of values
           also not given.

           Europe

           Boezen etal. (1998)
           Amsterdam, NL
           PM10, SO2, and NO2 measured.
           PM10 ranged from 7.9 to 242.2 //g/m3 with a median of 43.
Reported on analysis of 1844 school children in
grades 2-5 from six urban areas in the eastern
U.S., and from separate studies from Uniontown
and State College, PA. Lower respiratory
symptoms, and cough used as endpoints. The
authors replicated the models used in the original
analyses. CM and PM25 were used individually
and jointly in the analyses.  Sulfates fractions
were also used in the analyses. Details of the
models were not given.

In southwestern Virginia, 673 mothers were
followed June 10 to Aug. 31, 1995 for the daily
reports of present or absence of runny or stuffy
nose. PM indicator, O3, NO2 temp., and random
sociodemographic characteristics considered.
                                                       Study of 428 participants with mild airway
                                                       obstruction conducted during a Winnepeg
                                                       pollution episode. Gender specific odds ratios of
                                                       symptoms were calculated for differing PM10
                                                       levels using the Breslow-Day test.
                                                       Study of 75 symptomatic and asymp. adults near
                                                       Amsterdam for three months during winter 1993-
                                                       1994. An autoregressive logistic model was used
                                                       to relate PM10 to respiratory symptoms, cough,
                                                       and phlegm, adjusting for daily min. temp., time
                                                       trend, day of week.
                                                                                                     Sulfate fraction was highly correlated
                                                                                                     with PM25 (0.94), and not
                                                                                                     surprisingly gave similar answers.
                                                                                                                Of all pollutants considered, only the
                                                                                                                level of coarse particles as calculated
                                                                                                                (PM10 - PM2 5)  independently related
                                                                                                                to incidence of new episode of runny
                                              Cough, wheezing, chest tightness, and
                                              shortness of breath were all increased
                                              during the episode
                                              No relationship found with pulmonary
                                              function. Some significant
                                              relationships with respiratory disease
                                              found in subpopulations
                                                                                                                                                    PM2 5 was found to be significantly related to
                                                                                                                                                    lower respiratory symptoms even after
                                                                                                                                                    adjusting for CM, whereas the reverse was not
                                                                                                                                                    true. However, for cough, CM was found to be
                                                                                                                                                    significantly related to lower respiratory
                                                                                                                                                    symptoms even after adjusting for PM2 5,
                                                                                                                                                    whereas the reverse was not true.
O

-------
            TABLE 8B-7 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON SYMPTOMS
           	IN STUDIES OF NONASTHMATICS	

                                                                                                                                     Effect measures standardized to 50 ,ug/m3 PM10
                                                                                                                                     (25 ,ug/m3 PM2.5). Negative coefficients for
                                                                                                                                     lung function and ORs greater than 1 for other
                                                                                                                                     endpoints suggest PM effects
Reference citation, location, duration,
pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
Results and Comments
Effects of co-pollutants
Europe (cont'd)

Howel et al. (2001) study of children's respiratory
health in 10 non-urban communities of northern
England.  PM levels were measured using a single
continuous real-time monitor. PM10 levels ranged
from 5 to 54 ,ug/m3.
Roemeretal. (1998)
Mean PM10 levels measured at local sites ranged
11.2 to 98.8 ,ug/m3 over the 28 sites.
Roemer et al. (2000)
PM10 means for the 17 panels ranged 11.2 to
98.8 //g/m3.
SO2, NO2, and PM elemental content also measured.
Measurement methods were not described.
                                                  The study included 5 pairs of non-urban
                                                  communities near and not so near 5 coal mining
                                                  sites. 1405 children aged 1-11 years were
                                                  included.  275 of the children reported having
                                                  asthma. Diaries of respiratory symptoms were
                                                  collected over a 6 week period. PM10, measured
                                                  by a single continuous real-time monitor, ranged
                                                  from 5 to 54 ,ug/m3.

                                                  Pollution Effects on Asthmatic Children in
                                                  Europe (PEACE) study was a multi-center study
                                                  of PM10, BS, SO2, and NO2 on respiratory health
                                                  of children with chronic respiratory symptoms.
                                                  Results from individual centers were reported by
                                                  Kotesovec et al. (1998), Kalandidi et al. (1998),
                                                  Haluszka et al. (1998),  Forsberg et al. (1998),
                                                  Clench-Aas et al. (1998), and Beyer et al.  (1998).
                                                  Children with chronic respiratory symptoms
                                                  were selected into the panels.  The symptom with
                                                  one of the larger selection percentages was dry
                                                  cough (range over sample of study communities
                                                  29 to 92% [22/75; 84/91] with most values over
                                                  50%).  The group as a whole characterized as
                                                  those with chronic respiratory disease, especially
                                                  cough.

                                                  Combined results from 1208 children divided
                                                  among 17 panels studied. Endpoints included
                                                  symptoms as reported in a dairy and PEE.
                                                  Symptom variables analyzed as a panel instead of
                                                  using individual responses. The analysis was
                                                  treated as  a time series, adjusting for first order
                                                  autocorrelation. Parameter estimates were
                                                  combined using a fixed-effects model where
                                                  heterogeneity was not present and a random-
                                                  effects model where it was present.
                                             The associations found between daily
                                             PM10 levels and respiratory symptoms
                                             were frequently small and positive and
                                             sometimes varied by community.
                                             These studies modeled group rates and
                                             are an example of the panel data
                                             problem.
                                      OR wheeze = 1.16(1.05, 1.28(PM10)
                                      OR cough = 1.09(1.02, 1.16)(PM10)
                                      OR reliever use = 1.00 (0.94, 1.06) (PM10)
                                             Daily concentrations of most elements
                                             were not associated with the health
                                             effects.
                                      The analysis of PM1(
                                      paper.
                                                                                                                                                       was not a focus of this

-------
to
o
o
to
            TABLE 8B-7 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON SYMPTOMS
           	IN STUDIES OF NONASTHMATICS	

                                                                                                                                       Effect measures standardized to 50 ,ug/m3 PM10
                                                                                                                                       (25 fj.g/m1 PM2.5). Negative coefficients for
                                                                                                                                       lung function and ORs greater than 1 for other
                                                                                                                                       endpoints suggest PM effects
          Reference citation, location, duration,
          pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
                                                                                                           Results and Comments
                                                                                                           Effects of co-pollutants
OO
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
          Europe (cont'd)

          van der Zee etal. (1999)
          Netherlands
          PM10 averages ranged 20 to 48 ,ug/m3.
          BS, sulfate fraction, SO2, and NO2 also measured.
van der Zee et al. (2000)
Netherlands
Daily measurements of PM10, BS, fine sulfate, nitrate,
ammonium and strong acidity.
PM10 was measured using a Sierra Anderson 241
dichotomous sampler.
          Tiittanen etal. (1999)
          Kupio, Finland
          Median PM10 level: 28 (25th, 75th percentiles = 12,
          43).
          Median PM25: 15 (25th and 75th percentiles of 9 and
          23). Black carbon, CO, SO2, NO2, and O3 also
          measured.  PM was measured using single stage
          Harvard samplers.
          Keles etal. (1999)
          Istanbul, Turkey
          Nov. 1996 to Jan. 1997.
          TSP levels ranged from annual mean of 22 ,ug/m3 in
          unpolluted area to 148.8 //g/m3 in polluted area.
A panel study of 795 children aged 7 to 11 yr,
with and without chronic respiratory symptoms,
living in urban and nonurban areas in the
Netherlands. Respiratory symptoms measured for
3 winters starting 1992/1993. Symptom variables
analyzed as a panel instead of using individual
responses.  The analysis was treated as a time
series, adjusting for first order autocorrelation.
The number of subjects was used as a weight.
Minimum temp., day of week, and time trend
variables used as covariates. Lags of 0, 1 and 2
days used, as well as 5 day moving average.

Panel study of adults aged 50 to 70 yr during 3
consecutive winters starting in 1992/1993.
Symptom variables analyzed as a panel instead of
using individual responses. Analysis treated as a
time series,  adjusting for first order
autocorrelation. Number of subjects used as a
weight. Min. temp., day of week,  time trend
variables used as covariates. Lags 0, 1 and 2
days used, as well as 5 day moving average.

Six-week panel study of 49 children with chronic
respiratory disease followed in spring 1995 in
Kuopio, Finland.  Cough,  phlegm, URS, LRS and
medication use analyzed, using a random effects
logistic regression model (SAS macro
GLIMMFX). Covariates included  a time trend,
day of week, temp., and humidity. Lags of 0 to 3
days used, as well as 4-day moving average.
                                                  Symptoms of rhinitis and atopic status were
                                                  evaluated in 386 students grades 9 and 10 using
                                                  statistical package for the social sciences, Fisher
                                                  tests, and multiple regression model as
                                                  Spearman's coefficient of correlation.
                                                                                                In children with symptoms, significant
                                                                                                associations found between PM10, BS
                                                                                                and sulfate fraction and the health
                                                                                                endpoints. No analyses reported with
                                                                                                multiple pollutant models.
                                                                                                           BS was associated with upper
                                                                                                           respiratory symptoms.
                                                                                                Ozone strengthened the observed
                                                                                                associations. Introducing either NO2 or
                                                                                                SO2 in the model did not change the
                                                                                                results markedly.
                                              No difference found for atopic status in
                                              children living in area with different air
                                              pollution levels.
                                                                                                                                                  Lag 0, PM10, Urban areas
                                                                                                                                                   Cough OR = 1.04(0.95,  1.14)
                                                                                                                                                  Lag 2, PM10, Urban areas
                                                                                                                                                   Cough OR = 0.94 (0.89,  1.06)
                                                                                                                                                  5 day ave, PM10, Urban areas
                                                                                                                                                   Cough OR = 0.95 (0.80,  1.13)
Lag 0, Symptoms, Urban areas
 LRS OR = 0.98 (0.89, 1.08)
 URS OR = 1.04(0.96, 1.14)
Lag 2, Symptoms, Urban areas
 LRS OR = 1.01(0.93, 1.10)
 URS OR = 1.04(0.96, 1.13)
5 day ave, Symptoms, Urban areas
 LRS OR = 0.95 (0.82, 1.11)
 URS OR = 1.17(1.00, 1.37)

Lag 0, PM10:
 Cough OR = 1.00 (0.87, 1.16)
4 day ave, PM10
 Cough OR = 1.58(0.87,2.83)
Lag 0, PM2 5
 Cough OR = 1.04(0.88, 1.23)
4 day ave., PM25
 Cough OR = 2.01 (1.04,3.89)

-------
 to
 o
 o
 to
 TABLE 8B-7 (cont'd).  SHORT-TERM PARTICIPATE MATTER EXPOSURE EFFECTS ON SYMPTOMS
	IN STUDIES OF NONASTHMATICS	

                                                                                                                      Effect measures standardized to 50 ,ug/m3 PM10
                                                                                                                      (25 fj.g/m1 PM2.5). Negative coefficients for
                                                                                                                      lung function and ORs greater than 1 for other
                                                                                                                      endpoints suggest PM effects
           Reference citation, location, duration,
           pollutants measured, summary of values
Type of study, sample size, health outcomes
measured, analysis design, covariates included,
analysis problems, etc.
Results and Comments
Effects of co-pollutants
 OO
 td
 ij
 to
          New Zealand

          Harreetal. (1997)
          Christchurch, NZ
          SO2, NO2, PM10, and CO measured.
          Details on monitoring methods and pollutant ranges
          were not given.

          Asia

          Awasthietal. (1996)
          India
          Suspended particulate matter, SO2, nitrates, coal,
          wood, PM and kerosene measured. SPM was
          measured using a high-volume sampler.	
                                      Study of 40 subjects aged 55 years with COPD
                                      living in Christchurch, New Zealand during
                                      winter 1994. Subjects recorded completed diaries
                                      twice daily. Poisson regression model used to
                                      analyze symptom data.
                                      A cohort of 664 preschool children studied for
                                      two weeks each in northern India.  Ordinary least
                                      squares was used to relate a respiratory symptom
                                      complex pollutants.
                                            NO2 was associated with increased
                                            bronchodilator use.
                                     PM10 was associated with increased nighttime
                                     chest symptoms.
                                            A significant regression coefficient
                                            between PM and symptoms was found
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
o

-------
          Appendix 8B.8: Long-Term PM Exposure Effects On
       Respiratory Health Indicators, Symptoms, and Lung Function
April 2002                         8B-73     DRAFT-DO NOT QUOTE OR CITE

-------
to
o
o
to
 TABLE 8B-8. LONG-TERM PARTICIPATE MATTER EXPOSURE RESPIRATORY HEALTH INDICATORS:
	RESPIRATORY SYMPTOM, LUNG FUNCTION	

                                                                                                                   Effect estimates as reported by study
                                                                                                                   authors. Negative coefficients for lung
                                                                                                                   function and ORs greater than 1  for other
                                                                                                                   endpoints suggest effects of PM
         Reference citation, location, duration, type of
         study, sample size, pollutants measured,
         summary of values
Health outcomes measured, analysis design,
covariates included, analysis problems
Results and Comments
Effects of co-pollutants
          United States

          Abbey etal. (1998)
          California Communities
          20 year exposure to respirable particulates,
          suspended sulfates, ozone, and PM10.
          PM10 ranged from 1 to 145 ,ug/rn3 with a mean
          value of 32.8.
                                     Sex specific multiple linear regressions were
                                     used to relate lung function measures to
                                     various pollutants in long-running cohort study
                                     of Seven Day Adventists (ASHMOG Study).
                                            Sulfates were associated with
                                            decreases in FEV.
                                   Frequency of days where PM10 >
                                   100 Aig/m3 associated with FEV
                                   decrement in males whose parents had
                                   asthma, bronchitis, emphysema, or hay
                                   fever.  No effects seen in other subgroups.
oo
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
         Berglundetal.. (1999)
         California communities
         Peters etal. (1999a,b)
         12 demographically similar communities in
         So. California.
         O3, PM acids, and NO2 evaluated.
         PM was measured using a tapered element
         oscillating microbalance instrument.

         Avol etal. (2001)
         Subjects living in Southern California in 1993
         that moved to other western locations in 1998.
         Pollutants O3, NO2, PM10 differences 15 to
         66 ,ug/m3.
                                     Cohort study of Seventh Day Adventists.
                                     Multivariate logistic regression analysis of risk
                                     factors (e.g., PM) for chronic airway disease in
                                     elderly non-smokers, using pulmonary function
                                     test and respiratory symptom data.

                                     Stepwise logistic regression was used to relate
                                     prevalence rates for symptoms to community-
                                     specific ambient pollutants after adjustment for
                                     race, sex, asthma, body mass, hay fever, and
                                     membership in an insurance plan.


                                     Studied 110 children who were 10 yrs of age at
                                     enrollment and 15 at follow-up who had
                                     moved from communities filled out health
                                     questions and underwent spirometry.  Linear
                                     regression used to determine whether annual
                                     average change in lung function correlated
                                     with average changes in PM.
                                            Significant risk factors identified:
                                            childhood respiratory illness,
                                            reported ETS exposure, age, sex and
                                            parental history.
                                            Wheeze prevalence was associated
                                            with both acid and NO,.
                                            As a group, subjects who moved to
                                            areas of lower PM10 showed
                                            increased growth in lung function
                                            and subjects who moved to
                                            communities with a higher PM10
                                            showed decreased growth in lung
                                            function.
                                   For PM10 > 100Mg/m3, 42 d/yr:
                                   RR = -1.09 CT (0.92, 1.30) for
                                   obstructive disease determined by
                                   pulmonary function tests.
                                   No significant relationships were found
                                   between PM10 and symptoms.
                                   PM10 24 hr average
                                   PERFml/sperlOMg/m3
                                   mean = -34.9
                                   95% CI
                                   -59.8,-10.1
O

-------
 to
 o
 o
 to
 TABLE 8B-8 (cont'd).  LONG-TERM PARTICIPATE MATTER EXPOSURE RESPIRATORY HEALTH INDICATORS:
	RESPIRATORY SYMPTOM, LUNG FUNCTION	

                                                                                                                         Effect estimates as reported by study
                                                                                                                         authors. Negative coefficients for lung
                                                                                                                         function and ORs greater than 1  for other
                                                                                                                         endpoints suggest effects of PM
          Reference citation, location, duration, type of
          study, sample size, pollutants measured,
          summary of values
Health outcomes measured, analysis design,
covariates included, analysis problems
Results and Comments
Effects of co-pollutants
 oo
 td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          United States (cont'd)

          Gauderman et al. (2000)
          12 So. California communities 1993 to 1997
          Pollutants:  O3, NO2, PM10, and PM25.
          PM10 levels ranged from 16.1 to 67.6 Mg/m3
          across the communities.
          McConnell et al. (1999)
          12 Southern California communities
          1994 air monitoring data.
          PM10 (mean 34.8; range 13.0 - 70.
          PM2 5 (yearly mean 2 week averaged mean
          15.3 /-ig/m3', range 6.7 - 31.5 ,ug/m3).
                                          Studies of lung function growth of 3035
                                          children in 12 communities within 200-mile
                                          radius of Los Angeles during 1993 to 1997.
                                          Cohorts of fourth, seventh, and tenth-graders
                                          studied. By grade cohort, a sequence of linear
                                          regression models were used to determine over
                                          the 4yr of follow-up, if average lung function
                                          growth rate of children was associated with
                                          average pollutant levels.  Adjustment were
                                          made for height, weight, body mass index,
                                          height by age interaction, report of asthma
                                          activity or smoking. Two-pollutant models
                                          also used.

                                          Cross-sectional study of 3,676 school children
                                          whose parents completed questionnaires in
                                          1993 that characterized the children's history
                                          of respiratory illness.  Three groups examined:
                                          (1) history of asthma; (2) wheezing but no
                                          asthma; and (3) no history of asthma or
                                          wheezing. Logistic regression model used to
                                          analyze PM, O3, NO2, acid vapor effects. This
                                          study also described in Peters et al. (1999b,c).
                                            Lung growth rate for children in
                                            most polluted community, as
                                            compared to least polluted, was
                                            estimated to result in cumulative
                                            reduction of 3.4% in FEV! and
                                            5.0% in MMEF over 4-yr study
                                            period. Estimated deficits mostly
                                            larger for children spending more
                                            time outdoors. Due to the high
                                            correlation in concentrations across
                                            communities, not able to separate
                                            effects of each pollutant.  No sig.
                                            associations seen with O3.

                                            Positive association between air
                                            pollution and bronchitis and phlegm
                                            observed only among children with
                                            asthma.  As PM10 increased across
                                            communities, a corresponding
                                            increase in risk of bronchitis per
                                            interquartile range occurred.
                                            Strongest association with phlegm
                                            was for NO2.  Because of high
                                            correlation of PM air pollution,
                                            NO2, and acid, not possible to
                                            distinguish clearly which most
                                            likely responsible for effects.
                                   From the lowest to highest observed
                                   concentration of each pollutant, the
                                   predicted differences in annual growth
                                   rates were:  -0.85% for PM10 (p = 0.026);
                                   -0.64% for PM25 (p = 0.052); -0.90% for
                                   PM10.2.5 (p = 0.030); -0.77% forNO2 (p =
                                   0.019); and -0.73% for inorganic acid
                                   vapor (p = 0.042).
                                   PM10
                                    Asthma
                                        Bronchitis 1.4 CI( 1.1 -  1.8
                                        Phlegm     2.1(1.4-3.3)
                                        Cough     1.1(0.8- 1.7)
                                    No Asthma / No Wheeze
                                        Bronchitis 0.7 (0.4 - 1.0)
                                        Phlegm     0.8(0.6- 1.3)
                                        Cough     0.9(0.7- 1.2)
O

-------
to
o
o
to
 TABLE 8B-8 (cont'd).  LONG-TERM PARTICIPATE MATTER EXPOSURE RESPIRATORY HEALTH INDICATORS:
	RESPIRATORY SYMPTOM, LUNG FUNCTION	

                                                                                                                         Effect estimates as reported by study
                                                                                                                         authors.  Negative coefficients for lung
                                                                                                                         function and ORs greater than 1 for other
                                                                                                                         endpoints suggest effects of PM
         Reference citation, location, duration, type of
         study, sample size, pollutants measured,
         summary of values
Health outcomes measured, analysis design,
covariates included, analysis problems
                                                                                                 Results and Comments
                                                                                                 Effects of co-pollutants
oo
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          United States (cont'd)

          McConnell et al. (2002)
          12 Southern California communities
          1994-1997
          4-year mean cone. PM10 Mg/m3
          High community: 43.3(12.0)
          Low community: 21.6(3.8)
         Dockeryetal. (1996)
         24 communities in the U. S. and Canada.
         PM10, PM25, sulfate fraction, H+, ozone, SO2,
         and other measures of acid were monitored.
         PM was measured using a Harvard impactor.
         PM10 ranged form 15.4 to 32.7 with a mean of
         23.8. PM25 ranged form 5.8 to 20.7 Mg/m3
         with a mean of 14.5.

         Raizenne et al. (1996)
         24 communities in the U.S. and Canada
         Pollutants measured for at least one year prior
         to lung function tests: PM10, PM2l, particle
         strong acidity, O3, NO2, and SO2.  PM was
         measured with a Harvard impactor. For
         pollutant ranges, see Dockery et al. (1996).
                                          In 3,535 children assessed, the association of
                                          playing team sports with subsequent
                                          development of asthma during 4 yrs of follow-
                                          up.  Comparing high pollutant communities to
                                          low pollutant communities.  Relative risks of
                                          asthma adjusted for ethnic origin were
                                          evaluated for every pollutant with a
                                          multivariate proportional hazards model.  See
                                          also Peters et al. (1999b,c).
                                          Respiratory health effects among 13,369 white
                                          children aged 8 to 12 yrs analyzed in relation
                                          to PM indices.  Two-stage logistic regression
                                          model used to adjust for gender, history of
                                          allergies, parental asthma, parental education,
                                          smoking in home.
                                          Cross-sectional study of lung function. City
                                          specific adjusted means for FEV and FVC
                                          calculated by regressing the natural logarithm
                                          of the measure on sex, In height, and In age.
                                          These adjusted means were then regressed on
                                          the annual pollutant means for each city.
                                            Across all communities there was a
                                            1.8-fold increased risk (95% CI
                                            1.2-2.8) for asthma in children who
                                            had played three or more team
                                            sports in the previous year.  In high
                                            ozone (10:00 h to 18:00 h mean
                                            concentration) communities, there
                                            was a 3.3-fold increase risk of
                                            asthma in children playing three or
                                            more sports, an increase not seen in
                                            low ozone communities.

                                            Although bronchitis endpoint was
                                            significantly related to fine PM
                                            sulfates, no endpoints were related
                                            to PM10 levels.
                                            PM measures (e.g., particle strong
                                            acidity) associated with FEV and
                                            FVC decrement.
                                                                                                                                    The effect of team sports was similar in
                                                                                                                                    communities with high and low PM with
                                                                                                                                    a small increase in asthma among children
                                                                                                                                    playing team sports.
O

-------
to
o
o
to
  TABLE 8B-8 (cont'd).  LONG-TERM PARTICIPATE MATTER EXPOSURE RESPIRATORY HEALTH INDICATORS:
 	RESPIRATORY SYMPTOM, LUNG FUNCTION	

                                                                                                                         Effect estimates as reported by study
                                                                                                                         authors. Negative coefficients for lung
                                                                                                                         function and ORs greater than 1 for other
                                                                                                                         endpoints suggest effects of PM
         Reference citation, location, duration, type of
         study, sample size, pollutants measured,
         summary of values
Health outcomes measured, analysis design,
covariates included, analysis problems
Results and Comments
Effects of co-pollutants
oo
td
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
         Europe

         Ackermann-Liebrich et al. (1997)
         Eight Swiss regions
         Pollutants: SO2, NO2, TSP, O3, and PM10.
         PM was measured with a Harvard impactor.
PM10 ranged from 10 to 53 /
of37.
                                      3 with a mean
Braun-Fahrlander et al. (1997)
10 Swiss communities
Pollutants: PM10, NO2, SO2, and O3.
PM was measured with a Harvard impactor.
PM10 ranged from 10 to 33 ,ug/m3.
         Zempetal. (1999)
         8 study sites in Switzerland.
         Pollutants: TSP, PM10, SO2, NO2, and O3.
         PM was measured with a Harvard impactor.
         PM10 ranged from 10 to 33 ,ug/rn3 with a mean
         of21.
Long-term effects of air pollution studied in
cross-sectional population-based sample of
adults aged 18 to 60 yrs. Random sample of
2,500 adults in each region drawn from
registries of local inhabitants.  Natural
logarithms of FVC and FEV[ regressed against
natural logarithms of height, weight, age,
gender, atopic status, and pollutant variables.

Impacts of long-term air pollution exposure  on
respiratory symptoms and illnesses were
evaluated in cross-sectional study of Swiss
school children, (aged 6 to 15  years).
Symptoms  analyzed using a logistic regression
model  including covariates of family history
of respiratory and allergic diseases, number of
siblings, parental education, indoor fuels,
passive smoking, and others.

Logistic regression analysis of associations
between prevalences of respiratory symptoms
in random sample of adults and air pollution.
Regressions adjusted for age, BMI, gender,
parental asthma, education, and foreign
citizenship.
                                                                                       Significant and consistent effects on
                                                                                       FVC and FEV were found for PM10,
                                                                                       NO, and SO,.
Respiratory endpoints of chronic
cough, bronchitis, wheeze and
conjunctivitis symptoms were all
related to the various pollutants.
The colinearity of the pollutants
including NO2, SO2, and O3,
prevented any causal separation.
                                                                                       Chronic cough and chronic phlegm
                                                                                       and breathlessness were related to
                                                                                       TSP,PM10andN02.
                                  Estimated regression coefficient for PM10
                                  versus FVC = -0.035 (95% CI -0.041,
                                  -0.028). Corresponding value for FEV[
                                  -0.016 (95% CI -0.023 to -0.01).  Thus,
                                  10 Mg/ni3 PM10 increase estimated to lead
                                  to estimated 3.4 percent decrease in FVC
                                  and 1.6 percent decrease in FEV[.
                                                                                                                                   PM10
                                                                                                                                   Chronic cough OR 11.4 (2.8, 45.5)
                                                                                                                                   Bronchitis OR 23.2 (2.8, 45.5)
                                                                                                                                   Wheeze OR 1.41 (0.55, 3.58)
                                   Chronic cough, chronic phlegm and
                                   breathlessness were related to PM10, and
                                   TSP.
O

-------
to
o
o
to
  TABLE 8B-8 (cont'd). LONG-TERM PARTICIPATE MATTER EXPOSURE RESPIRATORY HEALTH INDICATORS:
 	RESPIRATORY SYMPTOM,  LUNG FUNCTION	
                                                                                                                          Effect estimates as reported by study
                                                                                                                          authors. Negative coefficients for lung
                                                                                                                          function and ORs greater than 1  for other
                                                                                                                          endpoints suggest effects of PM
Reference citation, location, duration, type of
study, sample size, pollutants measured,
summary of values
                                                     Health outcomes measured, analysis design,
                                                     covariates included, analysis problems
                                                                                                 Results and Comments
                                                                                                 Effects of co-pollutants
oo
td
ij
oo
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
         Europe (cont'd)

         Heinrich et al. (1999)
         Bitterfeld, Zerbstand Hettstedt areas of former
         East Germany,
         During Sept.  1992 to July 1993 TSP ranged
         from 44 to 65 /^ig/m3',
         PM10 measured October 1993 - March 1994
         ranged from 33 to 40; and BS ranged from
         26 to 42 ,ug/m3. PM was measured with a
         Harvard impactor.
Heinrich et al. (2000)
Three areas of former E. Germany
Pollution measures:  SO2, TSP, and some
limited PM10 data. TSP decreased from 65,
48, and 44 ,ug/m3 to 43, 39, and 36 Mg/m3 in
the three areas. PM was measured with a
Harvard impactor.
         Kramer etal. (1999)
         Six East and West Germany communities
         (Leipzig, Halle, Maddeburg, Altmark,
         Duisburg, Borken)
         Between 1991  and 1995 TSP levels in six
         communities ranged from 46 to 102 ,ug/m3.
         Each East Germany community had decrease
         in TSP between 1991 and 1995.  TSP was
         measured using a low volume sampler.
                                           Parents of 2470 school children ( 5-14 yr)
                                           completed respiratory health questionnaire.
                                           Children excluded from analysis if had lived
                                           < 2 years in their current home, yielding an
                                           analysis group of 2,335 children. Outcomes
                                           studied: physician diagnosis for asthma,
                                           bronchitis,  symptom, bronchial reactivity, skin
                                           prick test, specific IgE. Multiple logistic
                                           regression analyses examined regional effects.
                                           Cross-sectional study of children (5-14 yr).
                                           Survey conducted twice, in 1992-1993 and
                                           1995-1996; 2335 children surveyed in first
                                           round, and 2536 in second round. Only 971
                                           children appeared in both surveys. The
                                           frequency of bronchitis, otitus media, frequent
                                           colds, febrile infections studied.  Because
                                           changes measured over time in same areas,
                                           covariate adjustments not necessary.

                                           The study assessed relationship between TSP
                                           and airway disease and allergies by parental
                                           questionnaires in yearly surveys of children
                                           (5-8 yr) between February and May. The
                                           questions included pneumonia, bronchitis ever
                                           diagnosed by physician, number of colds,
                                           frequent cough, allergic symptoms.
                                           In all, 19,090 children participated. Average
                                           response was 87%. Analyses were conducted
                                           on 14,144 children for whom information on
                                           all covariates were available.  Variables
                                           included gender; parent education, heating
                                           fuel, ETS. Logistic regression used to allow
                                           for time trends and SO2 and TSP effects.
                                           Regression coefficients were converted to odds
                                           ratios.
                                                                                                 Controlling for medical, socio-
                                                                                                 demographic, and indoor factors,
                                                                                                 children in more polluted area had
                                                                                                 circa 50% increase for bronchitic
                                                                                                 symptoms and physician-diagnosed
                                                                                                 allergies compared to control area
                                                                                                 and circa twice the respiratory
                                                                                                 symptoms (wheeze, shortness of
                                                                                                 breath and cough).  Pulmonary
                                                                                                 function tests suggested slightly
                                                                                                 increased airway reactivity to cold
                                                                                                 for children in polluted area.

                                                                                                 PM and SO2 levels both decreased
                                                                                                 in the same areas; so results are
                                                                                                 confounded.
                                                                                       TSP and SO2 simultaneously
                                                                                       included in the model. Bronchitis
                                                                                       ever diagnosed showed a significant
                                                                                       association.  A decrease in raw
                                                                                       percentage was seen between the
                                                                                       start of the study and the end for
                                                                                       bronchitis. Bronchitis seemed to be
                                                                                       associated only with TSP in spite of
                                                                                       huge differences in mean SO2
                                                                                       levels.
                                                                                                                          No single pollutant could be separated out
                                                                                                                          as being responsible for poor respiratory
                                                                                                                          health.
The prevalence of all respiratory
symptoms decreased significantly in all
three areas over time.
Bronchitis ever diagnosed
TSP per 50 /-ig/m3
  OR 1.63 CI (1.37- 1.93)
  Halle (East)            %
      TSP Mg/m3    Bronchitis
1991     102           60.5
1992      73           54.7
1993      62           49.6
1994      52           50.4
1995      46           51.9

-------
to
o
o
to
  TABLE 8B-8 (cont'd).  LONG-TERM PARTICIPATE MATTER EXPOSURE RESPIRATORY HEALTH INDICATORS:
 	RESPIRATORY SYMPTOM, LUNG FUNCTION	

                                                                                                                         Effect estimates as reported by study
                                                                                                                         authors. Negative coefficients for lung
                                                                                                                         function and ORs greater than 1 for other
                                                                                                                         endpoints suggest effects of PM
         Reference citation, location, duration, type of
         study, sample size, pollutants measured,
         summary of values
Health outcomes measured, analysis design,
covariates included, analysis problems
Results and Comments
Effects of co-pollutants
         Europe (cont'd)
oo
td
ij
VO
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
         Baldietal. (1999)
         24 areas of seven French towns 1974-1976
         Pollutants: TSP, BS, and SO2, NO4
         3-year average TSP-mean annual values
         ranging 45-243 ^g/m3. TSP was measured by
         the gravimetric method.
Zeghnounetal. (1999)
La Havre, France during 1993 and 1996.
Daily mean BS levels measured in three
stations ranged 12-14 ^g/m3.
         Leonardi et al. (2000)
         17 cities of Central Europe
         Yearly average concentration (Nov. 1995 -
         Oct. 1996) across the 17 study areas varied
         from 41 to 96 /j,g/m3 for PM10, from 29 to 67
         Mg/m3 for PM25, and from 12 to 38 ,ug/rn3 for
         PM ,„.,,.
Reanalysis of Pollution Atmospheric of
Affection Respiratory Chroniques (PAARC)
survey data to search for relationships between
mean annual air pollutant levels and
prevalence of asthma in 1291 adult (25-59 yrs)
and 195 children (5-9 yrs) asthmatics. Random
effects logistic regression model used and
included age, smoking, and education level in
the final model.

Respiratory drug sales for mucolytic and
anticough medications (most prescribed by a
physician) were evaluated versus BS, SO2, and
NO2 levels.  An autoregressive Poisson
regression model permitting overdispersion
control was used in the analysis.
                                           Cross-sectional study collected blood and
                                           serum samples from 10-61 school children
                                           aged 9 to 11 in each community 11 April to
                                           May 1996. Blood and serum samples
                                           examined for parameters in relation to PM.
                                           Final analysis group of 366 examined for
                                           peripheral lymphocyte type and total
                                           immunoglobulin classes.  Association between
                                           PM and each log transformed biomarker
                                           studied by linear regression in two-stage
                                           model with adjustment for confounding factors
                                           (age, gender, number of smokers in house,
                                           laboratory, and recent respiratory illness). This
                                           survey was conducted within the frame work of
                                           the Central European study of Air Quality and
                                           Respiratory Health (CEASAR) study.
                                                                                      Only an association between SO2
                                                                                      and asthma in adults observed. No
                                                                                      other pollutant was associated. Nor
                                                                                      was relationship with children seen.
                                                                                      Meteorological variables and O3
                                                                                      not evaluated.
Respiratory drug sales associated
with BS, NO2, and SO2 levels. Both
an early response (0 to 3 day lag)
and a longer one (lags of 6 and
9 days) were associated.
                                            Number of lymphocytes (B, CD4+,
                                            CD8d, and NK) increased with
                                       10   increasing concentration of PM
                                            adjusted for confounders. The
                                            adjusted regression slopes are
                                            largest and statistically significant
                                            for PM2 5 as compared to PM10, but
                                            small and non statistically signif.
                                            for PM10_25.  Positive relationship
                                            found between concentration of IgG
                                            in serum and PM2 5 but not for PM10
                                            or PM10_25. Two other models
                                            produced  similar outcomes:  a
                                            multi-level linear regression model
                                            and an ordinal logistic regression
                                            model.
                                  For a 50 /j.g/w? increase in
                                  TSP
                                     Adult asthma prevalence
                                        OR 1.01 CI 0.92-1.11
                                  SO2
                                     Adult asthma prevalence
                                        OR 1.26 CI 1.04-1.53
                                  Adjusted
                                    Regression slope
                                      PM,.
                                      CD4+
                                      80% 95% CI (34; 143)
                                      p< 0.001

                                    Total IgG
                                      24%
                                      95% CI (2; 52)
                                      p 0.034

-------
to
o
o
to
  TABLE 8B-8 (cont'd). LONG-TERM PARTICIPATE MATTER EXPOSURE RESPIRATORY HEALTH INDICATORS:
 	RESPIRATORY SYMPTOM, LUNG FUNCTION	

                                                                                                                          Effect estimates as reported by study
                                                                                                                          authors. Negative coefficients for lung
                                                                                                                          function and ORs greater than 1 for other
                                                                                                                          endpoints suggest effects of PM
         Reference citation, location, duration, type of
         study, sample size, pollutants measured,
         summary of values
Health outcomes measured, analysis design,
covariates included, analysis problems
Results and Comments
Effects of co-pollutants
oo
td
oo
o
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
         Europe (cont'd)

         Tumovska and Kostiranev (1999)
         Dimitrovgrad, Bulgaria, May 1996
         Total suspended particulate matter (TSPM)
         mean levels were 520 ± I 61 Aig/m3 in 1986
         and 187 ± 9 Mg/m3 in 1996.  SO2, H2S, and
         NO, also measured.
Jedrychowski et al. (1999)
In Krakow, Poland in 1995 and 1997
Spacial distributions for BS and SO2 derived
from network of 17 air monitoring stations.
BS 52.6 Mg/m ±53.98 in high area and
33.23 ±35.99 in low area.
         Jedrychowski and Flak (1998)
         In Kracow Poland, in 1991-1995
         Daily 24 h concentration of SPM (black
         smoke) measured at 17 air monitoring
         stations.
         High areas had 52.6 /-ig/m3 mean compared to
         low areas at 33.2 ^g/m3.
Respiratory function of 97 schoolchildren
(mean age 10.4 ± 0.6 yr) measured in May
1996 as a sample of 12% of all four-graders in
Dimitrovgrad. The obtained results were
compared with reference values for Bulgarian
children aged 7 to 14 yr, calculated in the same
laboratory in 1986 and published
(Gherghinova et al., 1989; Kostianev et al.,
1994).  Variation analysis technique were used
to treat the data.

Effects on lung function growth studied in
preadolescent children. Lung function growth
rate measured by gain in FVC and FEV[ and
occurrence of slow lung function growth
(SLFG) over the 2 yr period defined as lowest
quintile of the distribution of a given test in
gender group. 1129 children age 9 participated
in first year and  1001 in follow-up 2 years
later. ATS standard questionnaire and PFT
methods used. Initially univariate descriptive
statistics of pulmonary function indices and
SLFG were established, followed by
multivariate linear regression analyses
including gender, ETS, parental education,
home heating system and mold. SO2 also
analyzed.

Respiratory health survey of 1,129 school
children (aged 9 yr).  Respiratory outcomes
included chronic cough, chronic phlegm,
wheezing, difficulty breathing and asthma.
Multi-variable logistic regression used to
calculate prevalence OR for symptoms
adjusted for potential confounding.
                                                                                       Vital capacity and FEV[ were
                                                                                       significantly lower (mean value. =
                                                                                       88.54% and 82.5% respectfully)
                                                                                       comparing values between 1986 and
                                                                                       1996. TSPM pollution had
                                                                                       decreased by 2.74 times to levels
                                                                                       still higher than Bulgarian and
                                                                                       WHO standards.
Statistically significant negative
association between air pollution
level and lung function growth
(FVC and FEVO over the follow up
in both gender groups.  SLFG was
significantly higher in the more
polluted areas only among boys.
In girls there was consistency in the
direction of the effect, but not stat.
significant. Could not separate BS
and SO2 effects on lung function
growth.  Excluding asthma subjects
subsample (size 917) provided
similar results.
                                                                                       The comparison of adjusted effect
                                                                                       estimates revealed chronic phlegm
                                                                                       as unique symptom related neither
                                                                                       to allergy nor to indoor variable but
                                                                                       was associated significantly with
                                                                                       outdoor air pollution category
                                                                                       (APL). No potential confounding
                                                                                       variable had major effect.
                                                                                                                                    Boys
                                                                                                                                      SLFG (FVC)
                                                                                                                                         OR = 2.15 (CI 1.25 - 3.69)
                                                                                                                                      SLFG (FEVO
                                                                                                                                         OR=1.90(CI1.12- 3.25)

                                                                                                                                    Girls
                                                                                                                                      FVC OR =1.50 (CI 0.84-2.68)
                                                                                                                                      FEV1 OR= 1.39 (CI 0.78 - 2.44)
                                   It was not possible to assess separately the
                                   contribution of the different sources of air
                                   pollutants to the occurrence of respiratory
                                   symptoms. ETS and household heating
                                   (coal vs. gas vs. central heating) appeared
                                   to be of minimal importance.

-------
to
o
o
to
  TABLE 8B-8 (cont'd).  LONG-TERM PARTICIPATE MATTER EXPOSURE RESPIRATORY HEALTH INDICATORS:
 	RESPIRATORY SYMPTOM, LUNG FUNCTION	

                                                                                                                         Effect estimates as reported by study
                                                                                                                         authors. Negative coefficients for lung
                                                                                                                         function and ORs greater than 1 for other
                                                                                                                         endpoints suggest effects of PM
Reference citation, location, duration, type of
study, sample size, pollutants measured,
summary of values
                                                    Health outcomes measured, analysis design,
                                                    covariates included, analysis problems
                                                                                                Results and Comments
                                                                                                Effects of co-pollutants
oo
td
oo
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
         Latin America

         Calderon-Garciduenas et al. (2000)
         Southwest Metropolitan Mexico City
         (SWMMC) winter of 1997 and summer of
         1998.
Australia

Lewis etal. (1998)
Summary measures of PM10 and SO2
estimated for each of 10 areas in steel cities of
New South Wales.  PM10 was measured using
a high volume sampler with size-selective
inlets.
         Asia

         Wong etal. (1999)
         Hong Kong, 1989 to 1991
         Sulfate concentrations in respirable particles
         fell by 38% after implementing legislation
         reducing fuel sulfur levels.
                                           Study of 59 SWMMC children to evaluate
                                           relationship between exposure to ambient
                                           pollutants (O3 and PM10) and chest x-ray
                                           abnormalities. Fishers exact test used to
                                           determine significance in a 2x2 task between
                                           hyperinflation and exposure to SWMMC
                                           pollutant atmosphere and to control, low-
                                           pollutant city atmosphere.
                                           Cross-sectional survey of children's health and
                                           home environment between Oct 1993 and Dec
                                           1993 evaluated frequency of respiratory
                                           symptoms (night cough, chest colds, wheeze,
                                           and diagnosed asthma). Covariates included
                                           parental education and smoking, unflued gas
                                           heating, indoor cats, age, sex, and maternal
                                           allergy. Logistic regression analysis used
                                           allowing for clustering by GEE methods.
                                           3405 nonsmoking, women (mean age 36.5 yr;
                                           SD ± 3.0) in a polluted district and a less
                                           polluted district were studied for six
                                           respiratory symptoms via self-completed
                                           questionnaires. Binary latent variable
                                           modeling used.
                                                                                      Bilateral symmetric mild lung
                                                                                      hyperinflation was significantly
                                                                                      associated with exposure to the
                                                                                      SWMMC air pollution mixture
                                                                                      (p>0.0004). This raises concern for
                                                                                      development of chronic disease
                                                                                      outcome in developing lungs.
                                                                                                SO2 was not related to differences in
                                                                                                symptom rates, but adult indoor
                                                                                                smoking was.
Night cough OR 1.34 (1.18, 1.53)
Chest colds OR 1.43 (1.12, 1.82)
Wheeze OR 1.13 (0.93, 1.38)
                                                                                       Comparison was by district; no PM
                                                                                       measurements reported. Results
                                                                                       suggest control regulation may have
                                                                                       had some (but not statistically
                                                                                       significant) impact.
O

-------
to
o
o
to
  TABLE 8B-8 (cont'd). LONG-TERM PARTICIPATE MATTER EXPOSURE RESPIRATORY HEALTH INDICATORS:
 	RESPIRATORY SYMPTOM, LUNG FUNCTION	

                                                                                                                          Effect estimates as reported by study
                                                                                                                          authors.  Negative coefficients for lung
                                                                                                                          function and ORs greater than 1 for other
                                                                                                                          endpoints suggest effects of PM
         Reference citation, location, duration, type of
         study, sample size, pollutants measured,
         summary of values
Health outcomes measured, analysis design,
covariates included, analysis problems
Results and Comments
Effects of co-pollutants
oo
td
oo
to
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
 O
         Asia (cont'd)

         Wang etal. (1999)
         Kaohsiung and Panting, Taiwan
         October 1995 to June 1996
         TSP measured at 11 stations, PM10 at 16
         stations.  PM10 annual mean ranged from 19.4
         to 112.81 ,ug/m3 (median =91.00 Mg/m3)
         TSP ranged from 112.81 to 237.82 Mg/m3
         (median = 181.00). CO,NO2, SO2,
         hydrocarbons and O3 also measured.
Guo etal. (1999)
Taiwan, October 1955 and May 1996
PM10 measured by beta-gauge.
Also monitoring for SO2, NO2,  O3, CO.
PM10 ranged from 40 to  110 Mg/ni3 with a
mean of 69.
         Wang etal. (1999)
         Chongquing, China
         April to July 1995
         Dichot samplers used to measure PM25.
         Mean PM2 5 level high in both urban
         (143 ^ig/m3) and suburban (139 /L/g/m3) area.
         SO2 also measured
Relationship between asthma and air pollution
examined in cross-sectional study among
165,173 high school students (11-16 yr).
Evaluated wheeze, cough and asthma
diagnosed by doctor.  Video determined if
student displayed signs of asthma.  Only
155,283 students met all requirements for
study analyses and, of these, 117,080 were
covered by air monitoring stations. Multiple
logistic regression analysis used to determine
independent effects of risk factors for asthma
after adjusting for age, gender, ETS, parents
education, area resident, and home incense use.

Study of asthma prevalence and air pollutants.
Survey for respiratory disease and symptoms in
middle-school students age < 13 to > 15 yr.
Total of 1,018,031 (89.3%) students and their
parents responded satisfactorily to the
questionnaire.  Schools located with 2 km of
55 monitoring sites.  Logistic regression
analysis conducted, controlling for age, hx
eczema, parents education.

Study examined relationship between PFT and
air pollution. Pulmonary function testing
performed on 1,075 adults (35 - 60 yr) who
had never smoked and did not use coal stoves
for cooking. Generalized additive model used
to estimate difference, between two areas for
FEVl5 FVC, and FEVj/FVCro with adjustment
for confounding factors (gender; age, height,
education, passive smoking, and occupational
exposures).
                                                                                       Asthma significantly related to high
                                                                                       levels of TSP, NO2, CO, O3 and
                                                                                       airborne dust. However PM10 and
                                                                                       SO2 not associated with asthma.
                                                                                       The lifetime prevalence of asthma
                                                                                       was 18.5% and the 1-year
                                                                                       prevalence was 12.5%.
Because of close correlation among
air pollutants, not possible to
separate effects of individual ones.
Factor analysis used to group into
two classes (traffic-related and
stationary fossil fuel-related).  No
association found between lifetime
asthma prevalence and nontraffic
related air pollutants (SO2, PM10).

Mean SO2 concentration in the
urban and suburban area highly
statistically significant different
(213 and 103 /-ig/m3 respectfully).
PM2 5 difference was small, while
levels high in both areas.  Estimated
effects on FEV1 statistically
different between the two areas.
                                   Adjusted OR

                                   PM10
                                    1.00(0.96-1.05)

                                   TSP
                                    1.29(1.24-1.34)
                                                                                                                          Difference between urban and suburban
                                                                                                                          area excluding occupational exposures:
                                                                                                                          FEV,
                                                                                                                           B- 119.79
                                                                                                                            SE28.17
                                                                                                                            t - 4.25
                                                                                                                            p<0.01
FVC
 B - 57.89
  SE 30.80
  t- 1.88
  p<0.05

-------
 to
 o
 o
 to
  TABLE 8B-8 (cont'd).  LONG-TERM PARTICIPATE MATTER EXPOSURE RESPIRATORY HEALTH INDICATORS:
 	RESPIRATORY SYMPTOM, LUNG FUNCTION	

                                                                                                                         Effect estimates as reported by study
                                                                                                                         authors. Negative coefficients for lung
                                                                                                                         function and ORs greater than 1 for other
                                                                                                                         endpoints suggest effects of PM
          Reference citation, location, duration, type of
          study, sample size, pollutants measured,
          summary of values
Health outcomes measured, analysis design,
covariates included, analysis problems
Results and Comments
Effects of co-pollutants
 oo
 td
 oo
 fe
 H
 6
 o
 o
 H
O
 O
 H
 W
 O
          Asia (cont'd)

          Zhang etal. (1999)
          4 areas of 3 Chinese Cities (1985 - 1988)
          TSP levels ranged from an annual arithmetic
          mean 137 /-ig/m3 to 1250 Mg/m3 using
          gravimetric methods.
Qian et al. (2000)
4 China cities
The 4 year average TSP means were 191, 296,
406, and 1067 Mg/m3. SO2 and NO2
measurements were also available.
TSP was measured gavimetrically.
A pilot study of 4 districts of 3 Chinese cities
in for the years 1985-1988, TSP levels and
respiratory health outcomes studied.  4,108
adults (< 49 yrs) examined by questionnaires
for couth, phlegm, wheeze, asthma, and
bronchitis. Categorical logistic—regression
model used to calculate odds ratio. SO2 and
NO2 were also examined. Other potential
confounding factors (age, education level,
indoor ventilation, and occupation) examined
in the multiple logistic regression model.

Pilot cross-sectional survey  of 2789 elementary
school children in four Chinese communities
chosen for their PM gradient.  Frequency of
respiratory symptoms (cough, phlegm, wheeze,
and diagnosed asthma, bronchitis, or
pneumonia) assessed by questionnaire.
Covariates included parental occupation,
education and smoking.  The analysis used
logistic regression, controlling for age, sex,
parental smoking, use of coal in home, and
home ventilation.
                                                                                       Results suggested that the OR's for
                                                                                       cough, phlegm, persistent cough
                                                                                       and phlegm and wheeze increased
                                                                                       as outdoor TSP concentrations did. .
                                   Wheeze produced largest OR for both
                                   mothers and fathers in all locations.
Results not directly related to
pollution levels, but symptom rates
were highest in highest pollution
area for cough, phlegm,
hospitalization for respiratory
disease, bronchitis, and pneumonia.
No gradient correlating with
pollution levels found for the three
lower exposure communities.
o

-------
 i                         9.  INTEGRATIVE SYNTHESIS
 2
 3
 4      9.1 INTRODUCTION
 5          This chapter focuses on integration of key information on exposure-dose-response risk
 6      assessment components drawn from the preceding detailed chapters, to provide a coherent
 7      framework for assessment of human health risks posed by ambient particulate matter (PM) in the
 8      United States. As such, the chapter updates the integrated assessment of available scientific
 9      information regarding ambient PM sources, exposures, and health risks as they pertain to the
10      United States that was provided in the 1996 Particulate Matter Air Quality Criteria Document
11      (1996 PM AQCD; U.S. Environmental Protection Agency, 1996a).
12          This chapter mainly uses the 10 Questions from the National Research Council (NRC)
13      Particulate Matter (PM) Research Agenda (NRC, 1998, 2001) as an organizing principle to
14      summarize and integrate key points derived from the material presented in detail in Chapters  1 to
15      8 of this document.  After providing certain background information, the chapter is then basically
16      organized to follow the Risk Assessment Framework (as shown in Figure  9-1), and it addresses
17      the NRC questions noted earlier in Chapter 1 within the context of discussing general topic areas
18      that follow the flow of that framework from sources/emissions to effects.  Some additional topics
19      in addition to the 10 NRC questions are also addressed.
20          Unlike the other criteria  pollutants (O3, CO, NO2, SO2, and Pb), PM  is not a specific
21      chemical entity but is a mixture of particles of different sizes,  compositions,  and properties.
22      Therefore, it is useful to present some background on the size, chemistry and physics of PM
23      before entering the Risk Assessment Framework. Thus, this chapter first provides background
24      information on key features of atmospheric particles, highlighting important distinctions between
25      fine- and coarse-mode particles with regard to size, chemical composition, sources, atmospheric
26      behavior, and potential human exposure relationships—distinctions that collectively continue to
27      suggest that fine- and coarse-mode particles should be treated as two distinct subclasses of air
28      pollutants. Recent trends in U.S. concentrations of different ambient PM size and composition
29      fractions (e.g., PM10, PM2 5, and PM10_2 5) and ranges of variability seen in U.S. regions and urban
30      airsheds are also summarized to place the ensuing human exposure and health effects discussions

        April 2002                                 9-1         DRAFT-DO NOT QUOTE OR CITE

-------
          Sources of Airborne
           Particulate Matter
             Or Gaseous
          Precursor Emissions
    Indicator in
Ambient (Outdoor) Air
    (e.g. Mass
   Concentration)
                    Mechanism determining emissions,
                    chemical transformation (including
                   formation of secondary particles from
                      gaseous precursors), and
                         transport in air
               Human time-activity
               patterns, Indoor (or
               microenvironmental)
               sources and sinks of
                particuiate matter
   Deposition,
 clearance, retention
 and disposition of
 particuiate matter
  presented to an
   individual
 Mechanisms of
damage and repair
        Figure 9-1.  A general framework for integrating particulate-matter research. Note that
                     this figure is not intended to represent a framework for research management.
                     Such a framework would include multiple pathways for the flow of
                     information.

        Source: National Research Council (2001), as modified from NRC (1983, 1994), Lioy (1990), and Sexton et al.
               (1992).
 1      in perspective. After discussing human exposure aspects, the chapter next summarizes key points

 2      regarding respiratory tract dosimetry, followed by a discussion of the extensive PM health

 3      database that has expanded greatly during recent years. The latter includes numerous new

 4      epidemiologic studies of populations throughout the world published since the  1996 PM AQCD

 5      that provide further evidence that serious health effects (mortality, exacerbation of chronic

 6      disease, increased hospital admissions, etc.) are  associated with exposures to ambient levels of

 7      PM found in contemporary U.S. urban air sheds. Evaluations of other possible explanations for

 8      the reported PM epidemiology results (e.g., other co-pollutants, choice of models, etc.) also are

 9      discussed, ultimately leading to the conclusion that the reported associations of PM exposure and

10      effects are valid.

11           New toxicologic evidence (derived from controlled exposure studies of humans and

12      laboratory animals) is  also discussed, which elucidates likely mechanisms of action and other

13      information that greatly enhances the plausibility of the epidemiologic findings in comparison to

14      1996. Quantitative evidence is then discussed that (a) further substantiates associations of such

15      serious health effects with U.S. ambient PM10 levels, (b) also more strongly establishes fine
        April 2002
                   9-2
DRAFT-DO NOT QUOTE OR CITE

-------
 1      particles (as indexed by various indicators, e.g., PM25) as likely being important contributors to
 2      the observed human health effects, and (c) now provides additional information on associations
 3      between coarse-fraction (PM10_2 5) particles and adverse health impacts. The overall coherence of
 4      the newer epidemiologic database also is discussed, which strengthens the 1996 PM AQCD
 5      evaluation suggesting a likely causal role of ambient PM in contributing to the reported effects.
 6           The nature of the observed  effects and the biological mechanisms that might underlie such
 7      effects then are discussed. The discussion of potential mechanisms of injury examines ways in
 8      which PM could induce health effects. The increased, but still limited, availability of new
 9      experimental evidence necessary to evaluate or directly substantiate the viability of hypothesized
10      mechanisms is noted. Information concerning possible contributions of particular classes of
11      specific ambient PM constituents also is summarized.
12           The chapter also provides information on the identification of susceptible population
13      groups at special risk for ambient PM effects and factors placing them at increased risk, which
14      need to be considered in generating risk estimates for the possible occurrence of PM-related
15      health events in the United States.
16
17
18      9.2  BACKGROUND
19      9.2.1 Basic Concepts
20           Atmospheric particles originate from a variety of sources and possess a range of
21      morphological, chemical, physical, and thermodynamic properties. Sources include combustion,
22      photochemical oxidation of precursors, and soil dust. Atmospheric particles contain inorganic
23      ions, metallic compounds, elemental carbon, organic compounds, and crustal  compounds.  Some
24      atmospheric particles are hygroscopic and contain particle-bound water. The  organic fraction is
25      especially complex, containing hundreds of organic compounds. Individual particles may be
26      composed by any number of the above and other components.
27
28      9.2.2 Particle Size Distributions
29           As discussed in Chapter 2, the distribution of particles with respect to size is an important
30      physical parameter governing their behavior.  Atmospheric particles vary in density and often are

        April 2002                                 9-3        DRAFT-DO NOT QUOTE OR CITE

-------
 1      not spherical. Therefore, their diameters are often described by an "equivalent" diameter (i.e.,
 2      that of a unit density sphere that would have the same physical behavior).  The aerodynamic
 3      diameter (Da) depends on the density of the particle and is defined as the diameter of a spherical
 4      particle with a density of 1 g/cm3 but with a settling velocity equal to that of the particle in
 5      question. The atmospheric deposition rates of particles, and therefore, their residence times in
 6      the atmosphere, are a strong function of their aerodynamic diameters. The aerodynamic diameter
 7      also influences deposition patterns of particles within the lung. The effects of atmospheric
 8      particles on visibility, radiative balance, and climate, will also be influenced by the size
 9      distribution of the particles. Atmospheric particles cover several orders of magnitude in particle
10      size. Therefore, size distributions often are expressed in terms of the logarithm of the particle
11      diameter on the X-axis and the measured differential concentration on the Y-axis. If the
12      differential concentration is plotted on a linear scale,  the number of particles (per cm3 of air), or
13      the surface area, the volume, or the mass of particles  (per m3 of air) having diameters in the size
14      range from log D to log(D  + AD), will be proportional to the area under that part of the size
15      distribution curve.
16           Averaged atmospheric size distributions are shown in Figure 9-2.  Figure 9-2a shows the
17      number distributions of particles, on a logarithmic scale, as a function of particle diameter for
18      several aerosols.  The particle volume distributions for two of these are shown in Figure 9-2b.
19      These distributions show that most of the particles are quite small, below 0.1 //m; whereas most
20      of the particle volume (and therefore most of the mass) is found in particles larger than 0.1 //m.
21
22      9.2.3 Definitions of Particle Size Fractions
23           Aerosol scientists use four different approaches or conventions in the classification  of
24      particles by size: (1) modes, based on the observed size distributions and formation mechanisms;
25      (2) cut point, usually based on the 50% cut point of the specific sampling device; (3) dosimetry
26      or occupational health sizes, based on the entrance into various compartments of the respiratory
27      system; and (4) legally specified, regulatory sizes for air quality standards.
28
29           Modal.  The modal classification, first proposed by Whitby (1978), is shown in Figure 9-3.
30      In polluted atmospheres, the nuclei mode can be seen clearly in the volume distribution only in
31      traffic or near traffic or other sources of nuclei mode particles. The observed modal structure is
        April 2002                                9-4          DRAFT-DO NOT QUOTE OR CITE

-------
1,000,000 -
X
   10,000 -
        03
       "55
        E
        CD
       b
       _0)
        o
       '•c
        CD
       Q_
                   •••,,-N
              100 -
       1 -
     0.01 -
       .2   0.0001 -
       5
       z
       T3
          0.000001 -
                       \
                       •
-------
         O)

         I
7

6

5

4  -


2

1
                                                                 Mechanically
                                                                  Generated
                DGV = 0.018
                o =1.6
                  1  ' I  ""I
              0.002       0.01

                      Nuclei Mode
                            0.1              1
                           Particle Diameter, Dp(|jm)
                           Accumulation Mode
                             Fine-Mode Particles
       1  I ' '"I     '   r
             10
         Coarse Mode
    Coarse-Mode Particles
100
      Figure 9-3.  Volume size distribution, measured in traffic, showing fine-mode and
                  coarse-mode particles and the nuclei and accumulation modes within the
                  fine-particle mode. DGV (geometric mean diameter by volume, equivalent to
                  volume median diameter) and og (geometric standard deviation) are shown for
                  each mode. Also shown are transformation  and growth mechanisms (e.g.,
                  nucleation, condensation, and coagulation).
      Source: Adapted from Wilson and Suh (1997).
1     0.1 //m. Toxicologists and epidemiologists use the term "ultrafme" and aerosol physicists and
2     material scientists use the term "nanoparticles" to refer to particles in the nuclei-mode size range.
3     Accumulation Mode: That portion of the fine particle mode with diameters above about 0.1 //m.
4          The major processes that influence the formation and growth of particles in the three modes
5     are also shown in Figure 9-3.  New particles may be formed by nucleation from gas phase
6     material.  Particles may grow by condensation as gas phase material condenses on existing
7     particles. Particles also may grow by coagulation as two particles combine to form one.  Gas
      April 2002
                                   9-6
DRAFT-DO NOT QUOTE OR CITE

-------
 1      phase material condenses preferentially on smaller particles, and the rate constant for coagulation
 2      of two particles decreases as the particle size increases. Therefore, nuclei mode particles grow
 3      into the accumulation mode, but accumulation mode particles do not normally grow into the
 4      coarse mode.
 5           Over the years, the terms fine and coarse, as applied to particle sizes, have lost the precise
 6      meaning given in Whitby's (1978) definition.  In any given article, therefore, the meaning of fine
 7      and coarse, unless defined, must be inferred from the author's usage. In particular, PM2 5 and
 8      fine-mode particles are not equivalent. In this document, the term "mode" is used with fine and
 9      coarse when it is desired to specify the distribution of fine-mode particles or coarse-mode
10      particles as shown in Figure 9-3.
11
12      Size-Selective and Occupational Health Size Fractions
13           Size-selective sampling refers to the collection of particles below or within a specified
14      aerodynamic size range, usually defined by the upper 50% cut point size, and has arisen in an
15      effort to measure particle size fractions with some special significance (e.g., health, visibility,
16      source apportionment, etc.). An example of a PM10  and a PM25 size cut are shown in Figure 9-4.
17      The subscripts, 10 and the 2.5, signify the 50% cut size, i.e., the size at which 50% of the
18      particles are collected and 50% of the particles are rejected.  As can be seen, the cut is not
19      perfectly sharp. Some particles larger than the 50%  cut point are collected; neither are all
20      particles smaller than the 50% cut point collected.
21           The occupational health community has defined size fractions for use in the protection of
22      human health. This  convention classifies particles into inhalable, thoracic, and respirable
23      particles according to their upper size cuts (also shown in Figure 9-4).  However, these size
24      fractions may also be characterized in terms of their entrance into various compartments of the
25      respiratory system. Thus, inhalable particles enter the respiratory tract, including the head
26      airways. Thoracic particles travel past the larynx and reach the lung airways and the
27      gas-exchange regions of the lung.  Respirable particles are a subset of thoracic particles that are
28      more likely to reach the gas-exchange region of the lung.
29
30           Regulatory Size Cuts. In 1987, the NAAQS for PM were revised to use  PM10, rather than
31      total suspended particulate matter (TSP), as the indicator for the NAAQS for PM (Federal

        April 2002                                  9-7        DRAFT-DO NOT QUOTE OR CITE

-------
                100
                  0
                                                                     APM10
                                                                     •  IPM
                                                                     •  TPM
                                                                     O  RPM
                                                                     VPM25
                                      4             10     20          50
                                     Aerodynamic Diameter (|jm)
                                 100
      Figure 9-4.  Specified particle penetration (size-cut curves) through an ideal (no-particle-
                  loss) inlet for five different size-selective sampling criteria. Regulatory size
                  cuts are defined in the Code of Federal Regulations; PM25 (2001a), PM10
                  (2001b). PM2 5 is also defined in the Federal Register (1997). Size-cut curves
                  for inhalable particulate matter (IPM), thoracic particulate matter (TPM) and
                  respirable particulate matter (RPM) size cuts are computed from definitions
                  given by American Conference of Governmental and Industrial Hygienists
                  (1994).
1     Register, 1987). The use of PM10 as an indicator is an example of size-selective sampling based

2     on a regulatory size cut (Federal Register, 1987). The selection of PM10 as an indicator was

3     based on health considerations and was intended to focus regulatory concern on those particles

4     small enough to enter the thoracic region of the human respiratory tract. The PM2 5 standard set

5     in 1997 is also an example of size-selective sampling based on a regulatory size cut (Federal

6     Register, 1997). The PM25 standard was based primarily on epidemiological studies using

7     concentrations measured with PM2 5 samplers as an exposure index. However,  the PM2 5 sampler
      April 2002
9-8
DRAFT-DO NOT QUOTE OR CITE

-------
1
2
3
4
5
6
7
was not designed to collect respirable particles. It was designed to collect fine-mode particles.
EPA is currently considering the possibility of a thoracic coarse particle standard with PM10_2 5 as
an indicator. Examples of regulatory size cuts are shown in Figure 9-5. Note also that, in the
range of particle aerodynamic diameter (Da) between 1.0 and 2.5 //m, there is overlap between
fine- and coarse-mode particles.  The degree of overlap depends on prevailing conditions of
humidity and the amount of soil dust in the atmosphere.
    E
    :o
       50  -
               40  -
               30  -
        20  -
        10  -
           0.1
                                                     Coarse-Mode Particles
                        Fine-Mode Particles
                         0.2
 i  • •  i •        i      i   i   ^       r
0.5     1.0     2         5      10      20
   Aerodynamic Particle Diameter (|jm)
                            Total Suspended Particles (TSP)
                                  PM
                                     10
                           PM
                              2.5
                                                       PM
                                                    10.2.5
100
Figure 9-5.  An idealized distribution of ambient particulate matter showing fine-mode
            particles and coarse-mode particles and the fractions collected by size-selective
            samplers. (WRAC is the Wide Range Aerosol Classifier which collects the
            entire coarse mode [Lundgren and Burton, 1995].)
Source: Adapted from Wilson and Suh (1997).
April 2002
                                                9-9
                         DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.3  CHARACTERIZATION OF EMISSION SOURCES
 2           What are the size distribution, chemical composition, and mass-emission rates of
 3     paniculate matter emitted from the collection of primary-particle sources in the United States,
 4      and what are the emissions of reactive gases that lead to secondary particle formation through
 5      atmospheric chemical reactions?
 6
 1           The linkages between airborne PM and its sources are not as well defined as they are for
 8      many other pollutants.  In large part this is because PM is not a well defined chemical entity but
 9      represents a complex mixture of primary and secondary components.  PM is called "primary" if it
10      is in the same chemical form in which it was emitted into the atmosphere. PM is called
11      "secondary" if it is formed by chemical reactions in the atmosphere.  Primary coarse particles are
12      usually formed by mechanical processes, such as the  abrasion of surfaces or by the suspension of
13      soil or biological material.  This includes material emitted in particulate form, such as wind-
14      blown dust, sea salt, road dust, and combustion-generated particles such as fly ash and soot.
15      PM10_25 is mainly primary in origin.  Primary fine particles are emitted from sources either
16      directly as particles or as vapors that rapidly condense to form ultrafme or nuclei-mode particles.
17      Secondary PM is formed by chemical reactions of free, adsorbed, or dissolved gases. Most
18      secondary fine PM is formed from condensable vapors generated by chemical reactions of
19      gas-phase precursors.  Secondary formation processes can result in either the formation of new
20      particles or the addition of condensable vapor to preexisting particles. Most of the sulfate and
21      nitrate and a portion of the organic compounds in atmospheric particles are formed by chemical
22      reactions in the atmosphere. Because precursor gases undergo mixing during transport from their
23      sources, it is difficult to identify individual sources of secondary constituents of PM.
24           Table 9-1 summarizes anthropogenic and natural sources for the major primary and
25      secondary aerosol constituents of fine and coarse particles. Anthropogenic sources can be further
26      divided into stationary and mobile sources.  Stationary sources  include fuel combustion for
27      electrical utilities, residential space heating and industrial processes;  construction and
28      demolition; metals, minerals, and petrochemicals; wood products processing; mills and elevators
29      used  in agriculture; erosion from tilled lands; waste disposal and recycling;  and fugitive dust
30      from paved and unpaved roads. Mobile,  or transportation-related, sources include direct
31      emissions of primary PM and secondary PM precursors from highway and off-highway vehicles
        April 2002                                 9-10        DRAFT-DO NOT QUOTE OR CITE

-------
>
TABLE 9-1. CONSTITUENTS OF ATMOSPHERIC PARTICLES AND THEIR MAJOR SOURCES1
*— Sources
§ Primary (PM <2 . 5 ^m) Primary (PM >2 . 5 ^m) Secondary PM Precursors (PM <2 . 5 /j,m)
Aerosol
species Natural Anthropogenic Natural
SO4= Sea spray Fossil fuel combustion Sea spray
Sulfate
Anthropogenic Natural
— Oxidation of reduced sulfur
gases emitted by the oceans and
wetlands and SO2 and H2S
emitted by volcanism and forest
fires
Anthropogenic
Oxidation of SO2 emitted
from fossil fuel combustion
       NO3-
       Nitrate
       Minerals
                                                                                  Oxidation of NO,, produced by
                                                                                  soils, forest fires, and lighting
                                                   Oxidation of NO,, emitted
                                                   from fossil fuel combustion
                                                   and in motor vehicle
                                                   exhaust
Erosion and     Fugitive dust paved
re-entrainment   and unpaved roads,
               agriculture, and
               forestry
                                    Erosion and re-entrainment
Fugitive dust, paved
and unpaved road
dust, agriculture, and
forestry
^


o
3>
'•Tj
H
6
o
2
0
H
O
c
o
~1
w
o
hrl
7s
H
W
NH4+
Ammonium

Organic
carbon (OC)



Elemental
carbon
(EC)
Metals


Bioaerosols


—

Wild fires




Wild fires


Volcanic
activity

Viruses and
bacteria

'Dash (-) indicates either very




—

Prescribed burning,
wood burning, motor
vehicle exhaust, and
cooking

Motor vehicle exhaust
wood burning, and
cooking
— —

— Tire and asphalt wear
and paved road dust



, — Tire and asphalt wear
and paved road dust

Emissions of NH3 from wild Emissions of NH3 from
animals, and undisturbed soil animal husbandry, sewage,
and fertilized land
Oxidation of hydrocarbons Oxidation of hydrocarbons
emitted by vegetation (terpenes, emitted by motor vehicles,
waxes) and wild fires prescribed burning, and
wood burning

— —


Fossil fuel combustion, Erosion, re-entrainment, — — —
smelting, and brake
wear



minor source or no known


and organic debris

Plant and insect fragments, —
pollen, fungal spores, and
bacterial agglomerates
source of component.











-------
 1      and nonroad sources.  In addition to fossil fuel combustion, biomass in the form of wood is
 2      burned for fuel.  Vegetation is burned to clear new land for agriculture and for building
 3      construction, to dispose of agricultural and domestic waste, to control the growth of animal or
 4      plant pests, and to manage forest resources (prescribed burning).  Also shown are sources for
 5      precursor gases whose oxidation forms secondary particulate matter.
 6           In general, the sources of fine PM are very different from those for coarse PM.  Some of the
 7      mass in the fine size fraction has been formed during combustion from material that volatilized
 8      in combustion chambers and then recondensed before emission into the atmosphere.  By and
 9      large, however, most ambient PM2 5 is secondary, having been formed in the atmosphere from
10      photochemical reactions involving precursor gases. Transport and transformations of precursors
11      can occur over distances of hundreds of kilometers. The coarse PM constituents have shorter
12      lifetimes in the atmosphere, so their effects tend to be more localized.  Only major sources for
13      each  constituent within each broad category shown at the top of Table 9-1 are listed.  Not all
14      sources are equal in magnitude.  Chemical characterizations of primary particulate emissions for
15      a wide variety of natural and anthropogenic sources (as shown in Table 9-1) were given in
16      Chapter 5 of the 1996 PM AQCD.  Summary tables of the composition of source emissions
17      presented in the  1996 PM AQCD and updates to that information are provided in Appendix  3D
18      of Chapter 3 in this document. The profiles of source composition are based largely on results of
19      various studies that collected signatures for use in source apportionment studies.
20           Natural sources of primary PM include windblown dust from undisturbed land,  sea spray,
21      and plant and insect debris. The oxidation of a  fraction of terpenes  emitted by vegetation  and
22      reduced sulfur species from anaerobic environments leads to secondary PM formation.
23      Ammonium (NH4+) ions, which play a major role in regulating the pH of particles, are derived
24      from emissions of ammonia (NH3) gas. Source categories for NH3 have been divided into
25      emissions from undisturbed soils (natural) and emissions that are related to human activities
26      (e.g., fertilized lands,  domestic and farm animal waste).  There is ongoing debate about
27      characterizing emissions from wild fires (i.e., unwanted  fire) as either natural or anthropogenic.
28      Wildfires have been listed in Table 9-1 as natural in origin, but land management practices and
29      other human actions affect the occurrence and scope of wildfires. For example, fire suppression
30      practices allow the buildup of fire fuels and increase the susceptibility of forests to more severe
31      and infrequent fires from whatever cause,  including lightning strikes.  Similarly, prescribed

        April 2002                                9-12       DRAFT-DO NOT QUOTE OR  CITE

-------
 1      burning is listed as anthropogenic, but can viewed as a substitute for wildfires that would
 2      otherwise eventually occur on the same land.
 3           The precursors to secondary PM have natural and anthropogenic sources, just as primary
 4      PM has natural and anthropogenic sources. Whereas the major atmospheric chemical
 5      transformations leading to the formation of particulate nitrate and sulfate have been relatively
 6      well studied, those involving the formation of secondary aerosol organic carbon are still under
 7      active investigation. A large number of organic precursors are involved, many of the kinetic
 8      details still need to be determined, and many of the actual products of the oxidation of
 9      hydrocarbons have yet to be identified.
10           However, over the past decade, a significant amount of research has been carried out to
11      improve the understanding of the atmospheric chemistry of secondary organic PM (SOPM)
12      formation. Although additional sources of SOPM might still be identified, there appears to be a
13      general consensus that biogenic compounds (monoterpenes, sesquiterpenes) and aromatic
14      compounds (toluene, ethylbenzene) are the most significant SOPM precursors. A large number
15      of compounds have been detected in biogenic and aromatic SOPM,  although the chemical
16      composition of these two categories has not been fully established, especially for aromatic
17      SOPM.  Transformations that occur during the aging of particles are still not adequately
18      understood. There are still large gaps in current understanding of a number of key processes
19      relating to the partitioning of semivolatile compounds between the gas phase and ambient
20      particles containing organic compounds, liquid water, inorganic salts, and acids.  In addition,
21      there is a general lack of reliable analytical methods for measuring multifunctional oxygenated
22      compounds in the gas and aerosol phases.
23           Emissions estimates for primary PM2 5 components shown in Table 9-1 are provided in
24      Table 9-2 and emissions of precursors of secondary PM2 5 are shown in Table 9-3. The values
25      shown are annual averages for the entire United States. As can be seen from a comparison of the
26      entries in the two tables, the emissions of precursor gases of secondary PM  are much larger than
27      those for primary PM. It should be noted here that the emissions estimates given above are
28      subject to a considerable degree of uncertainty, which varies from species to species. In addition,
29      there can be a great deal of temporal variability in the emissions.  See NARSTO (2002) for
30      further details regarding the calculation of emissions inventories.
31

        April 2002                                9-13        DRAFT-DO NOT QUOTE OR CITE

-------
           TABLE 9-2.  EMISSIONS OF PRIMARY PM7< BY VARIOUS SOURCES IN 1999
        Source
Emissions
(109 kg/y)     Maj or PM Components
                 Notes
        On-road vehicle        0.21
        exhaust
        Non-road vehicle       0.37
        exhaust
             Organic compounds,
             elemental carbon
             Organic compounds,
             elemental carbon
        Fossil fuel
        combustion
  0.36       Crustal elements, trace
             metals
        Industrial             0.35
        processes

        Biomass burning       1.2
        Waste disposal         0.48

        Fugitive dust          3.3
        Windblown dust        NA1
        Other                0.02

        Total                 6.2
             Metals, crustal material,
             organic compounds

             Organic compounds,
             elemental carbon
             Organic compounds,
             trace metals
             Crustal elements
             Crustal elements
             Organic compounds,
             elemental carbon
Exhaust emissions from diesel (72%) and
gasoline vehicles (28%).

Exhaust emissions from off-road diesel (57%)
and gasoline vehicles (20%); ships and boats
(10%); aircraft (7%); railroads (6%).
Fuel burning in stationary sources such as
power plants (33%); industries (39%);
businesses and institutions (25%); residences
(3%).

Metals processing (29%); mineral products
(27%); chemical mfg. (11%); other industries
(33%).

Managed burning (47%); residential wood
burning (28%); agricultural burning (7%);
wildfires (18%).
Open burning (91%); incineration (9%).


Dust raised by vehicles on paved (19%) and
unpaved roads (40%); construction (15%),
dust from raising crops (24%) and livestock
(2%).

Dust raised by wind on bare land.

Structural fires
        'NA = not available.

        Source: Adapted from U. S. Environmental Protection Agency (2001).
1           Although most emphasis in this section has been placed on sources within the United

2      States, it also should be remembered that sources outside the United States contribute to ambient

3      PM levels that can, at times, exceed the ambient NAAQS. Dense hazes, composed mainly of

4      dust, occur frequently during the summer in southern Florida.  This dust has been emitted in the

5      Sahara Desert and then transported across the Atlantic Ocean.  Large-scale dust storms in the

6      deserts of central Asia recently have been found to contribute to PM levels in the Northwest on

7      an episodic basis. Not  only dust but microbial pathogens and various pollutants are transported
       April 2002
                          9-14
   DRAFT-DO NOT QUOTE OR CITE

-------
         TABLE 9-3. EMISSIONS OF PRECURSORS TO SECONDARY PM2 5 FORMATION
                                     BY VARIOUS SOURCES IN 1999
        Precursor
                 Emissions
                 (109kg/y)
  Secondary PM
    Component
                     Notes
        S02
        NO''2
        Biogenic
        VOCs1
        NH,
                    17
        Anthropogenic      16
        VOCs
                   44
                   45
Sulfate
                   26       Nitrate
                            Various mainly
                            unidentified
                            compounds of 'OC'
Various mainly
unidentified
compounds of 'OC'

Ammonium
Exhaust from on-road (2%) and non-road (5%) engines
and vehicles; fossil fuel combustion by electrical utilities,
industries, other sources (85%); various industrial
processes (7%); and other minor sources (1%).

Exhaust from on-road (34%) and non-road (22%)
engines and vehicles; fossil fuel combustion by electrical
utilities, industries, other sources (39%); lightning (4%);
soils (4%); and other minor sources (5%).

Evaporative and exhaust emissions from on-road (29%)
and non-road (18%) vehicles; evaporation of solvents
and surface coatings (27%); biomass burning (9%);
storage and transport of petroleum and volatile
compounds (7%); chemical and petroleum industrial
processes (5%); other sources i
Approximately 98% emitted by vegetation. Isoprene
(35%); monoterpenes (25%); all other reactive and
non-reactive compounds (40%).

Exhaust from on-road and non-road engines and vehicles
(5%); chemical manufacturing (3%); waste disposal,
recycling, and other minor sources (5%); livestock
(82%); and fertilizer application (18%).
        'Includes estimates of natural sources from Guenther et al. (2000).
        2Emissions expressed in terms of NO2.

        Source:  Adapted from U. S. Environmental Protection Agency (2001).
1

2

3

4

5

6
during these events.  Uncontrolled biomass burning in central America and Mexico may have

contributed to elevated PM levels that exceeded the daily NAAQS level for PM in Texas; and

wildfires throughout the United States, Canada, Mexico, and Central America all contribute to

PM background concentrations in the United States.
       April 2002
                                              9-15
                               DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.4 AMBIENT CONCENTRATIONS
 2          What are the basic characteristics of ambient monitoring data used to draw inferences
 3      about the relations between health outcomes and air pollution?
 4
 5      9.4.1  Measurement of Particulate Matter
 6          It is possible to measure a variety of PM indicators with high precision. However, the
 7      absolute accuracy of a PM monitoring techniques cannot be established because no standard
 8      reference calibration material or procedure has been developed for suspended, atmospheric PM.
 9      Therefore, accuracy is defined as the degree of agreement between a field PM sampler and a
10      collocated PM reference method audit sampler. Intercomparison studies, therefore, are very
11      important for establishing the reliability of PM measurements.
12          One important measurement problem arises from the presence of semivolatile components
13      (i.e., species that exist in the atmosphere in dynamic equilibrium between the condensed phase
14      and gas phase) in atmospheric PM.  Important examples include ammonium nitrate, semivolatile
15      organic compounds, and particle-bound water.  Most filter-weighing techniques for PM,
16      including the U.S. Federal Reference Methods (FRM), require equilibration of collected material
17      at fixed,  near-room temperature (25  °C) and moderate relative humidity (40%) to reduce particle-
18      bound water. This also causes the loss of an unknown, but possibly significant fraction, of
19      ammonium nitrate and semivolatile  organic compounds. Some modest amount of particle-bound
20      water may be present at the 40 % relative humidity at which filter samples are equilibrated.
21      However, to avoid measurement of large amounts of particle-bound water that would be present
22      at higher relative humidities, continuous measurement techniques must reduce particle-bound
23      water in  situ. One technique is to stabilize PM at a specified temperature high enough to remove
24      all, or almost all, particle-bound water. This results in loss of much of the semivolatile PM.
25      Examples include the tapered element oscillating microbalance (TEOM) operated at 50 °C and
26      beta gauge monitors with heated inlets. Another technique is the use of a diffusion denuder to
27      remove water vapor without heating. Examples include the Brigham Young absorptive sampler
28      and Harvard pressure drop monitor.  The three approaches give different mass concentrations,
29      especially in air sheds with high nitrate, wood smoke, or secondary organic aerosols.  Current
30      PM standards are based on health effects studies mainly using filter techniques. However, the

        April 2002                               9-16        DRAFT-DO NOT QUOTE OR CITE

-------
 1      need to provide new real time information to the public and the economic pressure to replace
 2      filter samplers with continuous monitors will require a better understanding of the physics and
 3      chemistry of the semivolatile components of PM and studies of the potential health effects of
 4      these components.
 5
 6      9.4.2  Mass Concentrations
 7          Data for ambient PM2 5 and PM10 concentrations are obtained routinely by networks
 8      operated by various state and local agencies. Data are also collected as part of research efforts by
 9      governmental, academic and industrial groups. Data from state and local agencies are stored in
10      the AIRS (Aerometric Information Retrieval System) data base, maintained by the U.S.
11      Environmental Protection Agency. Concentrations of PM10_2 5 based on FRM PM10 and PM25
12      monitors are estimated by taking the difference between these two measurements. The spatial
13      coverage and frequency of sampling depends on the resources of the agency carrying out the
14      monitoring. Thus, the amount of data collected in a given urban area varies across the United
15      States.
16          The median PM2 5 concentration was 13 //g/m3 in the United States on a county basis, for
17      1999 and 2000.  The corresponding median PM10_25 concentration was about 10 //g/m3 for the
18      same period. However, there was a good deal of variability in the annual means in different
19      environments in the United States. The mean PM2 5 concentration was below 7 //g/m3 in 5% and
20      below 18 //g/m3 in 95% of counties that met minimum AIRS data completeness criteria for
21      calculation of an annual mean concentration (at least 11 days data for each calendar quarter).
22      The mean PM10_2 5 concentration was below 4 //g/m3 in 5% and below 21 //g/m3 in 95% of
23      counties meeting the criteria given above. Mean PM2 5 and PM10_2 5 concentrations reported by
24      the IMPROVE network were considerably lower than the lowest 5th percentile values reported by
25      state and local agencies.
26          An adequate characterization of the PM concentrations found in urban areas cannot be
27      obtained by considering only annual average concentrations for the whole urban area.  There can
28      be considerable spatial and temporal variability in the concentration fields.  Typically, annual
29      mean concentrations are within 5 //g/m3 of each other in urban areas (MSAs). The spread in
30      values can be much greater if CMSAs are considered.  Even within some MSAs, concentrations
31      measured at separate sites on individual days can differ by over 100 //g/m3.
        April 2002                                9-17        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Pairs of sites within MSAs are correlated with each other to varying degrees, depending on
 2      the urban area. There are some very general regional patterns evident in the data base in which
 3      sites tend to be more highly correlated with each other in the eastern United States and less well
 4      correlated with each other in the western United States.  Figure 9-6 shows an example for
 5      Philadelphia, PA.  The exceptions are frequent enough to prevent extrapolation from one city to
 6      another without first examining the data. Although sites may be highly correlated with each
 7      other within an MSA, this does not mean that the concentration fields are uniform, as illustrated
 8      by Figure 9-7 for three urban areas.  Concentrations for the three  site pairs chosen are all well
 9      correlated with each other (r>0.9), but the concentrations display different degrees of uniformity.
10
11      9.4.3 Physical  and Chemical Properties of Ambient  PM
12           Physical and chemical properties of fine-mode and coarse-mode particles that are produced
13      by sources listed in Table 9-1 are summarized in Table 9-4. It can readily be seen that fine- and
14      coarse-mode particles show striking differences in the nature of their sources, their composition,
15      and hence, their chemical properties, and in their removal processes. Differences in sources and
16      removal processes for fine- and coarse-mode particles account for many differences in their
17      behavior in the atmosphere.  The much shorter atmospheric lifetimes of coarse particles
18      compared to fine particles implies that fine particles can travel much further in the atmosphere
19      than coarse particles.  The more sporadic nature of the sources of coarse particles, in addition,
20      implies that coarse PM should be more highly spatially variable than fine PM. Elemental
21      compositions, including trace elements by X-ray flourescence analysis, for PM25 and PM10_2 5 in
22      two cities with different fine/coarse relationships are given in Table 9-5. The major chemical
23      components of PM25 from several sites in the eastern, interior, and western parts of the United
24      States are shown in Figure 9-8.
25
26
27
        April 2002                                 9-18        DRAFT-DO NOT QUOTE OR CITE

-------
                                       Phildelphia, PA MSA
                                                                  AIRS Site ID
                                                       Site A
                                                       SiteB
                                                       SiteC
                                                       SiteD
                                                       SiteE
                                    34-007-0003
                                    34-007-1007
                                    42-045-0002
                                    42-101-0004
                                    42-101-0136
                    b.
                               100 km
                              3400700031  3400710071  4204500021  4210100041  4210101361
                         Mean
                         Obs
                          SD
                           50-
                           40-
14.941
 108
8.524
15.427
 103
8.749
15.992
 112
8.265
14.823
284
8.537
14.718
 278
8.295
                                                                1234
                      c.  Site    A
          B
                  D
A
B
C
D
E
0.964 0.868
1 (3.3,0.082) (6.3,0.155)
95 98
0.849
1 (6.9,0.158)
94
1


0.88
(4.8, 0.129)
81
0.894
(3.7, 0.135)
77
0.868
(5.0,0.149)
85
1

0.868
(5.4, 0.147)
80
0.857
(6.4, 0.148)
79
0.818
(6.6, 0.154)
83
0.918
(4.9,0.13)
246
1
Figure 9-6.  Philadelphia, PA-NJ MSA.  (a) Locations of sampling sites by AIRS ID#;
             (b) Quarterly distribution of 24-h average PM2 5 concentrations; (c) Intersite
             correlation statistics, for each data pair, the correlation coefficient, (P90,
             coefficient of divergence) and number of measurements are given.
April 2002
             9-19
                 DRAFT-DO NOT QUOTE OR CITE

-------
                                 Columbia SC1999 & 2000

1
CD
O
O
0
"o
L_
CD
|
z



CO
8
CD
ecu mi
0
"5
L_
CD
.Q
E
z
120 -
100 -

80 -

60 -

40 -
20 -
Q

45-079-0007 vs 45-079-001
r = 0.97
COD = 0.06
Psf> = 2.7figlm>





h

(V nf N^ CK' 9} ^ ^ *> N *> I/
N?  N$> *$> <$>
Chicago IL2000
30 -,
25 -
20 -
15 -

10 -


5 -
n -.

17-031-2001 vs 17-031-4201
r = 0.94
COD = 0.1 4



•
•
1 .
II. II -
                                     Detroit Ml 2000
8 12-
c
£ 10-
3 R
o 8 -
0
•5 6-
te 4 -
1 2 -
z n







|























26-099-0009 vs 26-163-0033
r = 0.93
COD = 0.22








1
1
P90= 12.7/(g/m3
_
1
1 1
nil. ,i . i
                              Concentration Difference (//g/m3)
Figure 9-7.  Occurrence of differences between pairs of sites in three MSAs. The absolute
            differences in daily average PM2 5 concentrations between sites are shown on
            the x-axis and the number of occurrences on the y-axis. The MSA, years of
            observations, AIRS site I.D. numbers for the site pairs, Pearson correlation
            coefficients (r), coefficients of divergence (COD), and 90th percentile (P90)
            difference in concentration between concurrent measurements are also shown.

Source:  Pinto et at. (2002).
April 2002
9-20
DRAFT-DO NOT QUOTE OR CITE

-------
                 TABLE 9-4.  COMPARISON OF AMBIENT PARTICLES,
       FINE MODE (Nuclei Mode Plus Accumulation Mode) AND COARSE MODE
                                     Fine
                                                                 Coarse
                      Nuclei
                              Accumulation
Formation
Processes:

Formed by:
Composition:
 Solubility:
Atmospheric
half-life:

Removal
Processes:
Travel
distance:
            Combustion, high-temperature
         processes, and atmospheric reactions
Nucleation
Condensation
Coagulation
Sulfates
Elemental Carbon
Metal compounds
Organic compounds
with very low
saturation vapor
pressure at ambient
temperature
Probably less soluble
than accumulation
mode

Minutes to hours
Grows into
accumulation mode

-------
  TABLE 9-5. CONCENTRATIONS OF PM2 5, PM10 2 5 AND SELECTED ELEMENTS
                     IN THE PM?, AND PM,n,, SIZE RANGE
                                           10-2.5
Phoenix, AZ (n = 164)

Species
Mass
Al
Si
P
S
Cl
K
Ca
Ti
V
Cr
Mn
Fe
Co
Ni
Cu
Zn
As
Se
Br
Pb
Concentration
PM25
11,200
125
330
11
487
19
110
129
11
0.7
0.6
5.7
177
-0.4
0.6
5.2
17
1.9
0.4
3.8
6.6
(ng/m3)
PM10.2.5
27,600
1879
535
37
131
208
561
1,407
130
2.0
2.6
29
1,211
1.2
1.8
10.3
25
0.6
-0.02
0.8
4.6
Philadelphia, PA (n =

Species
Mass
Al
Si
P
S
Cl
K
Ca
Ti
V
Cr
Mn
Fe
Co
Ni
Cu
Zn
As
Se
Br
Pb
Concentration
PM25
29,800
109
191
15
3,190
23
68
63
8.7
9.7
1.4
3.2
134
0.8
8.5
7.7
56
0.4
1.3
14
28
20)
(ng/m3)
PM10.2.5
8,400
325
933
28
38
47
100
421
30
3.2
1.0
6.3
352
-0.2
2.0
14
52
0
-0.1
3.0
13
 Source: Zweidinger et al. (1998); Pinto et al. (1995).
April 2002
9-22
DRAFT-DO NOT QUOTE OR CITE

-------
       Figure 9-8.  Major chemical components of PM2 5 as determined in the pilot study for
                   EPA's national speciation network.
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
9.5  AIR QUALITY MODEL DEVELOPMENT AND TESTING
     What are the linkages between emissions sources and the biologically important
components of paniculate matter?

     Atmospheric models that address this question fit into two general categories. Either they
are process oriented and attempt to predict variables of interest based on the solution of equations
describing basic physical and chemical processes or they are statistically oriented and rely on the
statistical analysis of atmospheric data to infer information about the nature and relative
importance of different sources.  Although there are many sub-categories within each of these
two broad categories, the two main types of models that are under active development and
application are chemistry-transport models (CTMs) and receptor models.
       April 2002
                                        9-23
DRAFT-DO NOT QUOTE OR CITE

-------
 1
 2
 3
 4
 5
 9
10
11
12
13
14
15
     The main components of a CTM are summarized in Figure 9-9. Models such as the
CMAQ (Community Model for Air Quality) system and MAQSIP (Multiscale Air Quality
Simulation Platform) incorporate the processes shown in Figure 9-9 as numerical algorithms to
predict time dependent concentration fields of a wide variety of gaseous and paniculate phase
pollutants. Also shown in Figure 9-9 is the meteorological model used to provide the inputs for
calculating the transport of species in the CTM. The meteorological models such as the MM5
model, which supples these inputs to the CTMs mentioned above, also provide daily weather
forecasts.  The domains of these models extend typically over several thousand kilometers by
several thousand kilometers. Because these models are computationally intensive, it is often
impractical to run them over larger domains without sacrificing some features. For these
reasons, both the meteorological model and the CTM must have boundary conditions that allow
the effects of processes occurring outside the model domain to be felt. The entire system of
meteorological model emissions processors and output processors constituents the framework of
EPA'sModels-3.
                                        Meteorological
                                            Model
                                                        Emissions
                                                          Model
                                                        Anthropogenic
                                                        (point, area sources)
                                                            &
                                                    V  Biogenic Emissions
                                            Chemistry Transport Model
                                                 Visualization of Output
                                                   Process Analyses
       Figure 9-9.  Main components of a comprehensive atmospheric chemistry modeling
                   system, such as Models 3.
       April 2002
                                         9-24
DRAFT-DO NOT QUOTE OR CITE

-------
 1           The performance of models such as these must be evaluated by comparison with field data
 2      as part of a cycle of model improvements and subsequent evaluations. Discrepancies between
 3      model predictions and observations can be used to point out gaps in current understanding of
 4      atmospheric chemistry. Very often, however, the algorithms in the model are 'tuned' to improve
 5      agreement between the model predictions and a particular set of observations. Model evaluation
 6      does not merely involve a straightforward comparison between model predictions and the
 7      concentration field of the pollutant of interest. Even this task is not straightforward in the case of
 8      PM, because PM is composed of a number of different substances with different chemical and
 9      physical properties. A comparison of model predicted PM25 mass with measured PM25 mass
10      may not be very meaningful because there can be compensating errors in the model calculations
11      and there are significant artifacts affecting the collection and retention of a number of PM
12      components such as semi-volatile organic compounds and ammonium nitrate. Because of the
13      number and complexity of the parameterizations used in CTMs, there may be compensating
14      errors and tests of these parameterizations must be made for individual physical and chemical
15      processes.
16           Another issue relates to the averaging time that is used  for both the observations and the
17      model outputs. Model predictions can be made with time steps shorter than an hour, however, as
18      noted in Chapters 2 and 3, there is a considerable degree of uncertainty associated with individual
19      hourly observations by continuous monitors. Emissions inventories, as shown in Tables 9-2 and
20      9-3, represent annual averages;  and it is impractical, except in a few cases, to increase that
21      resolution down to even a few days. At least for modeling ozone, it has  been found that
22      agreement between model and observations is improved if seasonal averages, rather than
23      episodic averages are considered.
24           Models such as the CTMs discussed above have been under development for a number of
25      years. Discussions of these models have not been included in the earlier chapters because these
26      models are not yet being used to provide information about human exposures that could be
27      incorporated into this document. CTMs are being used to develop emissions control strategies
28      and to aid in implementation of existing air quality standards. The reader is referred to NARSTO
29      documents (NARSTO, 2002) for further details.
30           There are two main approaches to receptor modeling.  Receptor models such as the
31      chemical mass balance (CMB) model relate source category contributions to ambient

        April 2002                               9-25        DRAFT-DO NOT QUOTE OR CITE

-------
 1      concentrations based on analyses of the composition of ambient PM and source emissions
 2      samples. This technique has been developed for apportioning source categories of primary PM
 3      and was not formulated to include the processes of secondary PM formation. In the second
 4      approach, various forms of factor analysis are used.  They rely on the analysis of time series of
 5      compositional data from ambient samples to derive both the composition of sources and the
 6      source contributions. Standard approaches such as factor analysis or Principal Component
 7      Analysis (PCA) can apportion only the variance and not the mass in an aerosol composition data
 8      set.  Positive matrix factorization (PMF) is a recently developed multivariate technique that
 9      overcomes many of the limitations of standard techniques, such as principal components analysis
10      (PCA), by allowing for the treatment of missing data and data near or below detection limits.
11      This is accomplished by weighting elements inversely according to their uncertainties.  Standard
12      methods such as PCA weight elements equally regardless of their uncertainty. Solutions also are
13      constrained to yield nonnegative factors.  Both the CMB and the PMF approaches find a solution
14      based on least squares fitting and minimize an object function. Both methods provide error
15      estimates for the solutions based on estimates of the errors in the input parameters. It should be
16      remembered that the error estimates often contain subjective judgments.  For a complete
17      apportionment of mass, all of the major sources  affecting a monitoring site must be sampled for
18      analysis by CMB, whereas there is no such restriction in the use of PMF.
19           Among other approaches, the UNMIX model takes a geometric approach that exploits the
20      covariance of the ambient data to determine the number of sources, the composition and
21      contributions of the sources, and the uncertainties (Henry, 1997). A simple example may help
22      illustrate the approach taken by UNMIX. For example, in a two-element scatter plot of ambient
23      Al and Si, a straight line and a high correlation for Al versus Si can indicate a single source for
24      both species (soil), while the slope of the line  gives information on the composition of the soil
25      source. In the same data set, iron may not plot on a straight line against Si, indicating other
26      sources of Fe in addition to soil. More importantly, the Fe-Si scatter plot may reveal a lower
27      edge. The points defining this edge represent ambient samples collected on days when the only
28      significant source of Fe was soil.  Success of the UNMIX model hinges on the ability to find
29      these "edges" in the ambient data from which the number of sources and the source compositions
30      are extracted. UNMIX uses principal component analysis to find edges in m-dimensional space,
31      where m is the number of ambient species. UNMIX does not make explicit use of errors or

        April 2002                                9-26        DRAFT-DO NOT QUOTE OR CITE

-------
 1     uncertainties in the ambient concentrations, unlike the methods outlined above. This is not to
 2     imply that the UNMIX approach regards data uncertainty as unimportant, but rather that the
 3     UNMIX model results implicitly incorporate error in the ambient data. The underlying
 4     philosophy here is that the uncertainties are often unquantifiable, and hence it is best to make no
 5     a priori assumptions about what they are.
 6           For most practical purposes, the relative contributions of sources to ambient PM  samples
 7     are determined by receptor models.  Receptor models have  most successfully been applied to the
 8     determination of sources of primary PM.  The process based models are more flexible  and could
 9     be used for determining sources of secondary PM. However, they are computationally much
10     more intensive, and they rely on a large number of inputs, with varying degrees of uncertainty.
11     Arguably, emissions inventories represent the major source of uncertainty in the application of
12     CTMs (see e.g., Calvert et al., 1993).  However, significant uncertainty also exists in
13     photochemical transformations, in part because of a lack of data for many key reactions. Further
14     uncertainty is added in the methods that are used to reduce the literally thousands of reactions
15     involving hundreds of species occurring in the atmosphere to more tractable numbers through the
16     use of idealized chemical mechanisms. Issues concerning the gas phase mechanisms are relevant
17     because the free radicals that are involved in the formation  of photochemical oxidants are also
18     involved in the formation of secondary PM.
19
20
21     9.6  EXPOSURE TO PARTICULATE MATTER AND COPOLLUTANTS
22           What are the quantitative relationships between concentrations of particulate matter and
23     gaseous copollutants measured at stationary outdoor air-monitoring sites and the contributions
24     of these concentrations to actual personal exposures, especially for subpopulations and
25     individuals?
26
27           It will be useful to separate these relationships into two components:  (a) central site to
28     outdoor concentrations; and (b) outdoor concentrations to personal exposures.
29
30
31
       April 2002                                9-27        DRAFT-DO NOT QUOTE OR CITE

-------
 l      9.6.1  Central Site to Outdoor
 2           The first component to be examined is the relationship between ambient PM concentrations
 3      measured by a central monitor, located at a  site presumably representative of the community (or
 4      the average of several such sites), and the ambient PM concentration just outside an indoor
 5      microenvironment such as a home.
 6
 7      9.6.1.1 Exposure for Acute Epidemiology
 8           In acute time-series studies, daily deaths (or other health effects) are regressed against the
 9      daily ambient PM concentrations as measured at a single site (or the average of several sites) in a
10      city.  Spatial variations in daily exposure can lead to errors in the estimated relative risk.  Under
11      the assumption of a linear relationship between exposure and effect, analysis of exposure error
12      suggests that a key indicator of the effect on epidemiologic results of spatial variations in
13      exposure will be the strength of the daily site-to-site correlations of ambient PM concentrations.
14      Chapter 3 presents a substantial body of new monitoring data from AIRS. A range of
15      correlations of PM25 concentrations were found between monitoring sites in the cities chosen for
16      analysis. PM10 and TSP sites were frequently chosen to monitor specific local  point or area
17      sources. However, PM25 sites are chosen primarily to be representative of community
18      exposures.  Still it would be wise to check the representativeness of a site before choosing a site
19      or group of sites to provide a representative community concentration for exposure or
20      epidemiologic studies. As shown in Figure 9-10, site-to-site correlations tend to be higher for a
21      site pair where both of which are dominated by regional PM than for a site pair where one of
22      which is more strongly influenced by local sources.
23
24      9.6.1.2 Exposure for Chronic Epidemiology
25           In chronic studies, total or annual deaths in large cohorts in different cities are regressed
26      against long-term or annual average concentrations in the different cities. Few analyses of
27      exposure error have been performed for this case. However, the key consideration for chronic
28      studies might be differences in the annual (or seasonal) averages in different parts of a city.
29      Prior to NRC-1, there was little information on the variations of long term PM concentration
30      averages across cities.  Some information on the spatial variations in long-term (seasonal)
31      averages are reported in Chapter 3 of this document, based on data from AIRS.
        April 2002                                 9-28        DRAFT-DO NOT QUOTE OR CITE

-------
CD
'O
it
0
O
O
c
o
•j=  0.80 •
1C
•0

b
o
                   Chicago, IL
               X   XX   x
              XXXv X Y   X
              X **  / #
                   X
40      60

Distance (km)
                                      Washington, DC-MD-VA-WV
1.00-



^,
•g 0 90 -
CD
0
Sfc
CD
0
O

o
13
CD
O
O
0.70-


Xv v K
X «vX$X X v
\ xxxX
x Xf X X X
x * x x xx

% x x
X X
x x x xx

v X
-X
X
X
x x
X
X
X
                                                    0     20     40      60
                                                                                   100     120
                                                                   Distance (km)
                 Pittsburgh, PA
                                     Los Angeles - Long Beach, CA
1.0-
0.9-

.^

CD °°~
'o
£
0
0
O 0.7-
c
.0
15
CD 06-
O
O
05-


0.4-

« J x.x xy xXj( x „ ^x
X * y Y y X
A A "AW
xx xx x
X xxXX X
X A
A

" X
x x
X

X j,
x x x
X
X
x x



1 1 1 1
1.0-
0.9-

• — '

'§ 0.7-
0
o
O 0.6-
c
,0
IS 0.5-
0
b
0 Q.4-

0.3-

0.2-


X y X
XXX v
X~ v
" X
XX
x
x x x
x x

x X x x x x
x x x
X
X


X

I I 1 I
                    Distance (km)
                                                                   Distance (km)
Figure 9-10.  Correlograms showing the variation in site-to-site correlation coefficient for
              PM2 5 as a function of distance between sites for several cities.


Source: Fitz-Simons et al. (2000).
April 2002
                        9-29
DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.6.2 Home Outdoor Concentrations Versus Ambient Concentrations
 2            Indoors and the Ambient Contribution to Total Personal Exposure
 3           What is the relationship between the concentration of ambient PM outside a home and the
 4      concentration of ambient PM that has infiltrated into the home?
 5
 6      9.6.2.1  Mass Balance Model
 7           It will be useful to  review some concepts derived from the equilibrium mass balance
 8      model, discussed in detail in Chapter 5.  The ratio of the ambient PM concentration outdoors, C,
 9      to the concentration of ambient PM that has infiltrated indoors, C(AI), is given by the infiltration
10      factor where P is the particle penetration efficiency, a is the air exchange rate, and k is the
11      deposition rate.
12
13                          C(AI)/C = Pa/(a+k)  = Fmp (the infiltration factor)                (9-1)
14
15      As will be discussed later, P and k are functions of the particle size, so FINP will also depend on
16      particle size.  The mass balance equation may be modified to include particle removal by air
17      handling systems and to  account for nonequilibrium behavior.
18           While indoors, a person will be exposed to a concentration of ambient pollution given by
19      C • FINP. However, while outdoors a person will be exposed to the full ambient concentration.
20      The infiltration factor and the fraction of time outdoors may be used with the ambient
21      concentration to  estimate the ratio of the ambient PM exposure (while indoors and outdoors) to
22      the ambient PM concentration, where y = the fraction of time spent outdoors,
23
24                  A/C =y + (l-y)FINP =y + (l-y)Pa/(a+k) = a (the attenuation factor).        (9-2)
25
26      Since y and a may vary from day to day and person to person and P and k will vary with particle
27      size, a will also be a variable.
28           It is necessary to understand the infiltration factor, used to estimate the concentration of
29      ambient PM concentration indoors [C(AI) = C • FINP\, and the attenuation factor, used to estimate
30      the ambient exposure, i.e.,  personal exposure to  particles  of ambient original, [A = C •  a],


        April 2002                               9-30        DRAFT-DO NOT QUOTE OR CITE

-------
 1     because they may be estimated from exposure measurements and used to estimate A, the ambient
 2     component of total personal exposure.
 3
 4     9.6.2.2  Separation of Total Personal Exposure into its Ambient and Nonambient
 5             Components
 6           A person's total exposure to PM or other pollutants includes a nonambient component,
 7     usually divided into a component due to indoor-generated pollutants that are evenly distributed
 8     through out the house and a component, sometimes called the personal cloud, due to activities of
 9     the person that generate pollutants which influence that person more than other persons in the
10     same house. Thus, total personal exposure, T, equals the sum of ambient exposure, A, and
11     nonambient exposure, N:
12
13                                            T = A+N                                 (9-3)
14
15     As NRC Topic  1 makes clear, a key variable of interest is A, the ambient exposure, i.e., the
16     contributions of particulate matter and gaseous copollutants measured at stationary outdoor
17     air-monitoring sites to actual personal exposures, not  T, the total personal exposures due to
18     ambient and indoor-generated pollutants. However, it is not possible to measure A or N directly.
19     Only T and C can be measured directly.  It is necessary to understand the infiltration factor, used
20     to estimate the concentration of ambient PM concentration indoors, \C(AI) = C • FINF], and the
21     attenuation factor, used to estimate the ambient exposure, [A =  C • a], because these factors may
22     be estimated from exposure measurements and used to estimate A, the ambient component of
23     total personal exposure.
24           In recent years, the need to separate personal exposure into ambient and nonambient
25     components has been recognized (Wilson and Suh, 1997), techniques for separating total
26     personal exposure into its ambient and nonambient components have been recommended
27     (Wilson et al., 2000), several papers have reported average values of a and N, and one paper has
28     reported individual values of A.
29
30     Average Values
31           As shown in Figure 9-11, regression of individual measurements of personal exposure on
32     the corresponding measurements of ambient concentrations yields two components of total
       April 2002                                9-31         DRAFT-DO NOT QUOTE OR CITE

-------
        o
        Q.
        LJJ
        "cc
        c
        o
            250
        ,^  200-
        •»  150
        cu
        *,  100-
        0)
        Q_

        1
              50-
               0
r= 0.373
R2= 0.142
T, = 40.5+0.711C,
#= 147
                                               •   T= 90 + 01 C or T = N + aC = N + A
                  0
50            100           150
   Ambient PM Concentration, Ct
                                                     200
                            250
        Figure 9-11.  Regression analysis of daytime total personal exposures to PM10 versus
                     ambient PM10 concentrations using data from the PTEAM study. The
                     slope of the regression line is interpreted by exposure analysts as the
                     average a, where aC = A.
        Source: Wilson et al. (2000)
1     exposure, one dependent on concentration, one not (T= 60 + 6jC; Zeger et al., 2000). Exposure
2     analysts associate the component independent of concentration, 60, with cohort average
3     nonambient exposure and the component dependent on concentration, 6l3 with alpha, a, the ratio
4     of ambient exposure to ambient concentration (T = N + aC = N + A; Dockery and Spengler,
5     1981; Ott et al., 2000; Wilson et al., 2000).  Most exposure studies report the correlation between
6     ambient concentrations and personal exposure, and many of these also report the slope of the
7     relationship. Since the slope may be interpreted as the average alpha there are a number of
8     studies from which estimates of the average alpha  may be estimated. However, the slope may
9     not accurately reflect the average alpha unless the data has been examined for outliers.  Several
      April 2002
                            9-32
DRAFT-DO NOT QUOTE OR CITE

-------
 1
 2
 3
 4
 5
 9
10
11
12
13
14
15
studies have interpreted the slope and reported the average FINP or a for cohorts (Ott et al., 2000;
Wilson et al., 2000; Patterson and Eatough, 2000; Landis et al., 2001).

Individual Values
     The high correlations found between ambient sulfate and personal sulfate (which has few
indoor sources) suggest that a better relationship may be found between ambient concentrations
and ambient exposures than between ambient concentrations and total personal exposures to PM
(Figure 9-12). The PTEAM study provided sufficient information to permit estimation of
individual values of ambient PM10 exposure, A.  These individual values of A were found to be
highly correlated with the corresponding ambient PM10 concentration, C (Figure 9-13) (Wilson
et al., 2000).  It is also important to determine whether or not the nonambient exposure, N, is a
function of C, since if TV is not correlated with C, N cannot be a confounder in a regression of
health effects on ambient concentration (Figure 9-14) (Zeger et al., 2000).

1.00 -

0.75 -
£ 0.50 -
0)
0 0.25 -
1
o
0 0.00 -
0
•5; -0.25 -
O
f i
U -0.50 -
-0.75 -
_-i nn -
Ebeltetal.,2000
Pearson's "r"
pt
PM26 y
1












Sulfate













Percentile

90th Percentile
75th Percentile
Median

25th Percentile

10th Percentile








Sarnatetal., 2000
Spearman's "r"
PM25

r
i





T


—



1







gl

T
Sulfate











                                  PM2 5  Sulfate
                                                       PM2 5   Sulfate
       Figure 9-12.  Comparison of correlation coefficients for longitudinal analyses of personal
                     exposure for individual subjects versus ambient concentrations of PM25 and
                     sulfate.
       Source: Ebelt et al. (2000), Sarnat et al. (2000).
       April 2002
                                          9-33
DRAFT-DO NOT QUOTE OR CITE

-------
              100
             E 75-
          §.<
          uj 5
50-
               25-
                            50         100        150        200
                             Ambient PM Concentration, Ct (|jg/m3)
                                                     250
Figure 9-13. Regression analysis of daytime exposures to the ambient component of
            personal exposure to PM10 (ambient exposure) versus ambient PM10
            concentrations.

Source: Wilson et al. (2000).
CO
-t-» ^
CD 5
s_ _
0 -100-
UJ i
1 « 50
CD <:
°" § o-
^n
c
r = 0.051 *
R!= 0.0026 *
N = 39.S+0.086C •
# = 147
. . * . *
* * *
4 4 * +* *
4 74^f/ tr **: *
^vAxi^V *::^^ * 4 *
* * *
i i i
) 50 100 150 200 25
                        Ambient PM Concentration, C, (|jg/m3)

Figure 9-14. Regression analysis of daytime exposures to the nonambient component of
            personal exposure to PM10 (nonambient exposure) versus ambient PM10
            concentrations. The two variables are unrelated.

Source: Mage et al. (1999)
April 2002
                       9-34
DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.6.3 Variability in the Relationship Between Concentrations and
 2            Personal Exposures
 3           The values of the infiltration factor and alpha may vary from person-to-person as shown by
 4      the distribution of the infiltration factor and alpha in the PTEAM study (Figure 9-15) (Wilson
 5      et al., 2000). The average value of alpha may vary from season-to-season and from city-to-city.
 6      The variation in average alpha across cities, as estimated by city-to-city air-conditioning use, can
 7      explain some of the variation in the quantitative effects of particles on health across cities
 8      (Figure 9-16) (Janssen et al., 2002). For a given PM component, the air exchange rate, a, is a
 9      major factor in determining the relationship between outdoor and personal exposure. This has
10      been shown in a study in which personal exposure data were classified into three groups based on
11      home ventilation status. High values of alpha and high correlations were found for the well-
12      ventilated homes, lower values for moderately well-ventilated homes, and much lower values for
13      poorly ventilated homes.
14
15      9.6.4 Exposure Relations for Co-Pollutants
16           The key issue is whether the gaseous co-pollutants (CO, NO2, SO2, and O3) contribute to
17      the health effects attributed to PM or whether they merely serve as surrogates for PM. If the
18      gaseous co-pollutants were responsible for some or all of the health effects attributed to PM in a
19      single pollutant, community time-series epidemiologic analysis, they would be contributors, and
20      the health effects due to PM would be overestimated. However, if the gaseous co-pollutants
21      were surrogates for PM, i.e., significantly correlated with PM but not contributing to the health
22      effects attributed to PM in the analysis, in a multiple regression, the surrogate would share some
23      of the health effect with the causal agent, especially if the surrogate were measured more
24      accurately than the causal agent. Thus, use of a surrogate in a multiple regression would result in
25      an underestimation of the health effects due to PM.
26           In community, time-series epidemiology,  in which daily, community-average health effects
27      are regressed against daily ambient concentrations, there are several requirements that must be
28      met in order for a gaseous co-pollutant to be a contributor to the health effects attributed to PM.
29      (1) The gaseous co-pollutant must be capable of causing the effect at the level of the community
30      exposure, (2) the daily ambient concentrations of the gaseous co-pollutant must be related to (i.e.,
31      correlated with) the daily ambient concentrations of the PM indicator, and (3) the daily ambient
        April 2002                                9-3 5       DRAFT-DO NOT QUOTE OR CITE

-------
                        Fraction of Ambient PM10 Found Indoors (F,NF = C(AI)/C)
                     0.20    0.30   0.40   0.50    0.60    0.70    0.80   0.90
                     Fraction of Ambient PM10 Found in Personal Exposure (a =A/C)
                     0.20    0.30    0.40   0.50   0.60   0.70    0.80    0.90
Figure 9-15.   Distribution of individual, daily values of the infiltration factor, F^ =
              C(AI)/C and the attenuation factor, a = A/C, estimated using data from the
              PTEAM study. The distribution of the attenuation factor is shifted to higher
              values compared to the infiltration factor because people are exposed to the
              full ambient concentration when outdoors.

Source: Wilson et al. (2000).
April 2002
9-36
DRAFT-DO NOT QUOTE OR CITE

-------
                    0.0025
                    0.0020 -
                •    0.0015 -
                o
                o
                Q   0.0010 -
                O
                    0.0005 -
                    0.0000
                                                                Winter peaking cities
                                                                Non-winter peaking cities
                                             O
                                 10      20      30      40      50     60
                                           Central Air Conditioning (%)
                                                                             70      80
       Figure 9-16.  Percentage of homes with air conditioning versus the regression coefficient
                     for the relationship of cardiovascular-related hospital emissions to ambient
                     PM10 concentrations. The higher the percent air conditioning, the lower the
                     amount of personal exposure to ambient PM per unit of ambient PM
                     concentration, i.e., lower a, and there a lower regression coefficient (increase
                     in risk per until PM10 exposure).
       Source: Janssen et al. (2002).
 1     concentrations of the gaseous co-pollutant must be related to (correlated with) the personal
 2     exposures to that gaseous co-pollutant.  Requirements 1 and 2 are also requirements for being a
 3     "confounder" in epidemiologic and biostatistics terminology.  Whether or not requirement 3 is
 4     also a requirement for confounding will depend on the exact definition used for confounding.
 5     A fourth requirement, that may apply to confounding, is that the gaseous co-pollutant not be in
 6     the formation pathway of the PM. Since SO2 and NO2 are in the formation pathway for the
 7     sulfate and nitrate components of PM and O3 is a key chemical reactant in the formation of the
 8     sulfate, nitrate, and organic components of PM, this fourth requirement has implications for
 9     possible confounding of PM by gaseous co-pollutants that have not yet been adequately analyzed.
10           The exposure analyst is  concerned with requirements 2 and 3. How well are the daily
11     ambient concentrations of the gaseous co-pollutants correlated with the daily ambient
       April 2002
9-37
DRAFT-DO NOT QUOTE OR CITE

-------
 1      concentrations of PM (or specific PM components or indicators) and are the daily ambient
 2      concentrations of the gaseous co-pollutants correlated with the daily personal exposures to the
 3      ambient? In order to answer these questions quantitatively, information would be needed on the
 4      spatial variability of PM indicators and the gaseous co-pollutants and on the variability of the
 5      factors which control the infiltration factors (penetration factor and deposition or removal rates).
 6           Exposure relationships for gaseous co-pollutants were not reviewed in the exposure chapter
 7      (Chapter 5) of this document. Although there have been many exposure studies of the gaseous
 8      co-pollutants, there has been little analysis of the experimental data in terms relevant to
 9      epidemiology. Exposure studies for CO, NO2, and O3 have been reviewed in the respective Air
10      Quality Criteria Documents (U.S. Environmental Protection Agency, 1993, 1996b, 2000a) and
11      exposure studies of the gaseous co-pollutants and PM components have been reviewed by Monn
12      (2001).  Qualitative information on exposure relationships which may be inferred from the
13      studies reviewed in these publications are given in Table 9-6.
14
15
               TABLE 9-6.  QUALITATIVE ESTIMATES OF EXPOSURE VARIABLES

Highest
High
Medium
Low
Lowest
Spatial Homogeneity1
scv
PM25
NO2, O3, PM10.2 5, SO2
CO, EC4
UF5
Infiltration Factor2
CO
PM2 5, SO4=, EC4
NO2
PM10.2.5
UF5, 03, S02
Stability of the
Infiltration Factor3
CO
PM2 5, SO4=, EC4
N02, PM10.2 5, UF5
03, S02

         1. As indicated by the inverse size of the site-to-site correlation coefficient.
         2. As indicated by the value of the infiltration factor, inferred in the case of gaseous co-pollutants from
           indoor/outdoor ratios for homes without known indoor sources.
         3. As indicated by the inverse sensitivity of the deposition or removal rate to the surface to volume ratio and
           the chemical composition of the surface.
         4. Elemental carbon.
         5. Ultrafine particles.
        April 2002                                 9-3 8        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Based on the estimates in Table 9-6, it might be expected that the correlation between daily
 2      ambient concentrations of PM25 and sulfate and personal exposure to PM25 and sulfate would be
 3      high and statistically significant but that this relationship would not be as significant for the
 4      gaseous co-pollutants.  Two recent studies (Sarnat et al., 2000, 2001) provide new information
 5      relevant to the possible contribution of gaseous co-pollutants to the health effects attributed to
 6      PM.  Personal exposure measurements were made of NO2, O3, and sulfate (winter and summer)
 7      and of SO2 and EC (winter only).  Ambient measurements were made of these species (same
 8      seasons) and of CO (both seasons). Personal exposures to ambient PM2 5 were estimated by
 9      using the daily, individual ratios of personal exposure to sulfate to ambient concentrations of
10      sulfate as an estimate of the attenuation factor for PM25.  Correlations among ambient
11      concentrations, among personal exposures, and between ambient concentrations and personal
12      exposures were examined.
13           Daily personal exposures to NO2 and O3 were not significantly correlated with daily
14      ambient concentrations of those gaseous co-pollutants in either summer or winter. This suggests
15      that NO2 and O3 cannot be contributors to the health effects attributed to PM in an epidemiologic
16      analysis using daily ambient concentrations.  In the winter, daily personal exposures to SO2 were
17      negatively correlated with daily ambient concentrations of SO2.  Personal exposures to CO were
18      not reported. During summer, O3 and NO2 were positively and significantly associated with
19      PM25; the association with CO was positive but not significant.  During winter, CO and NO2
20      were positively and significantly associated with PM2 5 while O3 was negatively and significantly
21      associated with PM25; the association with SO2 was negative but not significant.  Similar
22      association of gaseous  co-pollutants were found with personal exposure to PM2 5 except that the
23      winter association with SO2 became significant. Also, the significant associations were more
24      significant with personal exposure to ambient PM2 5. This indicates that  daily ambient
25      concentrations of CO, NO2, O3 and SO2 can be surrogates for daily ambient concentrations of
26      PM2 5 but that exposure and epidemiologic analyses including O3 and SO2 need to examine
27      relationships on a seasonal basis.  These studies also indicate that daily ambient concentrations of
28      PM2 5, CO, NO2, O3 and SO2 serve as surrogates for daily personal exposures to PM2 5 and are
29      even better surrogates for daily personal exposures to ambient PM2 5.  Thus, in a multiple
30      regression using PM and a gaseous copollutant, both variables would be  surrogates  for personal
31      exposure to ambient PM.

        April 2002                                 9-39        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Sarnat et al. (2001) point out that "it is inappropriate to treat one variable as a confounder
 2      of another when both variables are actually surrogates of the same thing."  While the exposure
 3      results from these studies are based on a small number of non-randomly chosen subjects and
 4      therefore cannot be extrapolated with assurance to other situations, they do indicate the value of
 5      exposure analysis in identifying which of several collinear variables could possibly be causal.
 6      The work also suggests that neither NO2, O3, nor SO2 are likely to be the causal factor in the
 7      reported associations of ambient PM with health effects. No information was found on the
 8      correlation of ambient CO with personal exposure to CO in homes with no indoor CO sources.
 9      However, the low spatial homogeneity of ambient CO concentrations suggests that the
10      relationship would be weak.  Therefore, it seems likely, but not certain, that exposure
11      relationships would also indicate that CO is unlikely to be a contributor to the health effects
12      attributed to PM. It is important to understand that this does not indicate that these ambient
13      pollutants do not cause health effects of the type associated with PM in epidemiologic analyses.
14      It only indicates that community, time-series epidemiology using ambient concentrations cannot
15      provide information on the possible health effects of pollutants whose ambient concentrations are
16      not significantly correlated with personal exposure to that ambient pollutant.
17           Sarnat et al. (2001) also suggest that some of the gaseous co-pollutants may be acting as
18      surrogates for specific PM25 source categories or components.  "For subjects with COPD,
19      ambient CO and NO2 were not significantly associated with total personal PM2 5, but were
20      significantly associated with personal exposure to PM2 5 of ambient origin and also to personal
21      elemental carbon (EC). These significant associations may be due to the fact that motor vehicles
22      are a major source of CO, NO2, EC, and, to a lesser degree, to PM2 5 of ambient origin.
23      Conversely, ambient CO and NO2 were not significantly associated with personal sulfate, a
24      pollutant not associated with  motor vehicle emissions. O3, in contrast, was predominantly
25      associated with personal sulfate (positively in summer and negatively in winter) . . ." Thus, CO,
26      NO2, EC, and PM2 5 may be surrogates for personal exposure to pollutants from motor vehicles
27      and O3 may be a surrogate for regional sulfate. It should be noted that since PM2 5, CO, NO2, EC,
28      and PM  associated with motor vehicles are  all significantly correlated with each other, a
29      community, time-series epidemiologic analysis, in one community for one time period, cannot
30      tell whether a variable is actually responsible for relationship between concentration and health
31      effects observed in the analysis, or whether the variable is a surrogate for the causal variable.

        April 2002                                9-40        DRAFT-DO NOT QUOTE OR CITE

-------
 1      In order to more clearly differentiate between contributor and surrogate, it will be necessary to
 2      integrate information from toxicology and exposure analysis, as well as from epidemiologic
 3      studies in different time periods and different communities.
 4
 5      9.6.5 Summary
 6           For most cities, site-to-site correlations are high for PM2 5. However, the spatial
 7      distribution of PM should be investigated before beginning long term monitoring for exposure or
 8      epidemiologic studies.  The relationship between the concentrations of an ambient pollutant
 9      outdoors and the contribution of that ambient pollutant to personal exposure is given by the mass
10      balance model (Equation 9-2) and depends on the outdoor concentration, the time spent outdoors
11      and indoors, the air exchange rate, and penetration factor, and the indoor deposition or removal
12      rate.  For a given PM component, the major cause of variability in the relationship is the air
13      exchange rate. For gaseous co-pollutants, if the correlation between the ambient concentrations
14      and the personal exposures to the ambient concentrations are not statistically significant, that
15      gaseous co-pollutant cannot contribute to the health effect attributed to PM in a community,
16      time-series epidemiologic analysis.  However, if ambient concentration of the gaseous
17      co-pollutant is significantly correlated with the ambient concentration of PM, it may be as good
18      or better indicator of personal exposure to the lexicologically active component of PM as the
19      ambient PM concentration; and, thus, it may be a surrogate, i.e., it will falsely show an effect due
20      to its correlation with the personal exposure to the active component and, in a multiple
21      regression, it will appear to reduce the effect associated with PM.  Therefore, correlations among
22      ambient concentrations and ambient concentration-personal exposure relationships for PM and
23      co-pollutants are useful in interpreting the results of epidemiologic studies.
24
25
26      9.7  EXPOSURE TO BIOLOGICALLY IMPORTANT
27           CHARACTERISTICS OF PARTICULATE MATTER
28           What are the exposures to biologically important constituents and specific characteristics
29      of paniculate matter that cause responses in potentially susceptible subpopulations and the
30      general population?
31
        April 2002                                9-41        DRAFT-DO NOT QUOTE OR CITE

-------
 1           In their discussion of Topic 2, the NRC notes that in order to make such investigations
 2      practicable, it will be necessary to characterize susceptible subpopulations more fully, identify
 3      lexicologically important chemical constituents or particle-size fractions, develop and field-test
 4      exposure-measurement techniques for relevant properties of PM, and design comprehensive
 5      studies to determine population exposures.
 6
 7      9.7.1 Exposure Relationships for Susceptible Subpopulations
 8           Children, the elderly, and people  with pre-existing diseases such as diabetes, respiratory
 9      disease, and cardiovascular disease appear to constitute susceptible subpopulations.  A number of
10      studies of small cohorts drawn from these and other subpopulations have been conducted
11      recently by EPA and other organizations. Correlations between ambient concentrations and total
12      personal exposure have been presented for a few of these. However, most of the studies have not
13      yet been published, most of the studies have not reported the ambient exposure, and the studies
14      have not been analyzed to determine if there are indeed exposure differences between susceptible
15      groups and the general population.
16           An analysis of cohort exposure studies available in 1998 (Wallace, 2000) concluded that
17      the personal cloud component of nonambient exposure was less for subjects with  COPD than for
18      the general population, healthy elderly  subjects or children, presumably because of the higher
19      activity level of  younger or healthier subjects.  However, the relationship between ambient
20      concentrations and personal exposure for COPD patients was not better than that for other
21      cohorts.  Wallace (2000) noted that the desirable correlation is that "between personal exposure
22      to particles originating outdoors and outdoor concentrations." However, at that time there was
23      no information on the ambient component of personal exposure.  Unfortunately, there is still no
24      published information that would suggest differences in exposure relationships for healthy versus
25      susceptible populations.
26
27      9.7.2 Toxicologically Important Components of PM
28           Inherent in the NRC research agenda (NRC, 1998) was the consideration that one, or
29      perhaps a few, characteristics of PM would be associated with toxicity, and exposure monitoring
30      could concentrate on these components. However, it has not yet been possible to  identify any

        April 2002                                9-42        DRAFT-DO NOT QUOTE OR CITE

-------
1     PM characteristic as not being of toxicologic importance. Table 9-7 lists characteristics of PM
2     that have been found to be associated with toxicity either through epidemiologic or toxicologic
3     studies.
4
5
           TABLE 9-7. PARTICULATE MATTER CHARACTERISTICS POTENTIALLY
      	RELEVANT TO HEALTH	
         Particle number
         Particle surface area
         Mass-ultrafme PM [PM0 J
         Mass-fine PM [PM2 5 or PMl 0]
         Mass-thoracic coarse PM  [PM10_2 5 or PM1(M]
         Sulfate
         Strong acidity (H+)
         Nitrate
         Elemental carbon
         Organic carbon (many different compounds)
         Transition metals
         Specific toxic metals
         Bioaerosols
1     9.7.3  Exposure-Measurement Techniques
2          Measurement techniques, suitable for stationary monitors with 24-hour collection periods,
3     exist for the characteristics of PM listed in Table 9-7. For many of these measurements,
4     continuous or 1-hour-average stationary monitors also exist or are in development. However,
5     personal monitoring is usually limited to either PM2 5 or PM10. A few studies have included
6     passive monitors for NO2, O3, SO2, and CO. A roll-around monitor, which can be rolled around
7     to follow a person and thus simulate a personal exposure measurement with a more complete

      April 2002                               9-43       DRAFT-DO NOT QUOTE OR CITE

-------
 1      suite of measurements, has been used in recent studies. However, with the exception of personal
 2      light-scattering monitors, which do not have well-established relationships with PM mass, there
 3      are still no adequate personal monitors for continuous measurement of mass of other PM
 4      characteristics.
 5
 6      9.7.4  Comprehensive Studies to Determine Population Exposure
 7           Chapter 5 reports only  four exposure studies that have even attempted to provide
 8      statistically representative studies of population exposure of the general population or susceptible
 9      subgroups. However, only in the case of the PTEAM study has the exposure data been used to
10      estimate ambient and nonambient exposure separately (for PM10).  Even though statistically
11      representative studies are limited, available data from small cohorts allow some inferences
12      regarding differences in concentration-exposure relationships among different characteristics of
13      PM.  PM may be classified by particle size, by chemical composition, or by sources.
14      Concentration - exposure relationships may be different for different classes of particles.
15
16      Central Site to Outdoors
17           The 1996 PM AQCD reported information from a few cities (mostly eastern US) that
18      suggested that site-to-site correlation coefficients, r, were high  for sulfate and PM25 in some
19      cities; were lower but still relatively high for PM10 and TSP;  but were low  for PM10_2 5 (Figure
20      9-17). However, there was little information on site-to-site correlations of chemical components
21      of PM (except sulfate and strong acidity) or of orthogonal source-category factors.  New site-to
22      site correlation studies, using PM data from the AIRS data base, are presented in Chapter 3.
23      Some examples of the differences in site-to-site correlations  for PM2 5 and PM10_25, derived from
24      the data in Chapter 3, are shown in Figure 9-18.  It should be noted that the PM2 5 data is from
25      1999 and 2000 and satisfies  certain criteria for number of days  of data  per season. The PM10_2 5
26      data is from 2000 only and is less complete. In addition, some information on the site-to-site
27      correlations of PM25 components and source contributions are  now available (Figures 9-19 and
28      9-20). In order to reduce spatial variability, some cohort studies have used the concentration at
29      the nearest monitoring site or the distance to major traffic sources  for exposure information.
30
31
        April 2002                                9-44        DRAFT-DO NOT QUOTE OR CITE

-------
                1.0


                0.9


                0.8


                0.7

              o 0.6
              £
              
-------
         -0.2
                                         Site Pairs
                                                                                - Mass j
                                                                                - Sulfate |
                                                                                -OC  |
                                                                                -EC  j
                                                                                -Crustal |
                                                                                - Lead I
Figure 9-19.  Site-to-site correlation coefficients for PM25 mass and some chemical
              components of PM25 in 1994 in Philadelphia, PA.

Source: Pinto et al. (1995).
        o     1
                                                                   10     11
                                       Site Pair
                                                                              - Mass      |
                                                                              - Secondary   I
                                                                              - Motor Vehicles j
                                                                              - Crustal     |
                                                                              - Residual Oil  I
Figure 9-20.  Site-to-site correlation coefficients for PM2 5 mass and several source category
              factors in 1986 in the South Coast Basin (Los Angeles area).
Source: Wongphatarakul et al. (1998)

April 2002
9-46
DRAFT-DO NOT QUOTE OR CITE

-------
 1
 2
 3
 4
 5
 9
10
11
12
13
14
15
Outdoors to Indoors
     Information on the infiltration rate, FINP , as a function of particle size may be obtained as
follows.  Indoor and outdoor measurements of PM concentrations as a function of particle size
are made during the night when it is assumed that there are no indoor activities occurring that
might generate indoor PM.  Under this assumption the indoor concentration measurement is
C(AI) and C(AI)/C = FINP (Long et al., 2000). As can be seen in Figure 9-21, FINP is low for
ultrafme and coarse particles but high for accumulation mode particles. FINP also depends on the
air exchange rate, a,  FINP increases when a increases. The variation of P and k as a function of
particle size can also be determined by this technique (Figure 9-22) (Long et al., 2000).  There is
little information on ambient concentration - exposure relationships for specific chemical
components, except sulfate, or for specific source categories, other than what would be inferred
from the size distributions.  Infiltration ratios are low for components like strong acidity (FT) that
are neutralized by indoor-generated ammonia or like ammonium nitrate (NH4NO3) that evaporate
indoors.
                        1.1
                    o
                    ro
                    u.
                    c
                    _o
                    '•C
                    2
                    +-
                    t
                1.0 -
                0.9 -
                0.8 -
                0.7 -
                0.6 -
                0.5 -
                0.4 -
                0.3 -
                0.2 -
                0.1 -
                0.0
                                       0.1
                                          Summer  Fall
                                           LO  CNI  CO
                              (Dp  T—
                              0 O  o
                                                    O
                                                    co
                                                    CD
                                                    I--.
                                                    CD
                                                  Particle Diameter (|jm)
                      Source: Long, Sun and Koutrakis (2000)
        Figure 9-21.   Values of geometric mean infiltration factor, FINF = A/C, as a function of
                      particle diameter for hourly nighttime data (assuming no indoor sources) for
                      summer and fall seasons. Distribution of air exchange rates, a, for each
                      season are shown in the insert.
        Source:  Long et al. (2000).
        April 2002
                                           9-47
DRAFT-DO NOT QUOTE OR CITE

-------
5-
CD
'o
£
                  ts
                  C
                  d>
                  Q.
                      1.2
               1.0
               0.9
               0.8
               0.7
               0.6
               0.5
               0.4
               0.3
               0.2
               0.1
               0.0
                             S  §
                                   q CD  ^
                                     q  ,1  ^
                                           LO  CN CO
                                                             •si-  LO CD
                                                             m  -t ui
                  1.2
                  1.1
                  1.0
                  0.9
                  0.8
                  0.7
                  0.6
                  0.5
                  0.4
                  0.3
                  0.2
                  0.1
                  0.0
                                                                                   i
                                                                                   O
                                                                 O
                                                                 Q
                                              Size Interval (pm)
       Figure 9-22.  Values of penetration efficiency and deposition rate as a function of particle
                     diameter estimated from model of average nighttime indoor-outdoor
                     concentration data.
       Source: Long et al. (2000).
 1
 2
 3
 4
 5
 9
10
11
12
13
14
9.7.5  Air Pollutants Generated Indoors
     The NRC discussion of Research Topic 2 is clear that the primary purpose of the
investigations recommended should be "to examine the outdoor contributions to measurements
of total personal exposure."  However, they also recommended determining the exposure to "air
pollutants generated indoors." Total personal exposure includes both ambient and nonambient
sources.  Important sources of indoor PM are smoking, cooking, and cleaning. Because of the
variation of Finfwith particle size, ambient-infiltrated PM tends to be primarily in the
accumulation mode. As shown in Table 9-8, however, indoor PM is generated primarily in  the
ultrafme mode (smoking, other combustion sources,  most cooking) or the coarse mode  (cleaning,
sauteing). Another, possibly important indoor source, is the reaction of ambient-infiltrated ozone
with indoor emissions of terpenes from air fresheners or cleaning agents, e.g., cleaning  with Pine
Sol.  These particles are also generated largely in the ultrafme mode.  Ambient and indoor
generated PM also differ somewhat in their chemical composition as shown in Table 9-9.
       April 2002
                                          9-48
DRAFT-DO NOT QUOTE OR CITE

-------
             TABLE 9-8.  VOLUME MEAN DIAMETER (VMD) OF INDOOR
                                  PARTICLE SOURCESab
Particle Source
Cooking
Baking (Electric)
Baking (Gas)
Toasting
Broiling
Stir-Frying
Frying
Barbecuing
Sauteing, fine
Sauteing, coarse
Cleaning
Dusting
Vacuuming
Cleaning with Pine Sol
General Activities
Walking Vigorously (w/Carpet)
Sampling w/Carpet
Sampling w/o Carpet
Burning Candles
N

8
24
23
4
3
20
2
13
13

11
10
5

15
52
26
7
Indoor Activity - Mean VMD
(//m)

0.189
0.107
0.138
0.114
0.135
0.173
0.159
0.184
3.48

5.38
3.86
0.097

3.96
4.25
4.28
0.311
 Notes:
 Includes only individual particle events that were unique for a given time period and could be detected above
  background particle levels.
 Fine particle sizes calculated for PV0 02.0 5 using SMPS data; coarse particle sizes calculated for PV0 7_10 using
  APS data.

 Source:  Long et al. (2000).
April 2002                                   9-49         DRAFT-DO NOT QUOTE OR CITE

-------
          TABLE 9-9. CONCENTRATION DIFFERENCES BETWEEN CONSTITUENTS OF
         	NONAMBIENT (INDOOR-GENERATED) AND AMBIENT PM	
        Higher Concentration in Nonambient PM
   Higher Concentration in Ambient PM
        Mold Spores
        Endotoxin
        Animal Dander
        Biological Fragments
        (from insects, etc)
        Environmental Tobacco Smoke
        Resuspended Soil and House Dust
        Ultrafine Particles and Coarse-Mode Particles
   Pollen
   Transition Metals (non-soil Fe, Mn)
   Other Metals (Se, As, Ni, Cu)
   Oxygenated and Nitrated
   Polyaromatic Compounds
   Other Oxygenated Organic Compounds
   Sulfates and Nitrates
   Accumulation-Mode Particles
 1     9.8 DOSIMETRY:  DEPOSITION AND FATE OF PARTICLES IN THE
 2          RESPIRATORY TRACT
 3          What are the deposition patterns and fate of particles in the respiratory tract of individuals
 4     belonging to presumed susceptible subpopulations?
 5          Knowledge of the dose of particles delivered to a target site or sites in the respiratory tract
 6     is important for understanding possible health effects associated with human exposure to ambient
 7     PM and for extrapolating and interpreting data obtained from studies of laboratory animals. The
 8     dosimetry of particles of different sizes are subject to large differences in regional respiratory
 9     tract deposition, translocation, and clearance mechanisms and pathways and, consequently,
10     retention times. The  following sections summarize the current understanding of the physical
11     characteristics of particles and the biological determinants that affect particle dosimetry
12     mechanisms and pathways, as discussed in Chapter 6.
13
14     9.8.1  Particle Deposition in the Respiratory Tract
15          For dosimetry purposes, the respiratory tract can be divided into three regions:
16     (1) extrathoracic (ET), (2) tracheobronchial (TB), and (3) alveolar (A).  The ET region consists
17     of head airways (i.e.,  nasal and oral passages) through the larynx and represents the areas through
18     which inhaled air first passes. In humans, inhalation can occur through the nose or mouth (or
       April 2002
9-50
DRAFT-DO NOT QUOTE OR CITE

-------
 1      both, known as oronasal breathing). However, most laboratory animals commonly used in
 2      respiratory toxicological studies are obligate nose breathers.
 3           From the ET region, inspired air enters the TB region at the trachea. From the level of the
 4      trachea, the conducting airways then undergo branching for a number of generations. The
 5      terminal bronchiole is the most peripheral of the distal conducting airways and these lead,
 6      in humans, to the respiratory bronchioles, alveolar ducts, alveolar sacs, and alveoli (all of which
 7      comprise the A region). All of the conducting airways, except the trachea and portions of the
 8      mainstem bronchi, are surrounded by parenchymal tissue.  This is composed primarily of the
 9      alveolated structures of the A region and associated blood and lymphatic vessels. It should be
10      noted that the respiratory tract regions are comprised of various cell types and that there are
11      distinct differences in the cells of airway surfaces in the ET, TB, and A regions.
12           Particles deposit in the respiratory tract by five mechanisms: (1) inertial impaction,
13      (2) sedimentation, (3) diffusion, (4) electrostatic precipitation, and (5) interception.  Sudden
14      changes in airstream direction and velocity cause inhaled particles to impact onto airway
15      surfaces.  The ET and upper TB airways are dominant sites of inertial impaction, a key
16      mechanism for particles with aerodynamic diameter (Da) >1 //m.  Particles with Da > 0.5 //m
17      mostly are affected by sedimentation out of the airstream.  Both sedimentation and inertial
18      impaction influence deposition of particles in the same size range and occur in the ET and TB
19      regions, with inertial impaction dominating in the upper airways and gravitational settling
20      (sedimentation) increasingly more dominant in lower conducting airways. Particles with actual
21      physical diameters <  1 //m are increasingly subjected to diffusive deposition due to random
22      bombardment by air molecules, resulting in contact with airway surfaces. Particles between
23      0.3 and 0.5 //m in size are small enough to be little influenced by impaction or sedimentation and
24      large enough to be minimally influenced by diffusion, and  so, they undergo the least respiratory
25      tract deposition.  The interception potential of any particle  depends on its physical size; fibers are
26      of chief concern for interception, their aerodynamic size being determined mainly by their
27      diameter. Electrostatic precipitation is deposition related to particle charge; effects of charge on
28      deposition are inversely proportional to particle size and airflow rate.  This type of deposition is
29      likely small compared to effects of other deposition mechanisms and is generally a minor
30      contributor to overall particle deposition, but one recent study found it to be a significant TB
31      region deposition mechanism for ultrafine, and some fine, particles.

        April 2002                                 9-51        DRAFT-DO NOT QUOTE OR CITE

-------
 1           The ET region acts as an efficient filter that reduces penetration of inhaled particles to the
 2      TB and A regions of the lower respiratory tract. Total respiratory tract deposition increases with
 3      particle size for particles >1.0 //m Da, is at a minimum for particles 0.3 to 0.5 //m, and increases
 4      as particle size decreases below that range.  The ET deposition is higher with nose breathing than
 5      for mouth breathing, with increased ventilation rates associated with increasing levels of physical
 6      activity or exercise leading to more oronasal breathing and increased delivery of inhaled particles
 7      to TB and A regions in the lung.
 8           Hygroscopicity, the propensity of a material for taking up and retaining moisture, is a
 9      property of some ambient particle species and affects respiratory tract deposition.  Such particles
10      can increase in size in humid air in the respiratory tract and, when inhaled, deposit according to
11      their hydrated size rather than their initial size. Compared to nonhygroscopic particles  of the
12      same initial  size, deposition of hygroscopic aerosols in different regions varies, depending on
13      initial size:  hygroscopicity generally increases total deposition for particles with initial sizes
14      larger than «0.5 //m, but decreases deposition for particles between «0.01 and 0.5 and again
15      increases deposition for particles <0.01 //m.
16           Enhanced particle retention occurs on carinal ridges in the trachea and throughout the
17      segmental bronchi; and deposition "hot spots" occur at airway bifurcations or branching points.
18      Peak deposition sites shift from distal to proximal sites as a function of particle size, with greater
19      surface dose in conducting airways than in the A region for all particle sizes.  Whereas  both fine
20      (<2.5 //m) and thoracic coarse (2.5 to 10 //m) particles deposit to about the same extent on a
21      percent particle mass basis in the trachea and upper bronchi, a distinctly higher percent of fine
22      particles deposit in the A region. However, surface number dose (particles/cm2/day) is much
23      higher for fine than for coarse particles, indicating much higher numbers of fine particles
24      depositing, with the fine fraction contributing upwards of 10,000 times greater particle  number
25      per alveolar macrophage.
26           Ventilation rate, gender, age, and respiratory disease status are all factors that affect total
27      and regional respiratory tract particle deposition. In general, because of somewhat faster
28      breathing rates and likely smaller airway size, women have somewhat greater deposition of
29      inhaled particles than men in upper TB airways, but somewhat lower A region deposition than
30      for men.  Children appear to show four effects:  (1) greater total  respiratory tract deposition than
31      adults (possibly as much as 50% greater for those <14 years old  than for adults >14 years),

        April 2002                                 9-52         DRAFT-DO NOT QUOTE OR CITE

-------
 1      (2) distinctly enhanced ET region deposition (decreasing with age from 1 year), (3) enhanced TB
 2      deposition for particles < 5 //m, and (4) enhanced A region deposition (also decreasing with age).
 3      Overall, given that children have smaller lungs and higher minute volumes relative to lung size,
 4      they likely receive greater doses of particles per lung surface area than adults for comparable
 5      ambient PM exposures.  This and the propensity for young children to generally exhibit higher
 6      activity levels and associated higher breathing rates than adults likely contribute to enhanced
 7      susceptibility to ambient particle effects resulting from particle dosimetry factors.  In contrast,
 8      limited available data on respiratory tract deposition across adult age groups (18 to 80 years) with
 9      normal lung function do not indicate age-dependent effects (e.g., enhanced deposition in healthy
10      elderly adults).  Altered PM deposition patterns due to respiratory disease status may put certain
11      groups of adults (including some elderly) and children at greater risk for PM effects.
12           Both information noted in the 1996 PM AQCD and newly published findings discussed in
13      this document indicate that respiratory disease status is an especially important determinant of
14      respiratory tract particle deposition. Importantly, the pathophysiologic characteristics of chronic
15      obstructive pulmonary disease (COPD) contribute to more heterogenous deposition patterns and
16      differences in regional deposition. One study indicates that people with COPD tend to breath
17      faster and deeper than those with normal lungs (i.e., about 50% higher resting ventilation) and
18      had about 50% greater deposition than age-matched healthy adults under typical breathing
19      conditions,  with average deposition rates 2.5 times higher under elevated ventilation rates.
20      Enhanced deposition appears to be associated more with the chronic bronchitic than the
21      emphysematous component of COPD. In this and other new studies, fine-particle deposition
22      increased markedly with increased degree of airway obstruction (ranging up to 100% greater with
23      severe COPD).  With increasing airway obstruction and uneven airflow because of irregular
24      obstruction patterns, particles tend to penetrate more into remaining better ventilated lung areas,
25      leading to enhanced focal deposition at airway bifurcations and alveoli in those A region areas.
26      In contrast, TB deposition increases with increasingly more severe bronchoconstrictive states, as
27      occur with asthmatic conditions.
28           Differences  between species in particle deposition patterns were summarized in the 1996
29      PM AQCD and more recently by Schlesinger et al. (1997), as discussed in Chapter 6 of this
30      document.  These differences should be considered when relating biological responses obtained
31      in laboratory animal studies to effects in humans. Various species used in inhalation toxicology

        April  2002                                  9-53       DRAFT-DO NOT QUOTE OR CITE

-------
 1      studies serving as the basis for dose-response assessment may not receive identical doses in a
 2      comparable respiratory tract region (i.e., ET, TB, A) when exposed to the same aerosol at the
 3      same inhaled concentration. This is illustrated by mathematical modeling studies that evaluate
 4      interspecies differences in respiratory tract deposition. For example, Hofmann et al. (1996)
 5      found total deposition efficiencies for all particles (0.01, 1, and 10 //m) at upper and lower
 6      airway bifurcations to be comparable for rats and humans, but when higher penetration
 7      probabilities from preceding airways in the human lung were considered, bronchial deposition
 8      fractions were mostly higher for humans.  For all particle sizes, deposition at rat bronchial
 9      bifurcations was less enhanced on the carinas than in human airways. Numerical simulations of
10      three-dimensional particle deposition patterns within selected (species-specific) bronchial
11      bifurcations indicated that interspecies  differences in morphologic asymmetry is a major
12      determinant of local deposition patterns.  The dependence of deposition on particle size is similar
13      in rats and humans, with deposition minima in the 0.1- to l-//m size range for both total
14      deposition and deposition in the TB and A regions, but total respiratory tract and TB deposition
15      was consistently higher in the  human lung. Alveolar regional deposition in humans was lower
16      than in rat for 0.001- to 10-//m particles (deposition of such particles being highest in the upper
17      bronchial airways), whereas it was higher for 0.1- and l-//m particles in more peripheral airways
18      (i.e., bronchiolar airways in rat, respiratory bronchioles in humans). In a histology study, Nikula
19      et al. (2000) examined particle retention in rats (exposed to diesel soot) and humans (exposed to
20      coal dust). In both, the volume density of deposition increased with increasing dose.  In rats,
21      diesel exhaust particles were found mainly in lumens of the alveolar duct and alveoli, whereas in
22      humans, retained dust was mainly in interstitial tissue. Thus, in the two species, different lung
23      cells appear to contact retained particles and may result in different biological responses with
24      chronic exposure.
25           The probability of any biological  effect of PM in humans or animals depends on particle
26      dosimetry, and subsequent particle retention, as well as underlying dose-response relationships.
27      Interspecies dosimetric extrapolation must, therefore, consider differences in deposition,
28      clearance, translocation, and dose-response. Even similar deposition patterns may not result in
29      similar effects in different species, because dose also is affected by clearance mechanisms and
30      species sensitivity. Total number of particles deposited in the lung may not be the most relevant
31      dose metric by which to compare species; rather, the number of deposited particles per unit

        April 2002                                9-54        DRAFT-DO NOT QUOTE OR CITE

-------
 1      surface area may determine response.  Even if deposition is similar in rats and humans, there
 2      would be a higher deposition density in the rat because of the smaller surface area of the rat lung.
 3      Thus, species-specific differences in deposition density are important when attempting to
 4      extrapolate health effects observed in laboratory animals to humans.
 5
 6      9.8.2 Particle Clearance and Translocation
 7           Particles depositing on airway surfaces may be cleared from the respiratory tract completely
 8      or translocated to other sites within this system by regionally specific clearance mechanisms, as
 9      follow: ETregion—mucocialiary transport, sneezing, nose wiping and blowing, and dissolution
10      and absorption into blood; TB region—mucociliary transport, endocytosis by macrophages and
11      epithelial cells, coughing, and dissolution and absorption into blood and lymph; A region—
12      macrophages, epithelial cells, interstitial, and dissolution and absorption into blood and lymph.
13           Regionally specific clearance defense mechanisms operate to clear deposited particles of
14      varying particle  characteristics (size, solubility, etc.) from the ET, TB, and A regions and are
15      variously affected by different disease states. For example, particles are cleared from the ET
16      region by mucociliary transport to the nasopharynx  area, dissolution and absorption into the
17      blood, or sneezing, wiping or blowing of the nose; but such clearance is slowed by chronic
18      sinusitis, bronchiectasis, rhinitis, and cystic fibrosis. Also, in the TB region, poorly soluble
19      particles are cleared mainly by upward mucociliary  transport or by phagocytosis by airway
20      macrophages that move upward on the mucociliary  blanket,  followed by swallowing.  Soluble
21      particles in the TB region are absorbed mostly into the blood and some by mucociliary transport.
22      Although TB clearance is generally fast and much material is cleared in <24 h, the slow
23      component of TB clearance (likely associated with bronchioles 24 h and clearance
25      half-times of about  50 days.  Bronchial mucous transport is slowed by bronchial carcinoma,
26      chronic bronchitis, asthma, and various acute respiratory infections; these are disease conditions
27      that logically would be expected to increase retention of deposited particle material and, thereby,
28      increase the probability  of toxic effects from inhaled ambient PM components reaching the TB
29      region. Also, spontaneous coughing, an important TB region clearance mechanism, does not
30      appear to fully compensate for impaired mucociliary clearance in small airways and may become
31      depressed with worsening airway disease, as seen in COPD.
        April 2002                               9-55        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Clearance of particles from the A region by alveolar macrophages and their mucociliary
 2      transport is usually rapid (<24 h). However, penetration of uningested particles into the
 3      interstitium increases with increasing particle load and results in increased translocation to lymph
 4      nodes.  Soluble particles not absorbed quickly into the blood stream and translocated to
 5      extrapulmonary organs (e.g., the heart) within minutes may also enter the lymphatic system, with
 6      lymphatic translocation probably being increased as other clearance mechanisms (e.g., removal
 7      by macrophages) are taxed or overwhelmed under "particle overload" conditions.  Insoluble
 8      particles <2 //m clear to the lymphatic system at a rate independent of size; particles of this size,
 9      more so than those >5.0 //m, are deposited significantly in the A region. Translocation into the
10      lymphatic system is quite slow, and elimination from lymph nodes even slower (half-times
11      estimated in decades). Focal accumulations  of reservoirs of potentially toxic materials and their
12      slow release for years after initial ambient PM exposure may account partially for the observation
13      in epidemiologic studies that higher relative  risks are associated with long-term ambient PM
14      exposure than can be accounted for by additive effects of acute PM exposures. Alveolar region
15      clearance rates are decreased in human COPD sufferers and slowed by acute respiratory
16      infections, and the viability and functioning of alveolar macrophages are reduced in human
17      asthmatics and in animals with viral lung infections.  These observations suggest that persons
18      with asthma or acute lung infections are likely at increased risk for ambient PM exposure effects.
19           Differences in regional and total clearance rates between some  species reflect differences in
20      mechanical clearance processes. The importance of interspecies clearance differences is that
21      retention of deposited particles can differ between species and may result in differences in
22      response to similar PM exposures. Hsieh and Yu (1998) summarize  existing data on pulmonary
23      clearance of inhaled, poorly soluble particles in the rat, mouse,  guinea pig, dog, monkey, and
24      human. Two clearance phases, "fast" and "slow," in the A region are associated with mechanical
25      clearance along two pathways, the former with the mucociliary system and the latter with lymph
26      nodes.  Rats and mice are fast clearers, compared to other species.  Increasing initial lung burden
27      results in an increasing mass fraction of particles cleared by the slower phase. As lung burden
28      increases beyond 1 mg particles/g lung, the fraction cleared by the  slow phase increases to almost
29      100% for all species.  The rate for the fast phase is  similar in all species, not changing with
30      increasing lung burden, whereas the slow phase rate decreases with increasing lung burden.
31      At elevated burdens, the "overload" effect on clearance rate is greater in rats than in humans.

        April 2002                                 9-56        DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.8.3  Deposition and Clearance Patterns of Particles Administered by
 2             Inhalation Versus Intratracheal Instillation
 3           Inhalation is the most directly relevant exposure route for evaluating PM toxicity, but many
 4      studies deliver particles by intratracheal instillation. Because particle disposition is a determinant
 5      of dose, it is important to compare deposition and clearance of particles delivered by instillation
 6      versus inhalation. It is difficult to compare particle deposition and clearance among different
 7      inhalation and instillation studies because of differences in experimental methods and in
 8      quantification of particle deposition and clearance. Key points from a recent detailed evaluation
 9      (Driscoll et al., 2000) of the role of instillation in respiratory tract dosimetry and toxicology
10      studies are informative. In brief,  inhalation may result in deposition within the ET region, the
11      extent of which depends on the size of the particles used, but intratracheal instillation bypasses
12      this portion of the respiratory tract and delivers particles directly to the TB tree.  Although some
13      studies indicate that short (0 to 2  days) and long (100 to 300 days postexposure)  phases of
14      clearance of insoluble particles delivered either by inhalation or intratracheal instillation are
15      similar, others indicate that the percent retention of particles delivered by instillation is greater
16      than for inhalation, at least up to 30 days postexposure. Another salient finding is that inhalation
17      generally results in a fairly homogeneous distribution of particles throughout the lungs, but
18      instillation is typified by heterogeneous distribution (especially in the A region) and high levels
19      of focal particles. Most instilled material penetrates beyond the major tracheobronchial airways,
20      but the lung periphery is often virtually devoid of particles.  This difference is reflected in
21      particle burdens within macrophages, those from animals inhaling particles being burdened more
22      homogeneously and those from animals with instilled particles showing some populations of
23      cells with no particles and others  with heavy burdens, and is likely to impact clearance pathways,
24      dose to cells and tissues, and systemic absorption. Exposure method, thus, clearly influences
25      dose distribution that argues for caution in interpreting results from instillation studies.
26
27      9.8.4  Inhaled Particles as Potential Carriers of Toxic Agents
28           It has been proposed that particles also may act as carriers to transport toxic gases into the
29      deep lung.  Water-soluble gases, which would be removed by deposition to wet surfaces in the
30      upper respiratory system during inhalation, could dissolve in particle-bound water and be carried
31      with the particles into the deep lung. Equilibrium calculations indicate that particles do not
        April 2002                                 9-57         DRAFT-DO NOT QUOTE OR CITE

-------
 1      increase vapor deposition in human airways. However, these calculations do show that soluble
 2      gases are carried to higher generation airways (deeper into the lung) in the presence of particles
 3      than in the absence of particles. In addition, species such as SO2 and formaldehyde react in
 4      water, reducing the concentration of the dissolved gas-phase species and providing a kinetic
 5      resistence to evaporation of the dissolved gas.  Thus, the concentration of the dissolved species
 6      may be greater than that predicted by the equilibrium calculations. Also,  certain other toxic
 7      species (e.g., nitric oxide [NO], nitrogen dioxide [NO2], benzene, polycyclic aromatic
 8      hydrocarbons [PAH], nitro-PAH, a variety of allergens) may be absorbed onto solid particles and
 9      carried into the lungs.  Thus, ambient particles may play important roles not only in inducing
10      direct health impacts of their constituent components but also in facilitating delivery of toxic
11      gaseous pollutants or bioagents into the lung and may, thereby, serve as key mediators of health
12      effects caused by the overall air pollutant mix.
13
14      9.8.5 Summary of Particle Dosimetry
15           Although the current understanding of basic mechanisms of particle dosimetry,  clearance,
16      and retention has not changed since the 1996 PM AQCD, additional information has become
17      available on the role of certain biological determinants of these processes, such as gender and
18      age; and there has been an expansion of previous knowledge about the relationship between
19      regional deposition and translocation in regard to specific particle size ranges of significance to
20      ambient particulate exposure scenarios. There also has been significant improvement in the
21      mathematical and computational fluid dynamic modeling of particle dosimetry in the  respiratory
22      tract of humans.  Although the models have become more sophisticated and versatile, validation
23      of the models is still needed.
24           One of the areas that has improved since the 1996 PM ACQD is consideration of specific
25      and relevant ambient size particle ranges in deposition studies.  One such size mode is the nuclei
26      mode or ultrafine particles (< 0.1 //m).  While further information on respiratory deposition for
27      this size mode is still needed, there has been an improvement in the understanding of total
28      deposition as a function of particle size and breathing pattern and of certain aspects of regional
29      deposition of ultrafine particles. This new information indicates that the ET region, especially
30      the nasal passages, is a very efficient "filter" for these particles, reducing  the amount which
31      would be available for deposition in the TB and A regions of the respiratory tract.  Within the
        April 2002                                 9-58        DRAFT-DO NOT QUOTE OR CITE

-------
 1      thoracic region, the deposition distribution of ultrafme particles is highly skewed towards the
 2      proximal airway regions and resembles that of coarse particles.  In other words, deposition
 3      patterns of ultrafme particles are very much like those of coarse particles. Another example
 4      involves studies which attempt to evaluate the contribution of fine- and coarse-mode particles to
 5      deposition in various parts of the respiratory tract, although there have been only a few of these.
 6           It always has been clear that certain host factors affect deposition, and there has been
 7      improvement since the 1996 PM AQCD in the understanding of some of these factors,
 8      specifically gender, age, and health status. Recent information suggests that there are significant
 9      gender differences in the homogeneity of deposition as well as the deposition rate, and this  could
10      affect susceptibility. In regard to age, recent evaluations employed both mathematical models as
11      well as experimental studies, and most involved comparison of deposition in children compared
12      to adults. These studies generally indicate that children would receive greater doses of particles
13      per lung surface area than would adults. Unfortunately, deposition studies in another potentially
14      susceptible population, namely the elderly, are still lacking although there have been a number of
15      studies examining effects  of chronic pulmonary  disease on deposition.  These studies confirmed
16      that significant increases in deposition could occur in obstructed lungs.
17           Once deposited on airway surfaces, particles are subjected to translocation and clearance.
18      While the general pathways of clearance have been known for years, recent information has
19      improved the understanding of translocation of particles within size ranges which may be of
20      specific concern for ambient exposures. One such size mode, as noted above, is the ultrafme;
21      and recent studies indicate that ultrafme particles can be rapidly cleared from the lungs into the
22      systemic circulation and reach extrapulmonary organs.  This provides a mechanism whereby
23      inhaled particles may affect cardiovascular function, as noted in various epidemiological studies.
24           As with experimental studies, the major improvements in mathematical modeling of
25      dosimetry involve evaluation of realistic size modes for ambient conditions, as well as
26      improvements in the precision of these  models for more realistic depictions of respiratory tract
27      airflow patterns and detailed airway structures that may result in deposition "hot spots". These
28      improvements include more detailed evaluations of enhanced deposition at airway bifurcations,
29      use of parameters that allow determination of age differences in dosimetry, and improvement in
30      the modeling of clearance mechanisms.


        April 2002                                 9-59        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Thus, in general, while our understanding of specific aspects of particle dosimetry has
 2      improved since the 1996 PM AQCD, there are still areas in need of further evaluation. These
 3      include dosimetry in susceptible humans, better models for extrapolation between humans and
 4      animals used in inhalation studies, and better understanding of differences in the manner in
 5      which particles of different and relevant ambient size modes are handled following deposition.
 6      This latter research need is important for determining the potential of various particle types to
 7      exert effects systemically, rather than just locally within the respiratory tract.
10      9.9 ASSESSMENT OF PARTICULATE-MATTER PROPERTIES LINKED
11          TO HEALTH EFFECTS
12           What is the role ofphysicochemical characteristics ofparticulate matter in eliciting
13      adverse health effects?
14
15      9.9.1  Introduction
16          Ambient PM comprises a complex mix of constituents derived from many sources, both
17      natural and anthropogenic. Hence, the physicochemical composition of PM generally reflects the
18      major  contributing sources locally and regionally.  Within this framework of source or origin,
19      PM composition also varies significantly by the size-mode within which it is classified (ultrafme,
20      accumulation, or coarse).  It should be clear that any given particle can differ appreciably from
21      another individual particle of similar size, but that the region of origin with all of its contributing
22      sources determines the general composition of the generic PM in that classification mode. By its
23      nature then, exposure to airborne ambient PM constitutes an exposure to what is very clearly a
24      mixture of different particles of differing composition and to other gaseous co-pollutants that
25      coexist in that air-shed.
26          The epidemiology information reviewed in the 1996 PM AQCD and updated in this
27      document convincingly shows that a positive correlation exists between the levels of ambient PM
28      pollution and mortality/morbidity. However, this correlation is based mainly on a mass metric,
29      which is somewhat counter-intuitive considering the complexities in composition of PM and
30      given the perceptively low concentrations of most PM constituents, even when fractionated by
31      PM size. What has evolved since the 1996  PM AQCD is the advance in our understanding that
        April 2002                                9-60        DRAFT-DO NOT QUOTE OR CITE

-------
 1      the linkages between PM exposure and health impacts is most strongly related to accumulation
 2      mode particles, with combustion-derived PM typically being the most active of the source-based
 3      contributors. It is also appreciated that discovery of a "magic bullet" regarding PM
 4      physicochemical attributes is not likely to occur, and perhaps the sources from which the PM
 5      derive may be the best linkage one can achieve.
 6           Approaches to elucidating "causation" and "biological plausibility" have attempted to
 7      integrate the wealth of epidemiological data with the growing body of toxicology to reveal
 8      coherence among the findings to encourage the pursuit of sound hypotheses. Thus, while it is
 9      often difficult to  separate the physicochemical  attributes of PM that may be of health significance
10      from the mechanisms by which individual factor(s) may function in the response, a number of
11      hypotheses have  evolved espousing various PM characteristics as potentially significant
12      contributors to the observed health effects (reviewed by Dreher, 2000).  Each of the attribute-
13      based hypotheses has a sufficient data base to merit consideration and further investigation.
14      As the science progresses, it is important that any hypothesis be critically evaluated in the context
15      of the problem, and that the hypothesis provide reasonable responses to at least the following
16      generic, yet pertinent questions (Chapman et al., 1997).
17           •  Are there environmental sources that would lead to exposure to PM with the putative
18             constituent(s) or characteristic?
19           •  Is there evidence of personal exposure involving PM with that attribute and effect?
20           •  Does the  putative attribute possess or contribute to a toxic potential?
21           •  Is there evidence of an exposure-response relationship, especially at the low
22             concentrations found in the ambient environment?
23           •  How well does the hypothesis generalize from one PM sample, exposure, or locale to
24             another?
25           To date, toxicologic studies on PM have  provided important, albeit still limited, evidence
26      for specific PM attributes being primarily or essentially responsible for the cardiopulmonary
27      effects linked to ambient PM. In most cases, however, exposure concentrations in  laboratory
28      studies have been inordinately high compared to the exposures at which epidemiologic  studies
29      have found effects.  Reasons for this dosimetric discrepancy range from the limited numbers of
30      animals or human subjects that  can be practically studied, the uncertainty and narrow range of
31      responsiveness of the study groups and especially the typically limited use of young, elderly,

        April 2002                                9-61        DRAFT-DO NOT QUOTE OR CITE

-------
 1      unhealthy, or otherwise at-high-risk animals or humans, especially in light of poorly understood
 2      risk factors. Thus, most of the toxicology data-base resides in the "Hazard-Identification"
 3      compartment of the Risk Assessment paradigm.  However, sufficient coherence in the
 4      epidemiological and toxicological data has provided a level of "plausibility" to the observational
 5      studies and have opened new avenues for investigation to link PM properties and constituents to
 6      specific sources and to health outcomes. The primary PM properties thought to be related to
 7      health effects are discussed below.
 8
 9      9.9.2  Specific Properties of Ambient PM Linked to Health Effects
10      9.9.2.1 Physical Properties
11           Acid Aerosols:  There is relatively little new information on the effects of acid aerosols,
12      and the basic conclusions of the the 1996 PM AQCD remain unchanged. It previously was
13      concluded that acid aerosols cause little or no change in pulmonary function in healthy subjects,
14      but asthmatics may experience small  decrements in pulmonary function. Long-term exposures of
15      animals to acid aerosols,  on  the other hand, have been shown to alter airway morphology with
16      epithelial cell desquamation and an increase in secretory cells, but these changes have been
17      considered relatively minor. The conclusions about the acute health effects, however, are
18      supported by a recent study by Linn and colleagues (1997), in which healthy children (and
19      children with allergy or asthma) were exposed to sulfuric acid aerosol (100 //g/m3) for 4 hours.
20      While there were no significant effects on symptoms or pulmonary function when the entire
21      group was analyzed, the allergy group did have significant acid-related increases in symptoms,
22      although the acid concentrations were distinctly  higher than typical ambient concentrations.
23      These findings were consistent with those reported for adolescent asthmatics exposed to acid
24      aerosols in earlier studies reported in  the 1996 PM AQCD.
25           Although pulmonary effects of acid aerosols have been the subject of extensive research,
26      the cardiovascular effects of acid aerosols have received little attention. One example, which
27      raises the issue is a study of acetic acid fumes where reflex mediated increases in blood pressure
28      were found in normal and spontaneously hypertensive rats (Zhang et al., 1997).  Similarly, acidic
29      residual oil fly ash (ROFA) PM (which also contains a considerable amount of metal sulfates)
30      was found to alter ecocardiogram  (ECG) patterns in the same strain of rats at high air
31      concentrations (Kodavanti et al., 2000). Thus, acidic components should not be entirely
        April 2002                                9-62        DRAFT-DO NOT QUOTE OR CITE

-------
 1      dismissed as possible mediators of ambient PM health effects, since so little is known about
 2      potential cardiovascular impacts or impacts in compromised subjects.
 3
 4           Ultrafme Particles (Size. Surface Area. Number): The physical attributes of PM - size,
 5      surface area and number - are intimately interrelated. These properties influence lung deposition,
 6      penetrance and persistence in lung tissues, and systemic transport, and, in several studies,
 7      apparently the inherent toxicity of the particle itself. While a few epidemiological studies
 8      (Wichmann et al., 2000) show correlations between health outcomes and ultrafme (<100 nm)
 9      ambient PM, the bulk of the information regarding its toxic potential, and the role of surface
10      area, has derived from studies of surrogate insoluble particles, such as mineral oxides (e.g., TiO2)
11      and carbon black (Oberdorster et al., 1994; Osier and Oberdorster, 1997; Li et al., 1997, 1999).
12      These studies have shown that on an equivalent mass exposure-dose metric, ultrafme PM can
13      induce more acute lung injury than fine PM. Similarly, surrogate PM with high surface areas
14      induced more toxicity than those of like composition, but having smaller surface areas (Lison
15      et al., 1997).  On the other hand, studies have shown that composition also matters; for example
16      MgO ultrafmes produce less injury than ZnO (Kuschner et al., 1997), as did sparked carbon
17      versus similarly generated metal oxides (Elder et al., 2000).
18           As with acid aerosols, studies of ultrafme particles have focused largely on effects in the
19      lung, but inhaled ultrafme particles may also have the potential to be distributed systemically and
20      have effects that are independent of lung effects. Recent epidemiological studies evaluating
21      blood viscosity as a biologic correlate of ultrafme exposures, have reported slight increases that
22      raise the prospect of potential cardiovascular implications (Wichmann et al., 2000).
23
24           Fine and Thoracic Coarse Particles:  In contrast to ultrafme particles, the respective roles  of
25      fine (<2.5 //m) and thoracic coarse (2.5-10 //m) particles in defining health outcomes have
26      garnered considerable research attention because they are the most frequently measured size-
27      fractions of ambient PM and for which most health effects data exist.  The fine fraction
28      comprises most of the combustion-related constituents discussed below under chemicals and
29      most readily penetrates deeply into the respiratory tract - at least in terms of a mass metric dose.
30      Naturally, the fine fraction had greater surface area than the thoracic coarse fraction, but much
31      less surface area and particle number than the ultrafme fraction. To the extent that inhaled PM

        April 2002                                 9-63         DRAFT-DO NOT QUOTE OR CITE

-------
 1      may carry chemicals or reactive species on their surfaces, these smaller size fractions may have
 2      an additional dimension to their toxicity (in terms of surface chemical bioavailablilty) that is not
 3      found with coarse PM. For example, acute exposure to sulfate-coated carbon black was found to
 4      impair alveolar macrophage phagocytosis and intrapulmonary bactericidal activity in mice (Jakab
 5      et al., 1996; Clarke et al., 2000).  On the other hand, coarse PM usually is of mineral  (earthen) or
 6      biologic (discussed below) origin and, thus, has a less complex bioavailable chemical matrix than
 7      the finer PM mode. The relative toxicity of most earthen-derived PM has been observed to be
 8      less than that of the finer combustion-derived or surrogate ultrafme particles.  However, because
 9      ambient coarse PM would tend to impact on the airways of humans, it is thought this fraction
10      may be adverse to those with airways sensitivities or disease (e.g., asthma).
11
12      9.9.2.2  Chemical Properties
13           Inorganic Constituents:  The inorganic constituents of ambient PM comprise a number of
14      compounds and elements that derive from either natural or combustion sources.  The earthen or
15      natural  constituents of PM are  typically silicates that contain surface and matrix bound metals
16      such as calcium, magnesium, aluminum, and iron. As noted above, most of these silicates do not
17      appear to contribute much toxicity to ambient PM, as considered in this document.  Sulfate and
18      nitrate anions derived from combustion or photochemical processes usually complex with other
19      constituents in PM - often more water-soluble ammonium ions or organic acids, as well as
20      elemental cations, such as metals. The intrinsic, independent toxicities of sulfates (as per above)
21      and nitrates appear to be rather low, but they  may influence the toxicity or bioavailability of other
22      PM components.  Of the cations, metals represent a potential class of causal constituents for
23      PM-associated health effects that have received considerable attention (discussed in more detail
24      below).  Sulfate, nitrate, ammonium, and metals make up a substantial part of the mass of
25      ambient PM, often with a silicate or carbonaceous (see below) core, layering, or matrix. The
26      majority of PM-associated metals in fine PM are derived from stationary or mobile combustion
27      sources whereas particle sulfate,  nitrate and ammonium originate from secondary atmospheric
28      transformation reactions of involving SO2, NOX and biomass ammonia emissions. Organic PM
29      has both primary  and  secondary sources.
30


        April 2002                                9-64         DRAFT-DO NOT QUOTE OR CITE

-------
 1           Metals:  The 1996 PM AQCD relied on data from occupational exposures to initially
 2      evaluate the potential toxicity of metals in PM air pollution. Since that time, in vivo and in vitro
 3      studies using ROFA or soluble transition metals have contributed substantial new information on
 4      the health effects of PM-associated soluble metals.  The metals of most interest, notably the
 5      transition metals of iron, vanadium, copper, nickel, chromium, cadmium, arsenic, are ubiquitous
 6      constituents of PM-derived from anthropogenic fossil fuel emissions.  Exposure seems to be
 7      widespread with studies in autopsy specimens (1980's) showing dramatic increases in the content
 8      of the first row transition metals in lung tissues of Mexico City residents since the 1950's
 9      consistent with industrialization and pollution (Fortoul et al., 1996). Similar studies in North
10      America show metals in  the lung tissues of urban dwellers. Although there remain uncertainties
11      about the differential effects of one transition metal versus another, water-soluble or bioavailable
12      metals leached from ROFA or bulk ambient PM cause a variety of biological effects. Many
13      studies show that the action of instilled ROFA and constituent metals  are pro-inflammatory
14      (cells, mediators, and molecular signaling processes - in vivo and in vitro), and recently, they
15      have been shown to induce cardiac arrhythmias in animal models (both healthy and diseased).
16      In studies in which various ambient and emission source PM were instilled into rats, the soluble
17      metal content appeared to be the primary determinant of lung injury (Costa and Dreher, 1999).
18      However, these and the related findings on metal toxicity generally have derived from relatively
19      high dose instillation or inhalation exposures, lending them to criticism as to their relevancy for
20      ambient PM that is low in metal content.
21           Nevertheless, a series of studies associated with the closing of a  metal smelter in Utah
22      Valley, where ambient PM extracts (containing metals and other soluble constituents) were
23      instilled into the lungs of humans (Ohio and Devlin, 2001) and animals (Dye et al., 2001), as
24      well as tested in vitro (Frampton et al., 1999), showed remarkable coherence with
25      epidemiological studies of hospitalization and mortality (Pope,  1989; Pope et al., 1999b) in the
26      same area and at the same times of the PM samples used in the laboratory studies. The response
27      patterns in each study paralleled the metal  content.  Furthermore, recent application of novel
28      statistical approaches to the study of source-associated constituents (often metals are the
29      elemental markers) have shown promise in linking sources with their associated emission
30      profiles (including metals) to health outcomes in both humans (Laden et al., 2000)  and animals


        April 2002                                 9-65        DRAFT-DO NOT QUOTE OR CITE

-------
 1      (Clarke et al., 2000).  Thus, while metals appear to be one component involved in PM associated
 2      health effects, the full story is incomplete.
 3
 4           Organic Constituents:  Published research on the acute effects of PM-associated organic
 5      carbon constituents is conspicuous by its relative absence, except for diesel exhaust particles
 6      (DEP). Like metals, organics are common constituents of combustion-generated PM and are
 7      found in ambient PM samples over a wide geographical range. Organic carbon constituents
 8      comprise a substantial portion of the mass of ambient PM (10 to 60% of the total dry mass
 9      [Turpin, 1999]).  Although the organic fraction of PM is a poorly characterized heterogeneous
10      mixture of a widely varying number of different compounds, strategies have been proposed for
11      examining the health effects of potentially important organic constituents (Turpin, 1999).
12      In contrast, the mutagenic effects of ambient PM and evidence of DNA-adducts have had more
13      extensive study and have been linked to specific organic fractions  (Binkova et al., 1999; Chorqzy
14      et al.,  1994; Izzotti et al., 1996).  The extent to which organic constituents of ambient PM
15      contribute to adverse health effects identified by current epidemiology studies is not known.
16      Nevertheless, organic constituents remain of concern regarding PM health effects due in large
17      part to the contribution of DEP to the fine PM fraction and the health effects associated with
18      exposure to these particles.
19
20           Diesel Exhaust Particles (DEP): There is growing toxicological evidence that DEP
21      exacerbates the allergic response to inhaled antigens.  The organic fraction of diesel exhaust has
22      been linked to eosinophil degranulation and induction of cytokine production suggesting that the
23      organic constituents of DEP are responsible for the immune effects. It is known that the
24      adjuvant-like activity of DEP is not unique, and that certain metals have analogous adjuvant
25      effects (Lambert  et al., 2000). It is important to compare the immune effects of other source-
26      specific emissions, as well as concentrated ambient PM, to DEP to determine the extent to which
27      exposure to diesel exhaust may contribute to the incidence and severity of allergic rhinitis and
28      asthma.  Other types of noncancer and carcinogenic (especially lung cancer) effects are of
29      concern with regard to DEP exposures, as discussed in a separate EPA Health Assessment
30      Document for Diesel Exhaust (U.S. Environmental Protection Agency, 2002).
31

        April 2002                                 9-66        DRAFT-DO NOT QUOTE OR CITE

-------
 1           Biogenic Constituents: Recent studies support the conclusion of the 1996 PM AQCD that
 2     bioaerosols, at the concentrations present in the ambient environment, are unlikely to account for
 3     the health effects of ambient PM. Dose-response inhalation studies in healthy volunteers
 4     exposed to 0.55 and 50 //g endotoxin showed the threshold for pulmonary and systemic effects
 5     for endotoxin to be between 0.5 and 5.0 //g (Michel et al., 1997). Urban ambient air PM contains
 6     variable amounts of endotoxin, but the levels typically are several orders of magnitude less. The
 7     in vitro toxicological studies that have shown endotoxin associated with ambient PM to be pro-
 8     inflammatory, inducing cytokine expression in human and rat alveolar macrophages, appear to
 9     relate to the endotoxin dose to cell ratio (Becker et al., 1996; Dong et al., 1996). However,
10     endotoxin content does appear to vary by size-mode. Monn and Becker (1999) demonstrated
11     cytokine induction by human monocytes, characteristic of endotoxin activity, in the coarse size
12     fraction of outdoor PM, but not in the fine fraction. Interestingly, while studies in animals
13     models also require more endotoxin than typically found in ambient PM to induce inflammation,
14     recent studies suggest endotoxin may have a priming effect on PM-induced inflammatory
15     processes (Imrich et al., 1999).  Thus, the role of biogenic material like endotoxin  may have a
16     subtle role that is poorly understood.
17
18     9.9.2.3 Summary
19           Toxicological studies have provided considerable supportive evidence that certain
20     physicochemical particle attributes can provide elements of "causality" to observed health effects
21     of ambient PM.  A primary causative attribute may not exist but rather many attributes may
22     contribute to a complex mechanism driven by the nature of a given PM and its contributing
23     sources. The multiple interactions that may occur in eliciting a response in a host may make the
24     identification of any single causal component difficult and may account for the fact that mass as
25     the most basic metric shows the relationships to health outcomes that it does.
26
27     9.9.3  Chemical Components and Source Categories Associated with Health
28             Effects in Epidemiologic Studies
29           Epidemiologic studies using either individual chemical species or classes or using source
30     category factors (SCF)  derived  from factor analysis have identified a variety of species whose


       April 2002                               9-67        DRAFT-DO NOT QUOTE OR CITE

-------
 1     ambient concentrations are statistically associated with either total mortality or more specific
 2     mortality groupings.
 3
 4     9.9.3.1  Individual Chemical Species
 5           Table 9-10 lists the various gaseous co-pollutants, size fractions, chemical element or ions,
 6     and organic fractions that have been found to be associated with mortality in regressions using
 7     one pollutant species at a time.
             TABLE 9-10. CHEMICAL SPECIES ASSOCIATED WITH MORTALITY IN
                                     EPIDEMIOLOGIC STUDIES
Co-Pollutants
CO
NO2
SO2
03


PM Size Fractions
TSP
PM10
PM25
PM10.2.5
PM0,
number
Ions/Elements
SO4=
NO3
Ni
Pb


Carbon/Organic Fractions
TC (Total Carbon)
EC (elemental Carbon)
BC (Black Carbon)
COH (Coefficient of Haze)
OC Organic Carbon)
CX (Cyclohexene-extractable Carbon)
 1     9.9.3.2  Source Category Factors
 2           There are also three studies in which factor analysis has been used to identify several
 3     specific source category factors.  In two cases (Laden et al., 2000 and Tsai et al., 2000), the
 4     source category factors (SCF) were then used in a multiple regression, the nonsignificant factors
 5     were eliminated, and the multiple regression was rerun with only the significant factors. In the
 6     third case (Mar et al., 2000), relative risk values are reported for regression with SCF one at a
 7     time but the paper states that "Regression analysis with all of the factors included in a
 8     multi-source model produced similar results."  The similar results in single and multiple
 9     regressions and the low correlation between SCF indicates that there is low potential for
10     confounding among the various SCFs.

       April 2002                                9-68        DRAFT-DO NOT QUOTE OR CITE

-------
1           Source categories that have been found to be significantly associated (p < 0.05) with total,
2     cardiovascular, or cardiovascular plus respiratory mortality in one or more cities are shown in
3     Table 9-11.  A source category associated  with motor vehicles was found in all four studies. The
4     epidemiological studies do not provide sufficient information to determine whether the causal
5     factor is one or both of the gaseous co-pollutants (CO and NO2); soot particles from cars
6     (indexed by BS, COH, or EC); organic PM from vehicles, transition metals emitted by vehicle
7     (Mn, Fe, Zn); or other particles generated or resuspended by vehicular traffic.
           TABLE 9-11.  SOURCE CATEGORIES ASSOCIATED WITH MORTALITY IN
                                   EPIDEMIOLOGIC STUDIES
        Source Category                     Tracers
        Tsai et al. (2000)
           Motor vehicles                  CO
           Fuel Oil Combustion             Ni, V
           Sulfate                         S
           Industrial                       Zn, Cd
        Laden et al. (2000)
           Motor Vehicles                  Pb
           Coal Burning (sulfate)            Se, (S)
        Mar et al. (2000)
           Motor Vehicles                  CO, NO2; EC, OC; Mn, Fe, Zn, Pb
           Vegetative Burning              OC, non-soil K
           Sulfate                         S
        Ozkaynak et al. (1996)
           Motor vehicles                  CO, COH, NO2
1           The three studies that investigated multiple source categories also found a sulfate factor.
2     The factor reported by Laden et al. (2000) as "coal burning" contains high loadings of both
3     selenium and sulfur and could also have been called "regional sulfate". Mar et al. (2000) refer to
      April 2002                                9-69        DRAFT-DO NOT QUOTE OR CITE

-------
 1      the factor with high sulfate specifically as "regional sulfate".  They were able to make this
 2      connection because they also had a factor with a high loading of SO2 which they called a "local
 3      SO2" factor.  The regression with the chemical species S (assumed due to sulfate) was not
 4      significant, but the regression with the regional sulfate factor was  significant. This may be
 5      because the factor analysis will tend to remove other more localized sulfate sources such as
 6      CaSO4 and Na2SO4, leaving only acid sulfates ([NH4]2SO4, NH4HSO4, and H2SO4) for a regional
 7      sulfate factor. (In Phoenix, there was a modest loading of S in the soil factor.)  Therefore, all
 8      three sulfate factors should be considered as regional  sulfate.
 9           The studies of specific chemical components and source categories are especially important
10      because they indicate the association of health effects with the three major components of PM
11      mass: sulfate, nitrate, and organic PM. Examination of PM25 and nitrate effects, alone and in
12      multiple regressions, indicates that PM2 5 and nitrate were not confounded by NO2, CO or O3 in
13      Santa Clara, CA (Fairley, 1999).  Examination of the  lag structure from the Phoenix study reveals
14      that neither the regional sulfate factor nor the vegetative burning factor was confounded by NO2,
15      CO, SO2,  or O3.  The epidemiologic results suggest the need for toxicologic studies of the sulfate,
16      nitrate, and organic components of PM, including studies with compromised or susceptible
17      subjects.
18           All of the studies that investigated multiple source categories found a soil or crustal source
19      that was negatively associated with mortality.  This suggests that the components of natural soil
20      may have minimal toxicity unless contaminated by anthropogenic sources, such transition metals
21      or polyaromatic hydrocarbons.  In any event, the epidemiologic associations suggest additional
22      PM components that should be investigated in toxicologic studies.
23
24
25      9.10  SUSCEPTIBLE SUBPOPULATIONS
26           What subpopulations are at increased risk of adverse health outcomes from particulate
27      matter?
28
29
        April 2002                                9-70        DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.10.1  Introduction
 2           The 1996 PM AQCD identified several population groups potentially being at increased
 3      risk for experiencing health impacts of ambient PM exposure.  Elderly individuals (>65 years)
 4      were most clearly identified, along with those having preexisting cardiovascular or respiratory
 5      disease conditions.  Smokers and ex-smokers likely comprise a large percentage of individuals
 6      with cardiovascular and respiratory disease, e.g., chronic obstructive pulmonary disease (COPD).
 7      Individuals with asthma, especially children, also were identified as a potential susceptible
 8      population group. The studies appearing since the 1996 PM AQCD provide additional evidence
 9      to substantiate the above named groups as likely being at increased risk for ambient PM-related
10      morbidity or mortality effects.  There is even evidence, though quite limited at this time, of
11      prenatal effects on cardiac development and potential mortality impacts on infants in the first two
12      years of life.
13           While the identification of susceptible population groups is a critical element of the risk
14      paradigm, characterizing risk factors that underlie susceptibility and that may be common to
15      multiple groups would better substantiate risk estimates and provide better predictability to PM
16      responsiveness.  Information relating to these factors, as gleaned from recent epidemiology and
17      toxicology studies, suggests contributing host attributes that may be useful in gaining perspective
18      on their relative public health impact.
19
20      9.10.2  Preexisting Disease as a Risk Factor for Particulate Matter
21             Health Effects
22           The information reviewed in the 1996 PM AQCD is now augmented by  numerous new
23      studies which substantiate the finding that preexisting disease conditions represents an  important
24      risk factor for ambient PM health effects. Cardiovascular and respiratory diseases continue to
25      appear to be of greatest concern in relation to increasing risk for PM mortality and morbidity.
26      Indeed, the fact that these disease 'entities' often involve both organ systems, albeit to varying
27      degrees, might argue for their compilation under a broader classification of 'cardiopulmonary'
28      disease. Nevertheless, as they are diagnosed and reported separately, Table 9-12 shows the 1996
29      numbers of U.S. cases reported for COPD, asthma, heart disease, and hypertension.
30
31
        April 2002                                 9-71         DRAFT-DO NOT QUOTE OR CITE

-------
TABLE 9-12. INCIDENCE OF SELECTED CARDIORESPIRATORY DISORDERS BY AGE AND
                       BY GEOGRAPHIC REGION, 1996
to
o
o
to






oo
1
to


o
!>
H
6
O
0
H
O
O
O
o
H
W
(reported as incidence per thousand population and as number of cases in thousands)
Chronic Condition/Disease
COPD*
Incidence/1,000 persons
No. cases x 1,000
Asthma
Incidence/1,000 persons
No. cases x 1,000
Heart Disease
Incidence/1,000 persons
No. cases x 1,000
HD-ischemic
Incidence/ 1,000 persons
No. cases x 1,000
HD -rhythmic
Incidence/ 1,000 persons
No. cases x 1,000
Hypertension
Incidence/1,000 persons
No. cases x 1,000

All Ages

60.4
15,971

55.2
14,596

78.2
20,653

29
7,672

33
8,716
107.1
28,314

Under 45

50.6
9,081

58.9
10,570

33.1
5,934

2.5
453

24.3
4,358
30.1
5,391
Age
45-64

72.3
3,843

48.6
2,581

116.4
6,184

51.6
2,743

40.7
2,164
214.1
11,376

Over 65

95.9
3,047

45.5
1,445

268.7
8,535

140.9
4,476

69.1
2,195
363.5
11,547

Over 75

99.9
1,334

48.0
641

310.7
4,151

154.6
2,065

73.1
977
373.8
4,994
Regional
NE MW S W

57.8 67.6 59.4 56.6


61.8 56.6 51.8 52.9


88.5 78.0 77.0 70.4

28.9 30.0 30.7 25.0


40.2 34.0 28.1 32.9
109.3 108.2 113.5 93.7
'Total chronic bronchitis and emphysema.

Source: Adams etal. (1999).




















-------
 1      9.10.2.3 Ambient PM Exacerbation of Respiratory Disease Conditions
 2           Many time-series studies have shown that pre-existent chronic lung diseases as a group (but
 3      especially chronic obstructive pulmonary disease - COPD) constitutes a risk factor for mortality
 4      with PM exposure. Studies with humans that might reveal more specific data have been limited
 5      both ethically, as well as by the absence of good biomarkers of response (such as ECG's serve
 6      cardiac disease).  Measures of blood-gas saturation and lung function appear not to be
 7      sufficiently revealing or sensitive to mild physiologic changes in those with moderate disease
 8      conditions who might be amenable to lab study. In the field, assessing the degree of underlying
 9      disease and how that relates to responsiveness of these biomarkers is unclear. However, subjects
10      with COPD and asthma have been studied with inert aerosols for the purpose of assessing
11      distribution of PM within the lung, and it is now quite clear that airways disease leads to very
12      heterogeneous distribution of PM deposited within the lung. Studies have shown up to 10-fold
13      higher than normal deposition at airway bifurcations,  thus creating "hot-spots" that may well
14      have biologic implications, especially if the individual already has diminished function or other
15      debility due to the underlying disease, even CVD.  Thus the  dosimetry of PM within the lung
16      must be considered an important element of the susceptibility paradigm with most any
17      cardiopulmonary disease condition.
18           There are several reports of associations between short-term fluctuations in ambient PM
19      and day to day frequency of respiratory illness. In most cases, notably in children and young
20      people, exacerbation of preexisting respiratory illness and related symptoms has been assessed
21      rather than de novo acute respiratory infections, with asthma apparently an additional  risk factor.
22      The use of inhalers has also been shown to increase in many young asthmatics in response to air
23      pollution, with PM often noted as the primary correlate, and as a result school absenteeism
24      increases, again especially in asthmatic children.  Interestingly, acute respiratory infections in the
25      elderly with cardiopulmonary disease appears to result in complications of underlying cardiac
26      disorders when PM exposure is involved (Zanobetti et al., 2000), and likewise is linked to
27      subsequent hospitalization. Animal studies with surrogate PM, however, show varied impact on
28      the induction of infection, but in general can alter lung phagocyte functions, which might worsen
29      the condition. Thus, while there appears to a strong likelihood that infections may be worsened
30      by exposure to PM, general statements regarding interaction of PM with response to infectious


        April 2002                                 9-73        DRAFT-DO NOT QUOTE OR  CITE

-------
 1      agents are difficult given the unique attributes of various infectious agents and the immune status
 2      of the host.
 3           The underlying biology of lung diseases might also lead to heightened sensitivity to PM
 4      (apart from the dose issue noted above), but this attribute of disease remains hypothetical in the
 5      context of PM. The functional linkages with the cardiac system for maintenance of adequate gas
 6      exchange and fluid balance notwithstanding, the role of inflammation in the diseased respiratory
 7      tract (airways and alveoli) could play a key role. Studies in animals genetically or exogenously
 8      altered to induce inflammation are sometimes intrinsically more responsive to surrogate or
 9      concentrated ambient PM. While a PM-induced response may on the one hand be cumulative
10      with the underlying injury or condition, the responses may, on the other hand, be magnified by
11      any number of mechanisms that are poorly understood. There is sufficient basic biological data
12      to hypothesize that the exudated fluids in the airspaces may either interact differently with
13      deposited PM (e.g., to generate oxidants - Costa and Dreher,  1999; Ohio et al., 2001) to augment
14      injury, or predispose the lung (e.g., sensitize receptors - Undem and Carr, 2002) to enhance the
15      response to a stereotypic PM stimulus through otherwise normal pathways. Less appreciated is
16      the loss of reserve - functional or biochemical - where the susceptible individual is incapable of
17      sufficient compensation (e.g., antioxidant responses - Kodavanti et al., 2000). Any of these or
18      related mechanisms may contribute to "susceptibility" and may indeed be a common factor that
19      can be attributable to other susceptible groups.  Understanding these will ultimately aid in
20      addressing true risk of susceptible groups to PM.
21           Again, even a small percentage reduction in PM health impacts on respiratory-related
22      diseases could calculate out to a large number of avoided cases. In 1997, there were 3,475,000
23      U.S. hospital discharges for respiratory diseases: 38% for pneumonia, 14% for asthma, 13% for
24      chronic bronchitis, 8% for acute bronchitis, and the remainder not specified (Lawrence and Hall,
25      1999).  Of the 195,943 deaths recorded as caused by respiratory diseases, 44% resulted from
26      acute infections, 10% from emphysema and bronchitis, 2.8% from asthma, and 42% from
27      unspecified COPD (Hoyert et al., 1999).
28
29      9.10.2.4 Ambient PM Exacerbation of Cardiovascular Disease Conditions
30           Exacerbation of cardiovascular (CVD) has been associated epidemiologically, not only
31      with ambient PM, but also with other combustion-related ambient pollutants such as CO. Thus,

        April 2002                                9-74       DRAFT-DO NOT QUOTE OR CITE

-------
 1      while leaving little doubt that ambient PM exposures importantly affect CVD mortality and
 2      morbidity, the quantitation of the proportion of risk for such exacerbation specifically attributable
 3      to ambient PM exposure is difficult.  Recent studies (e.g., concentrated ambient particle studies
 4      [CAPS]) have demonstrated cardiovascular effects in response to ambient particle exposures, and
 5      studies utilizing animals and  other approaches also have produced results suggesting plausible
 6      mechanisms leading to cardiovascular effects. However, much remains to be resolved with
 7      regard to delineation of dose-response relationships for the induction and extrapolation of such
 8      effects to estimate appropriate and effective human equivalent PM (or specific constituent/s)
 9      exposures.
10           The recent appreciation for underlying cardiovascular dysfunction as a risk factor for PM
11      health effects derives from a  growing and diverse body of literature.  While many time-series
12      studies have revealed stronger associations between PM exposures and mortality when a
13      subpopulation was segregated for pre-existent cardiac disease, no direct and plausible evidence
14      had been available.  However, recent panel studies of human subjects with CVD (Peters et al.,
15      2000) have shown correlations between air pollution levels, notably PM, and intervention
16      discharge frequency of implanted cardiac defribrillators. Analogously, Pope and colleagues
17      (2001) have noted altered autonomic control of cardiac electrocardiograms (in terms of Heart
18      Rate Variability) over a wide age- range of ostensibly healthy subjects when they were
19      introduced into a room with active smokers.  Evidence of vascular narrowing with exposure to
20      concentrated ambient PM (CAPS) has likewise been reported suggesting parallel cardiovascular
21      responses (Brook et al., 2002).  Collectively, these and previous studies that have shown ambient
22      PM-induced alterations in cardiac physiology (Pope et al, 1999a,b; Liao et al., 1999; Peters et al.,
23      1999a;  Gold et al., 2000) in human subjects, complemented with animal studies (Godleski et al.,
24      1996; Watkinson et al., 1998, 2001; Kodavanti et al., 2000), reinforce the notion of significant
25      cardiac responses to PM. Moreover, indications of changes in plasma viscosity (Peters et al.,
26      1997a) and other factors involved in clotting function (Ohio et al., 2000) provide a plausible
27      cascade of events that could culminate in a sudden cardiac events in some individuals.
28           The HEI report on an epidemiologic study in Montreal, Canada by Goldberg et al. (2000),
29      provides interesting new information regarding types of medical conditions potentially
30      predisposing susceptible individuals to increased risk for PM-associated mortality.   It is
31      specifically suggestive that other diseases involving cardiovascular complications could also

        April 2002                                 9-75        DRAFT-DO NOT QUOTE OR CITE

-------
 1      contribute to PM risk. First, the immediate causes of death, as listed on death certificates, were
 2      evaluated in relation to various ambient PM indices (TSP, PM10 estimated PM2 5, COH, sulfates,
 3      and extinction coefficients) lagged for 0 to 4 days. Significant associations were seen between
 4      each of the PM measures and total nonaccidental deaths, respiratory diseases, and diabetes, with
 5      an approximate 2% increase in excess nonaccidental mortality being observed per 9.5 //g/m3
 6      interquartile increase in 3-day mean estimated PM2 5 exposure. When underlying clinical
 7      conditions identified in the decedents' medical records were then evaluated in relation to ambient
 8      PM measures, all three measures (COH, sulfate, and estimated PM2 5) were associated with acute
 9      lower respiratory disease, congestive heart failure, and any cardiovascular disease. Predicted
10      PM25 and COH also were reported to be associated with cancer, chronic coronary artery disease,
11      and any coronary artery disease, whereas sulfate was associated with acute and chronic upper
12      respiratory disease. None of the three PM measures were related to airways disease, acute
13      coronary artery disease, or hypertension. These results both tend to confirm previous findings
14      identifying those with preexisting cardiopulmonary diseases as being at increased risk for
15      ambient PM effects and implicate another possible risk factor, diabetes (which involves
16      cardiovascular complications as it progresses), as a potential susceptibility condition putting
17      individuals at increased risk for ambient PM effects. Zanobetti and Schwartz (2001) have
18      likewise found, perhaps more  directly, that those with diabetes are at increased risk, presumably
19      related to the cardiac and vascular complications associated with this disease.
20           To the extent that the observed associations between ambient PM and heart disease
21      exacerbation are causal and specific, the impact on public health could be dramatic.  In 1997,
22      there were about 4,188,000 U.S. hospital discharges with heart disease as the first-listed
23      diagnosis (Lawrence and Hall, 1999). Among these, about 2,090,000 (50%) were for ischemic
24      heart disease, 756,000 (18%) for myocardial infarction or heart attack (a subcategory of ischemic
25      heart disease), 957,000 (23%) for congestive heart failure, and 635,000 (15%) for cardiac
26      dysrhythmias. Also, there were 726,974 deaths from heart disease (Hoyert et al., 1999).  Thus,
27      even a small percentage reduction in PM-associated admissions or deaths from heart disease
28      would predict a large number of avoided cases.
29
30


        April 2002                                 9-76       DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.10.3  Age-Related At-Risk Population Groups:  The Elderly and Children
 2           The very young and the very old apparently constitute another group especially affected by
 3      PM air pollution. As noted above, a major factor in increased susceptibility to air pollution is the
 4      presence of a preexisting illness, as discussed by Zanobetti and Schwartz (2000).
 5           The impact of PM pollution is well-documented in time-series studies with mortality risk in
 6      studies where age is a factor in the analysis, risk increases above the age of 45 and continues to
 7      increase significantly throughout the remainder of life. Cardiopulmonary diseases more common
 8      to the elderly play into the risk within older age groups, but panel studies of morbidity focusing
 9      on generally healthy people in retirement homes or elderly volunteers exposed to concentrated
10      ambient PM in chambers show subtle alterations of autonomic control of cardiac function (i.e.,
11      slight depression of heart rate variability) and blood factors concordant with a putative response
12      to ambient PM levels. Though small, these changes are considered clinically significant based on
13      studies of risk in cardiac patients and general population  studies of cardiac disease progression.
14      Moreover, these changes are in contrast to the lack of similar physiologic changes in healthy
15      young people.  Over the long term, innate differences in metabolism or other mechanisms may
16      impact the likelihood of chronic outcomes, e.g., COPD or lung cancer.  To what extent
17      progression occurs with repeated PM exposures and how much disease or other risk factors add
18      to or complicate the magnitude of response remains uncertain.
19           Although infection as a risk factor for PM has already been discussed, it is important to
20      emphasize that there are clear  age differences in both the incidence and type of infections across
21      age groups.  Young children have the highest rates of respiratory illnesses related to infection
22      (notably respiratory synctial virus), while adults are affected by other infectious agents such as
23      influenza that may also lend susceptibility to PM.  Data to fully address the importance of these
24      differences is incomplete. The distribution of infectious  lung diseases in the U.S. in 1996,
25      summarized in the Table 9-13, provides a good overview of the diversity of this category of
26      preexisting lung disease.
27           In  addition to their higher incidences of preexisting respiratory conditions, several other
28      factors may render children and infants more susceptible to PM exposures, including more time
29      spent outdoors, greater activity levels and ventilation, higher doses per body weight and lung
30      surface area, and the potential  for irreversible effects on the developing lung.  For example, PM
31      doses on a per kilogram body weight basis are much higher for children than for adults as is
        April 2002                                 9-77        DRAFT-DO NOT QUOTE OR CITE

-------
               TABLE 9-13.  NUMBER OF ACUTE RESPIRATORY CONDITIONS PER
                     100 PERSONS PER YEAR, BY AGE: UNITED STATES, 1996
45 Years and Over
Type of Acute Condition
Respiratory Conditions
Common Cold
Other Acute Upper Respiratory
Infections
Influenza
Acute Bronchitis
Pneumonia
Other Respiratory Conditions
All
Ages
78.9
23.6
11.3
36.0
4.6
1.8
1.7
Under 5
Years
129.4
48.6
13.1
53.7
*7.2
*3.9
*2.9
5-17
Years
101.5
33.8
15.0
44.3
4.3
*1.7
*2.4
18-24
Years
86.0
23.8
16.1
40.5
*3.9
*1.4
*0.4
25-44
Years
76.9
18.7
11.6
38.1
5.1
*1.3
*2.0
Total
53.3
16.1
7.0
23.3
3.8
*2.0
*1.1
45-64
Years
55.9
16.4
7.5
26.1
3.5
*0.9
*1.5
65 Years
and Over
49.0
15.7
6.1
18.6
*4.4
*3.8
*0.5
         Source: Adams etal. (1999).
 1     displayed graphically in Figure 9-23. The amount of air inhaled per kilogram body weight
 2     decreases dramatically with increasing age, due in part to ventilation differences (in cubic meters
 3     per kilogram a day) of a 10-year-old being roughly twice that of a 30-year-old person, even
 4     without the consideration of activity level. Child-adult dosage disparities are even greater when
 5     viewed on a per lung surface-area basis.
 6          As to potential lung developmental impacts of PM, there exist both experimental and
 7     epidemiologic data, which although limited, suggest that the early post-neonatal period of lung
 8     development is a time of high susceptibility for lung damage by environmental toxicants.
 9     In experimental animals, for example, elevated neonatal susceptibility to lung-targeted toxicants
10     has been reported at doses "well below the no-effects level for adults" (Plopper and Fanucchi,
11     2000); and acute injury to the lung during early postnatal development may impair normal repair
12     processes, such as down-regulation of cellular proliferation (Smiley-Jewel et al., 2000, Fanucchi
13     et al., 2000). These results in animals appear concordant with recent findings for young  children
14     growing in the Los Angeles area where both oxidants and high PM prevail (Gauderman et al.,
15     2000).
16          These and other types of health effects in children are emerging as potentially more
17     important than appreciated in the 1996 PM AQCD. Unfortunately, relatively little is known
       April 2002
9-78
DRAFT-DO NOT QUOTE OR CITE

-------
                     0.6
                     0.5
                 E   0.4
                 (D
                 -I— »
                 as
                 03
                     0.3
                 -^   0.2
                           T	
                     0.1
                          0
 I
10
 i
20
30
40
 i
50
60
 i
70
80
                                                   Age (y)
       Figure 9-23. Inhalation rates on a per body-weight basis for males (•) and females
                    by age (Layton, 1993).
 1     about the relationship of PM to these and other serious health endpoints (low birth weight,
 2     preterm birth, neonatal  and infant mortality, emergency hospital admissions and mortality in
 3     older children). The recent report by Ritz et al. (2002) linking CO exposures of mothers in
 4     Los Angeles with fetal  cardiac defects raises concerns for PM, which was inconclusively linked
 5     in the study.  Similarly, little is yet known about the involvement of PM exposure in the
 6     progression from less serious childhood conditions, such as asthma and respiratory symptoms, to
 7     more serious disease endpoints later in life. Thus, the loss of productive life-years that add to the
 8     costs to society may be more than just those indexed by PM-related mortality and/or hospital
 9     admissions/visits.
10          In summary, host variability may come to be the most important factor in determining the
11     response profile of any population exposed to PM.  Studies to date suggest that certain
12     subpopulations are indeed more acutely responsive to PM, perhaps due to differences in lung
       April 2002
                 9-79
                      DRAFT-DO NOT QUOTE OR CITE

-------
 1      deposition (either in terms of dose and/or intrapulmonary distribution) or other biologic aspects
 2      of the cardiopulmonary system or disease thereof.  The role of innate attributes of risk grounded
 3      in one's genetic code is largely unknown but potentially of great importance. Animal models
 4      have been used to show clear differences in response to PM and other pollutants, and the critical
 5      involvement of varied genes in the induction of asthma, emphysema, and many other ailments is
 6      widely accepted, but poorly understood.
 7
 8
 9      9.11  MECHANISMS OF INJURY
10           What are the mechanisms by which acute exposure to PM causes adverse health effects?
11
12           Numerous epidemiologic studies have shown statistically significant associations between
13      ambient PM levels and a variety of human health endpoints, including mortality, hospital
14      admissions, emergency department visits, respiratory illness, and symptoms measured in
15      community  surveys. These associations have been observed with both short and long-term PM
16      exposure. There was little information available in the 1996 PM AQCD to provide biologically
17      plausible mechanisms to support the epidemiologic observations. However, in the intervening
18      years significant progress has been made in identifying pathophysiological effects in humans and
19      animals exposed to various PM that can provide insight into the mechanisms by which PM may
20      exert its effects.  Potential mechanisms include neural mechanisms affecting the autonomic
21      nervous system (ANS) via direct pulmonary reflexes or through pulmonary inflammatory
22      processes, direct effects of PM or its components on ion channel function in myocardial cells,
23      ischemic responses of the myocardium, or systemic responses including inflammation that can
24      trigger endothelial cell dysfunction,  and thrombosis via alterations in the coagulation cascade.
25      The interactions between these pathways which may lead to sudden cardiac death is shown in the
26      Figure 9-24. However, it must be noted that PM is a complex mixture of many different
27      components and it is possible that different components may stimulate different mechanistic
28      pathways. Thus exposure to PM may result in one or more pathways being activated, depending
29      on the chemical and physical makeup of the PM.
30

        April 2002                               9-80       DRAFT-DO NOT  QUOTE OR CITE

-------


, ,


Pulmonary Reflexes 9 Pulmonary Inflammation
i
Autonomic Nervous

System ~^*C Heart ) •«— Systemic Inflammation 	
/~^ \
* Endothe ial Cell
Conduction/Repolarization Dysfunction

1 1 1
Heart Rate Cardiac Rhythm Plaque
1 1
Brad yea rdia Thro
Tachycardia *
Ventricular Fibrillation ^V /

Sudden Cardiac
Death
i
^ Platelet
Activation
1
Rupture
1
.




Clotting
F acto rs

Viscosity

       Figure 9-24.  Schematic representation of potential pathophysiological pathways and
                     mechanisms by which ambient PM may increase risk of cardiovascular
                     morbidity and/or mortality.
 1          There is now ample evidence that inhaled particles can affect the heart through the ANS.
 2     Direct input from the lungs to the ANS via pulmonary afferent fibers can affect both heart rate
 3     (HR) and heart rate variability (HRV). The heart is under the constant influence of both
 4     sympathetic and parasympathetic innervation from the ANS; and monitoring changes in HR and
 5     HRV can provide insight into the balance between those two ANS subdivisions. During recent
 6     decades a large clinical database has developed describing a significant relationship between
 7     autonomic dysfunction and sudden cardiac death. One measure of this dysfunction, low HRV,
 8     has been implicated as a predictor of increased cardiovascular morbidity and mortality.  Several
 9     independent epidemiologic panel studies of elderly volunteers (some having cardiovascular or
10     pulmonary disease) have reported associations between PM concentrations and various measures
11     of HR and HRV. Although there are some differences among the studies, in general they report
12     an association between PM levels and a reduction in the standard deviation of normal to normal
13     beat intervals (SDNN), a time-domain variable of which the reduction was associated in the
       April 2002
9-81
DRAFT-DO NOT QUOTE OR CITE

-------
 1      Framingham Heart Study with a higher risk of death. Some studies also reported an association
 2      between PM and decreased HRV in the high frequency (HF) range, which is a reflection of
 3      parasympathetic modulation of the heart. Other studies have reported a positive association
 4      between PM and HR; elevated HR has been associated with hypertension, coronary heart disease,
 5      and death. Thus taken as a whole, evidence from panel studies indicates that PM can directly
 6      affect the ANS in such as way as to alter heart rate and heart rate variability. However, it should
 7      be noted that lowered HRV has primarily been used as a predictor of subsequent increased
 8      mortality and morbidity.  It is not clear whether a single reversible acute change in HRV places a
 9      person more at risk for an immediate adverse cardiac event. Whether changes in HRV associated
10      with exposure to PM represent an independent risk or is just a marker of exposure is not yet
11      known.
12           PM as also been shown to induce changes in conductance and repolarization of the heart as
13      well. Repolarization duration and morphology may reflect subtle changes in myocardial
14      substrate and vulnerability governed by changes in ion channel function.  There is considerable
15      evidence linking changes in T wave morphology, QT and T wave variability, T wave Alternans,
16      and changes in ST segment height, to the risk of sudden death. In some studies, rodent models of
17      susceptibility (monocrotaline injected,  spontaneously hypertensive) exposed to ROFA showed
18      exacerbated ST segment depression, a factor reflecting T wave morphology during repolarization
19      and which as been useful in diagnosing patients with ischemic heart disease. Healthy dogs
20      exposed to CAPS also showed changes in ST segment elevation; this was exacerbated in dogs
21      with coronary artery occlusion.
22           While PM-induced changes in HRV and HR, as well as changes associated with
23      repolarization and conductance, have the potential to progress to malignant arrhythmias, there is
24      now evidence from both human and animal studies that PM exposure may be linked with severe
25      events directly associated with sudden cardiac death. A recent epidemiology study of patients
26      with implanted cardiac defibrillators reported associations between PM and increased
27      defibrillator discharges.  Presumably, some of these patients would have suffered a fatal event
28      had they not had an implanted defibrillator. A second study reported that the risk for myocardial
29      infarction (MI) onset increased in association with PM levels in the 2 hours preceding the MI.
30      PM exposure has also been linked with malignant arrhythmia in some toxicology studies.
31      Healthy rodents exposed to ROFA demonstrated an increase in serious arrhythmic events,

        April 2002                                9-82        DRAFT-DO NOT QUOTE OR CITE

-------
 1      including bradycardia.  Rats treated with monocrotaline had significantly exacerbated
 2      arrhythmias, and some animals even died within 24 hours following exposure.  Older rats,
 3      exposed to both ROFA and PM collected from Ottowa, also experienced increased arrhythmias.
 4      Dogs exposed to CAPS experienced a slight bradycardia following exposure.  Some of these
 5      studies involved instillation of a specific PM component (ROFA) at high concentrations, making
 6      it uncertain that these observations would hold true using ambient PM at more realistic
 7      concentrations. Nevertheless, at least one study used ambient particles collected from Ottowa,
 8      and other studies exposed animals by inhalation to CAPS.  Taken as a whole, these studies
 9      provide convincing evidence that exposure of animals to high levels of PM can affect
10      conductance and repolarization, potentially leading to fatal arrhythmias. However, it remains to
11      be  seen if these mechanisms, that can potentially explain acute mortality associated with PM
12      exposure, operate at the lower concentrations of ambient PM to which most people are exposed.
13           Particulate matter could potentially affect the ANS by direct interaction with nerve ending
14      in the lung, or indirectly through the production of inflammatory mediators. Numerous studies
15      have  documented that exposure of rodents to ROFA results in substantial lung inflammation and
16      injury. However, due to the levels of ROFA used in many of these studies and the fact that
17      ROFA only makes up a small portion of most airsheds,  studies with ambient air particles may be
18      more relevant.  There are several studies in which humans, dogs, or rodents have been exposed to
19      CAPS and mild pulmonary inflammation observed. Other studies have shown similar effects
20      when ambient PM collected on filters was used. However, the level of inflammation was quite
21      low in most of these studies, certainly lower than reported in humans or animals exposed to
22      ozone, and it is not yet clear whether lung inflammation plays a role in PM-induced changes in
23      the ANS.
24           In addition to affecting the ANS via the lung, it is also possible that PM or its components
25      could directly attack the myocardium. There is substantial evidence that chronic exposure to
26      fibers encountered in the workplace (e.g., asbestos) result in deposition of fibers in organs other
27      than the lung. Some recent studies have suggested that ultrafme PM may exit the lung and
28      deposit in other organs, including the liver and heart. So far these studies have used sources of
29      particles not naturally found in the air (e.g., silver colloid, latex) so it is not yet clear to what
30      extent PM actually leaves the lung or, if it does, how it interacts directly with the heart.
31      However, there is some evidence of direct changes in the myocardium following PM exposure.

        April 2002                                9-83        DRAFT-DO NOT QUOTE OR CITE

-------
 1      For example, rats exposed to ROFA, which is made up mostly of soluble transition metals, have
 2      increased pro-inflammatory cytokine expression in the left ventricle.  In another study, dogs
 3      living in highly-polluted Mexico City had histopathology changes in heart tissue compared with
 4      dogs living in areas with low air pollution.  Substantial deposits of particulate matter could be
 5      seen throughout the myocardium in the Mexico City dogs.  Though preliminary, these
 6      observations point to a need for additional work to better define PM-induced changes in
 7      myocardial tissue.
 8           Acute coronary events frequently occur as a result of thrombus formation in the site of a
 9      ruptured atherosclerotic plaque. Increased levels of clotting and coagulation factors, platelet
10      aggregability, and blood viscosity, together with reduced fibrinolytic activity and endothelial cell
11      dysfunction can promote a pro-coagulant state which could potentially contribute to thrombus
12      formation. C reactive protein, a marker of systemic inflammation which correlates with some
13      cardiac events, is positively associated with PM in several panel studies. Some of these studies
14      also report associations between PM and enhanced blood viscosity or increased fibrinogen, a
15      known risk factor for ischemic heart disease. Controlled human and animal exposure studies
16      have also reported that exposure to CAPS (in humans) or ROFA (in animals) results in increased
17      levels of blood fibrinogen.  These studies suggest that PM may alter the coagulation pathways in
18      such a way as to trigger cardiovascular events in susceptible individuals.
19           Panel studies have also reported associations between PM and changes in white blood cells,
20      although these findings are not easy to interpret since some studies report positive  associations
21      while others report negative associations. Animal studies are similarly unclear, with some
22      studies (rodents exposed to CAPS) reporting increased numbers of blood platelets  and white
23      blood cells and others (rodents exposed to ROFA) reporting decreased numbers of white blood
24      cells.  In one study, rabbits  instilled with colloidal  carbon had an increase in neutrophils released
25      from the bone marrow.  The same research group found an association between PM and elevated
26      band neutrophil counts (a marker for bone marrow precursor release) in humans exposed to high
27      levels of carbon from biomass burning during the 1997 Southeast Asian smoke-haze episodes.
28           Endothelial cell dysfunction may contribute to  myocardial ischemia in some  susceptible
29      populations.  The vascular endothelium secretes multiple factors that control vascular tone,
30      modulate platelet activity, and influence thrombogenesis.  A recent study has reported endothelial
31      cell dysfunction in humans exposed to CAPS, as measured by dilation of the brachial artery.

        April 2002                                 9-84        DRAFT-DO NOT QUOTE OR CITE

-------
 1      This vasoconstriction could be caused by an increase in circulating endothelin-1, which has been
 2      described in rats exposed to PM.
 3           Taken as a whole, these studies are difficult to interpret but clearly indicate that PM can
 4      affect the circulatory system.  However, a complete understanding of the pathways by which very
 5      small concentrations of inhaled ambient PM can produce vascular changes that can contribute to
 6      increased mortality/morbidity remains to be more fully elucidated.
 7
 8
 9      9.12 HEALTH EFFECTS OF AMBIENT PARTICULATE MATTER
10           OBSERVED IN POPULATION STUDIES
11           How are exposures to ambient PM quantitatively related to increased risks of health effects
12      (mortality/morbidity) among  general human populations and susceptible subgroups ?
13
14      9.12.1 Introduction
15           This section assesses available scientific evidence regarding the physiologic and health
16      effects of exposure to ambient PM as observed in epidemiologic (human population) studies.
17      The main objectives of this evaluation are (1) to summarize and evaluate strengths and
18      limitations of available epidemiologic findings; (2) to summarize quantitative relationships
19      between ambient PM exposures and increased human health risks; (3) to assess the biomedical
20      coherence of findings across  studied endpoints; and (4) to note the increased biologic plausibility
21      of the available epidemiologic evidence in light of (a) linkages between specific PM components
22      and health effects and (b) various dosimetric, mechanistic, and pathophysiologic considerations
23      discussed earlier in this chapter.
24           Numerous epidemiologic studies have shown statistically significant associations of
25      ambient PM levels with  a variety of human health endpoints, including mortality, hospital
26      admissions,  emergency department visits, other medical visits, respiratory illness and symptoms
27      measured in community surveys, and physiologic changes in pulmonary function.  Associations
28      have been consistently observed between both short- and long-term PM exposure and these
29      endpoints. The general internal consistency of the epidemiologic database and available findings
30      demonstrate well that notable human health effects are associated with exposures to ambient PM
31      at concentrations currently found in many geographic locations across the United States.
        April 2002                                9-85        DRAFT-DO NOT QUOTE OR CITE

-------
 1      However, many difficulties still exist with regard to delineating the magnitudes and variabilities
 2      of risk estimates for ambient PM, the ability to attribute observed health effects to specific PM
 3      constituents, the time intervals over which PM health effects are manifested, the extent to which
 4      findings in one location can be generalized to other locations, and the nature and magnitude of
 5      the overall public health risk imposed by ambient PM exposure.
 6           The etiology of most air pollution-related health outcomes is highly multifactorial, and the
 7      impact of ambient air pollution exposure on these outcomes is often small in comparison to that
 8      of other etiologic factors (e.g., smoking). Also, ambient PM exposure usually is accompanied by
 9      exposure to many other pollutants, and PM itself is composed of numerous physical/chemical
10      components. Assessment of the health effects attributable to PM and its constituents within an
11      already-subtle total air pollution effect is difficult even with well-designed studies. Indeed,
12      statistical partitioning of separate pollutant effects may not characterize fully the etiology of
13      effects that actually depend on simultaneous exposure to multiple air pollutants. In this regard,
14      several viewpoints existed at the time of the 1996 PM AQCD regarding how best to interpret the
15      epidemiology data: one saw the PM exposure indicators as surrogate measures of complex
16      ambient air pollution  mixtures and the reported PM-related effects as representative of those of
17      the overall mixture; another held that reported PM-related effects are attributable to PM
18      components (per se) of the air pollution mixture and reflect independent PM effects; and a third
19      viewpoint holds that PM can be viewed both as a surrogate indicator, as well as a specific cause
20      of health effects.
21           Several other key issues and problems also must be  considered when attempting to interpret
22      the data reviewed in this document. For example, although the epidemiology data provide strong
23      support for the associations mentioned  above, questions remain regarding potential underlying
24      mechanisms.  Although much progress has been made toward identification of anatomic sites at
25      which particles trigger specific health effects and elucidation of biological mechanisms that
26      underlie induction of such effects, this area of scientific inquiry is still at an early stage.  Still,
27      compared to the lack  of much solid evidence available in the 1996 PM AQCD, there now is a
28      stronger basis for assessing biologic plausibility of the epidemiologic observations given notable
29      improvement in conceptual formulation of reasonable mechanistic hypotheses and  evidence
30      bearing on such hypotheses. New evidence related to several hypotheses was discussed earlier
31      with regard to possible mechanisms by which ambient PM may exert human health effects,

        April 2002                                 9-86        DRAFT-DO NOT QUOTE OR CITE

-------
 1      which tends to support the likelihood of a causal relationship between low ambient
 2      concentrations of PM and observed increased mortality or morbidity risks. At the same time,
 3      much still remains to be done to identify more confidently specific causal agents among typical
 4      ambient PM constituents.
 5
 6      9.12.2 Community-Health Epidemic logic Evidence for Ambient Particulate
 7              Matter Effects
 8           In recent years, epidemiologic studies showing associations of ambient air pollution
 9      exposure with mortality,  exacerbation of preexisting illness, and pathophysiologic changes have
10      increased concern about the extent to which exposure to ambient air pollution exacerbates or
11      causes harmful health outcomes at pollutant concentrations now experienced in the United
12      States.  The PM epidemiology studies assessed in the 1996 PM AQCD implicated ambient PM
13      as a likely key contributor to mortality and morbidity effects observed epidemiologically to be
14      associated with ambient air pollution exposures. New studies appearing since the 1996 PM
15      AQCD are important in extending results of earlier studies to many more cities and in confirming
16      earlier findings.
17           In epidemiologic studies of ambient air pollution, small positive estimates of air pollutant
18      health effects have been observed quite consistently, frequently being statistically significant at
19      p < 0.05.  If ambient air pollution promotes or produces harmful health effects, relatively small
20      effect estimates from current PM concentrations in the United States and many other countries
21      would generally be expected on biological and epidemiologic grounds.  Also, magnitudes and
22      significance levels of observed air pollution-related effects estimates would be expected to vary
23      somewhat from place to place, if the observed epidemiologic associations denote actual effects,
24      because (a) not only would the complex mixture of PM vary from place to place, but also
25      (b) affected populations may differ in characteristics that could affect susceptibility to air
26      pollution health effects.  Such characteristics include sociodemographic factors, underlying
27      health status, indoor-outdoor activities, diet, medical care access, exposure to risk factors other
28      than ambient air pollution (such as extreme weather conditions), and variations in factors (e.g.,
29      air-conditioning) affecting human exposures to ambient-generated PM.
30           Although it has been argued by some that the observed effects estimates for ambient air
31      pollution are not sufficiently constant across epidemiologic studies and that epidemiologic

        April 2002                                 9-87        DRAFT-DO NOT QUOTE OR CITE

-------
 1      studies are trustworthy only if they show relatively large effects estimates (e.g., large relative
 2      risks), these arguments have only limited weight in relation to ambient air pollution studies.
 3      Also, in any large population exposed to ambient air pollution, even a small relative risk for a
 4      widely prevalent health disorder could result in a substantial public health burden attributable to
 5      air pollution exposure, as was noted earlier (see Section 9.10).
 6           As noted above, small health effects estimates generally have been observed for ambient air
 7      pollutants, as would be expected on biological and epidemiologic grounds. In contrast to effects
 8      estimates derived for the 1952 London smog episode with relative risk (RR) exceeding 4.0 (i.e.,
 9      400% increase over baseline) for extremely high (>2 mg/m3) ambient PM concentrations, effects
10      estimates in most current epidemiology studies at distinctly lower PM concentrations (often
11      < 100 //g/m3) are relatively small.  The statistical estimates (1) are more often subject to small
12      (but proportionately large) differences in estimated effects of PM and other pollutants; (2) may
13      be sensitive to a variety of methodological choices; and (3) sometimes may not be statistically
14      significant, reflecting low statistical power of the study design to detect a small but real effect.
15           The ambient atmosphere contains numerous air pollutants, and it is important to continue to
16      recognize that health effects associated statistically with any single pollutant may actually be
17      mediated by multiple components of the complex ambient mix.  Specific attribution of effects to
18      any single pollutant may therefore be overly simplistic. Particulate matter is one of many air
19      pollutants derived from combustion sources, including mobile sources.  These pollutants include
20      PM, carbon monoxide, sulfur oxides, nitrogen oxides, and ozone, all of which have been
21      considered  in various epidemiologic studies to date. Many volatile organic compounds (VOCs)
22      or semivolatile compounds (SVOCs) also emitted by combustion sources or formed in the
23      atmosphere have not yet been systematically considered in relation to noncancer health outcomes
24      usually associated with exposure to criteria air pollutants.  In many newly available
25      epidemiologic studies, harmful health outcomes are often associated with multiple combustion-
26      related or mobile-source-related air pollutants, and some investigators have raised the possibility
27      that PM may be a key surrogate or marker for a larger subset of the overall ambient air pollution
28      mix.  This possibility takes on added potential significance to the extent that ambient aerosols
29      indeed may not only exert health effects directly attributable to their constituent components, per
30      se, but also serve as carriers for more efficient delivery of water soluble toxic gases (e.g., O3,
31      NO2, SO2) deeper into lung tissue, as noted earlier in Section 9.8.4.  This suggests that airborne

        April 2002                                 9-88         DRAFT-DO NOT QUOTE OR CITE

-------
 1      particle effects may be enhanced by the presence of other toxic agents or mistakenly attributed to
 2      them if their respective concentrations are highly correlated temporally. Thus, although
 3      associations of PM with harmful effects continue to be observed consistently across most new
 4      studies, the newer findings do not fully resolve issues concerning relative contributions to the
 5      observed epidemiologic associations of (a) PM acting alone, (b) PM acting in combination with
 6      gaseous co-pollutants, (c) the gaseous pollutants per se, and (d) the overall  ambient pollutant
 7      mix.
 8           It seems likely that, for pollutants whose concentrations are not highly correlated, effects
 9      estimates in multipollutant models would be more biologically and epidemiologically sound than
10      those in single-pollutant models, although it is conceivable that single-pollutant models also
11      might be credible if independent biological plausibility evidence supported designation of PM or
12      some other single pollutant as likely being the key toxicant in the ambient pollutant mix
13      evaluated. Because neither of these possibilities have been definitively demonstrated  and there is
14      not yet full scientific consensus as to optimal  interpretation of modeling outcomes for time
15      series-air pollution studies, the choice of appropriate  effects estimates to employ in risk
16      assessments for ambient PM effects remains a difficult issue. Issues related to confounding by
17      co-pollutants, along with issues related to time scales of exposure and response and
18      concentration-response function, still apply to new epidemiologic studies relating concentrations
19      of PM or correlated ambient air pollutants to hospital admissions, exacerbation of respiratory
20      symptoms, and asthma in children, to reduced pulmonary function in children and adults, and to
21      changes in heart rate, and heart rate variability in adults.  However, with considerable  new
22      experimental evidence now in hand, it is possible to hypothesize various ways in which ambient
23      exposure to PM acting alone or in combination with others could plausibly be involved in the
24      complex chain of biological events leading to harmful health effects in the human population.
25      This newer experimental evidence, coupled with new exposure analyses results, add considerable
26      support for interpreting the epidemiologic findings discussed below as likely being indicative of
27      causal relationships between exposures to ambient PM and consequent associated increased
28      morbidity and mortality risks.
29
30
31

        April 2002                                 9-89        DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.12.2.1 Short-Term Particulate Matter Exposure Effects on Mortality
 2           This section focuses primarily on discussion of short-term PM exposure effects on
 3      mortality, but also highlights some morbidity effects in relation to the mortality findings.
 4      Morbidity effects are discussed more fully after discussion of long-term mortality effects in the
 5      section following this one.
 6
 7      Summary of Previous Findings on Short-Term Particulate Matter Exposure-Mortality Effects
 8           Time series mortality studies reviewed in the 1996 PM AQCD provided strong evidence
 9      that ambient PM air pollution is associated with increased daily mortality. The 1996 PM AQCD
10      summarized about 35 PM-mortality time series studies published between 1988 and 1996.  The
11      available information from those studies was consistent with the hypothesis that PM is a causal
12      agent in the mortality impacts of air pollution.  The PM10 relative risk estimates derived from the
13      PM10 studies reviewed in the 1996 PM AQCD suggested that an increase of 50 //g/m3 in the 24-h
14      average of PM10 is associated with an increased risk of premature total mortality (total deaths
15      minus accidents and injuries) mainly on of the order of relative risk (RR) = 1.025 to 1.05 (i.e.,
16      2.5 to 5.0% excess risk) in the general population, with statistically significant increases being
17      reported more broadly across the range of 1.5 to 8.5%  per 50 //g/m3 PM10. Higher relative risks
18      were indicated for the elderly and for those with preexisting respiratory conditions. Also, based
19      on the then recently published Schwartz et al. (1996) analysis of Harvard Six City data, the 1996
20      PM AQCD found the relative risk for excess total mortality in relation to 24-h fine-particle
21      concentrations to be in the range of RR = 1.026 to 1.055 per 25 Mg/m3 PM25 (i.e., 2.6 to 5.5%
22      excess risk per 25 //g/m3 PM2 5). Relative risk estimates for morbidity and mortality effects
23      associated with standard increments in ambient PM10 concentrations and for fine-particle
24      indicators  (e.g., PM25, sulfates,  etc.) were presented in Chapters 12 and 13 of the 1996 PM
25      AQCD (see Appendix 9 A), and those effect estimates  are updated below in light of the extensive
26      newly available evidence discussed in Chapter 8 of this document.
27           Although numerous studies reported PM-mortality associations, several important issues
28      needed to be addressed in interpreting those relative risks.  The 1996 PM AQCD extensively
29      discussed the following critical  issues:  (1) seasonal confounding and  effect modification,
30      (2) confounding by weather, (3) confounding by co-pollutants, (4) measurement error,


        April 2002                                9-90        DRAFT-DO NOT QUOTE OR CITE

-------
 1      (5) functional form and threshold, (6) harvesting and life shortening; and (7) the roles of specific
 2      PM components.
 3           Season-specific analyses are often not feasible because of small magnitudes of expected
 4      effect size or small sample sizes (low power) available for some studies.  Some studies had
 5      earlier suggested possible season-specific variations in PM coefficients, but it was not clear if
 6      these were caused by peak variations in PM effects from season to season, varying extent of PM
 7      correlations with other co-pollutants, or weather factors during different seasons.  The likelihood
 8      of PM effects being accounted for mainly by weather factors was addressed by various methods
 9      that controlled for weather variables in most studies (including some involving  sophisticated
10      synoptic weather pattern evaluations), and that possibility was found to be very unlikely.
11           Many early PM studies considered at least one co-pollutant in the mortality regression, and
12      an increasing number have examined multiple pollutants. Usually, when PM indices were
13      significant in single-pollutant models, addition of a co-pollutant diminished the PM effect size
14      somewhat, but did not eliminate PM associations. In multiple-pollutant models performed by
15      season, the PM coefficients became less stable,  again possibly because of varying correlations of
16      PM with co-pollutants among seasonal or smaller sample sizes. However, in many studies, PM
17      indices showed the highest significance in both single- and multiple-pollutant models. Thus,
18      PM-mortality associations did not appear to be seriously distorted by co-pollutants.
19           Interpretation of the relative significance of each pollutant in mortality regression in
20      relation to its relative causal strength was difficult, however, because of lack of quantitative
21      information on pertinent exposure measurement errors among the air pollutants. Measurement
22      errors can influence the size and significance of air pollution coefficients in time series
23      regression analyses, an issue also important in assessing confounding among multiple pollutants,
24      because the varying extent of such errors among pollutants may influence corresponding relative
25      significance. The 1996 PM AQCD discussed several types of exposure measurement and
26      characterization errors, including site-to-site variability and site-to-person variability.  These
27      errors are thought to bias the estimated PM coefficients downward in most cases, but there was
28      insufficient quantitative information available at the time to allow estimation of such bias.
29           The 1996 PM AQCD also reviewed evidence for threshold and various other functional
30      forms of short-term PM mortality associations.  Some studies indicated that associations were
31      seen monotonically to even below the PM standards. It was considered difficult, however, to

        April 2002                                9-91        DRAFT-DO NOT QUOTE OR CITE

-------
 1      statistically identify a threshold from available data because of low data density at lower ambient
 2      PM concentrations, potential influence of measurement error, and adjustments for other
 3      covariates. Thus, use of relative risk (rate ratio) derived from log-linear Poisson models was
 4      deemed adequate.
 5           The extent of prematurity of death, i.e., mortality displacement (or harvesting) in observed
 6      PM-mortality associations has important public health policy implications. At the time of the
 7      1996 PM AQCD review, only a few studies had investigated this issue. Although one of the
 8      studies suggested that the extent of such prematurity might be only a few days, this may not be
 9      generalized because this estimate was obtained for identifiable PM episodes. Insufficient
10      evidence then existed to suggest the extent of prematurity for nonepisodic periods, from which
11      most of the recent PM relative risks were derived.
12           Only a few PM-mortality studies had analyzed fine particles and chemically specific
13      components of PM. The Harvard Six Cities Study (Schwartz et al.,  1996) analyzed size-
14      fractionated PM (PM25, PM10/15, and PM10/15.2 5) and PM chemical components (sulfates and H+).
15      The results suggested that PM2 5 was  associated most significantly with mortality among the PM
16      components. Although H+ was not significantly associated with mortality in this and earlier
17      analyses, the smaller sample size for  H+ than for other PM components made direct comparison
18      difficult. Also, certain respiratory morbidity studies showed associations between hospital
19      admissions and visits with components of PM in the fine-particle range.  Thus, the 1996 PM
20      AQCD concluded that there was adequate evidence to suggest that fine particles play especially
21      important roles in observed PM mortality effects.
22           Overall, then, the outcome of assessment of the above key issues in the 1996 PM AQCD
23      can be thusly summarized:  (1) observed PM effects are not likely seriously biased by inadequate
24      statistical modeling (e.g., control for  seasonality); (2) observed PM effects are not likely
25      significantly confounded by weather; (3) observed PM effects may be  confounded or modified to
26      some extent by co-pollutants, and such extent may vary from season to season; (4) determining
27      the extent of confounding and effect modification by co-pollutants requires knowledge of relative
28      exposure measurement/characterization error among pollutants (there was not sufficient
29      information on this); (5) no clear evidence for any threshold for PM-mortality associations was
30      reported (statistically identifying a threshold from existing data also was considered difficult, if
31      not impossible); (6) some limited evidence for harvesting, a few days of life-shortening, was

        April 2002                                9-92        DRAFT-DO NOT QUOTE OR CITE

-------
 1      reported for episodic periods (no study was conducted to investigate harvesting in nonepisodic
 2      U.S. data); and (7) only a relatively limited number of studies suggested a causal role of fine
 3      particles in PM-mortality associations, but in light of historical data, biological plausibility, and
 4      results from morbidity studies, a greater role for fine particles than coarse particles was suggested
 5      as being likely.
 6
 7      Updated Epidemiologic Findings for Short-Term Ambient Particulate Matter
 8      Exposure Effects on Mortality
 9           With regard to updating the assessment of PM effects in light of new epidemiologic
10      information published since the 1996 PM AQCD, the most salient key points on relationships
11      between short-term PM exposure and mortality (drawn from Chapter 8 discussions in this
12      document) can be summarized as follows.
13           Since the 1996 PM AQCD, there have been more than 80 new time-series PM-mortality
14      analyses, several of which investigated multiple cities using consistent data analytical
15      approaches.  With only few exceptions, the estimated mortality relative risks in these studies are
16      generally positive, many are statistically significant, and they generally comport well with
17      previously reported PM-mortality effects estimates delineated in the 1996 PM AQCD.  There are
18      also now numerous additional studies demonstrating associations between short-term (24-h) PM
19      exposures and various morbidity endpoints.
20           Several new studies conducted time series analyses in multiple cities. The major advantage
21      of these studies over meta-analyses for multiple "independent" studies is the consistency in data
22      handling and model specifications, thus eliminating variation in results attributable to study
23      design. Also, many of the cities included in these studies were ones for which no earlier time
24      series analyses had been conducted.  Therefore, unlike regular meta-analysis, they likely do not
25      suffer from omission of negative studies caused by publication bias. Furthermore, any spatial or
26      geographic variability  of air pollution effects can be systematically evaluated in such multi-city
27      analyses.
28
29           PM10 Effect Size Estimates. In the NMMAPS (Samet et al., 2000a,b) analysis of the
30      90 largest U.S. cities, the combined nationwide relative risk estimate was about a 2.3% increase
31      in total mortality per 50-//g/m3 increase in PM10.  The  NMMAPS effect size estimates did vary

        April 2002                                 9-93         DRAFT-DO NOT QUOTE OR CITE

-------
 1      somewhat by U.S. region (see Figures 8-3 and 8-5), with the largest estimate being for the
 2      Northeast (4.5% for a 1-day lag, the lag typically showing maximum effect size for most U.S.
 3      regions).  Various other U.S. multi-city analyses, as well as single-city analyses, obtained PM10
 4      effect sizes mainly in the range of 2.5 to 5.0% per 50-//g/m3 increase in PM10.  There is some
 5      evidence  that, if the effects over multiple days are considered, the effect size may be larger.
 6      What heterogeneity  existed for the estimated PM10 risks across NMMAPS cities could not be
 7      explained with the city-specific explanatory variables (e.g., as the mean levels of pollution and
 8      weather), mortality rate,  sociodemographic variables (e.g.,  median household income),
 9      urbanization, or variables related to measurement error.
10           Original results reported for the multi-city APHEA study showed generally consistent
11      associations between mortality and both SO2 and PM indices in western European cities, but not
12      for central and eastern European cities. More recent studies from APHEA n analyses, however,
13      found analogous increased risks to be associated with PM exposures in central and eastern
14      Europe as in western European cities.  The pooled estimate of PM10-mortality relative risks for
15      European cities comport well with estimates derived from U.S. data.
16           Certain other individual-city studies using similar methodology in analyses for each city
17      (but not generating combined overall pooled effect estimates) also report variations in PM effect
18      size estimates between cities and in their robustness to inclusion of gaseous copollutants in
19      multi-pollutant models.  Thus, one cannot entirely rule out that real differences may exist in
20      excess risk levels associated with varying size distributions, number, or mass of the chemical
21      constituents of ambient PM; the combined influences of varying co-pollutants present in the
22      ambient air pollution mix from location to location or season to season; or to variations in the
23      relationship between exposure and ambient PM concentration.
24           Nevertheless, there still appears to be reasonably good consistency among the results
25      derived from those several new multi-city studies providing pooled analyses of data combined
26      across multiple cities (thought to yield  the most precise effect size estimates).  Such analyses
27      indicate the percent  excess total (nonaccidental) deaths estimated per 50 //g/m3 increase in 24-h
28      PM10 to be 2.3% in the 90 largest U.S.  cities (4.5% in the Northeast region); 3.4% in 10 U.S.
29      cities; 3.5% in the eight largest Canadian cities; and about 2.0% in European cities (using PM10
30      = TSP*0.55). These combined estimates are reasonably consistent with the range of PM10
31      estimates previously reported in the 1996 PM AQCD (i.e.,  1.5 to 8.5% per 50 //g/m3 PM10).

        April 2002                                9-94        DRAFT-DO NOT QUOTE OR CITE

-------
 1      These and other excess risk estimates from many other individual-city studies comport well with
 2      a number of new studies confirming increased cause-specific cardiovascular- and respiratory-
 3      related mortality, and those noted below as showing ambient PM associations with increased
 4      cardiovascular and respiratory hospital admissions and medical visits.
 5
 6           Fine and Coarse Particle Effect Size Estimates.  Table 9-14 summarizes effects
 7      estimates (RR values) for increased mortality and/or morbidity associated with variable
 8      increments in short-term (24-h) exposures to ambient fine particles indexed by various fine PM
 9      indicators (PM25, sulfates, H+, etc.) in U.S. and Canadian cities.  Table 9-15 shows analogous
10      effect size estimates for inhalable thoracic fraction coarse particles (i.e., PM10_25). In both tables,
11      studies that were highlighted in comparable tables in the 1996 PM AQCD are indicated by
12      italics. For purposes of comparison across studies, results of single-pollutant models are
13      presented in these tables; co-pollutant model results are presented and discussed in more detail in
14      Chapter 8.
15           The effect size estimates derived for PM25  as an ambient fine particle indicator (especially
16      those based on directly measured versus estimated PM2 5 levels) generally appear to fall in the
17      range of 2.0 to 8.5% increase in total (nonaccidental) deaths per 25-//g/m3 increment in 24-h
18      PM25 for U.S. and Canadian cities. Cause-specific effects estimates appear to fall mainly in the
19      range of 3.0 to 7.0% per 25 //g/m3 24-h PM25 for cardiovascular or combined cardiorespiratory
20      mortality and 2.0 to 7.0% per 25 //g/m3 24-h PM2 5 for respiratory mortality in U.S. cities.
21           In the 1996 PM AQCD, there was only one study, the Harvard Six Cities study, in which
22      the relative importance of fine and coarse particles was examined.  That study suggested that fine
23      particles, but not coarse particles, were associated with daily mortality.  Now, more than
24      10 studies have analyzed both PM2 5 and PM10_25 for their associations with mortality (see
25      Figure 9-25).  Although some of these studies (e.g., the Santa Clara County, CA, analysis and the
26      eight largest Canadian cities analysis) suggest that PM2 5 is more important than PM10_2 5 in
27      predicting mortality fluctuations, several others (e.g., the Mexico City and Santiago, Chile
28      studies) seem to suggest that PM10_2 5 may be as important as PM2 5 in  certain locations (some
29      shown to date being drier, more arid areas). Seasonal  dependence of PM components'
30      associations observed in some of the locations (e.g., higher coarse [PM10_25] fraction estimates for
31      summer than winter in Santiago, Chile) hint at possible contributions  of biogenic materials (e.g.,

        April 2002                                 9-95        DRAFT-DO NOT QUOTE  OR CITE

-------
  TABLE 9-14. EFFECT ESTIMATES PER VARIABLE INCREMENTS IN 24-HOUR
     CONCENTRATIONS OF FINE PARTICLE INDICATORS (PM2 5, SO^, H+)
                  FROM U.S. AND CANADIAN STUDIES*
Study Location
Indicator
RR(±CI)**per25-,wg/m3
PM Increase or 15-^g/m3
SOJ Increase or 75-nmol/m3 H+ increase
Reported PM
Levels Mean
(Min, Max)***
Acute Total Mortality
Six City:A
Portage, WI
Topeka, KS
Boston, MA
St. Louis, MO
Kingston/Knoxville, TN
Steubenville, OH
Overall Six-City Results
Six U.S. Cities8
Santa Clara County, CAC
Buffalo, NYD
Philadelphia, PAE
Detroit, MIF
Phoenix, AZG
Phoenix, AZH
Los Angeles, CA1
San Bernadino and
Riverside Counties, CAJ
Coachella Valley, CAK
Boston, MAL
Three New Jersey Cities :M
Newark, NJ
Camden, NJ
Elizabeth, NJ
Eight Canadian CitiesN

PM25
PM25
PM25
PM2.5
PM25
PM25
PM25
PM25
PM25
so;
PM25
PM25
PM25
PM25
PM25
Est. PM2 5
PM25
PM25

PM25
PM25
PM25
PM25

1.030(0.993, 1.071)
1.020(0.951, 1.092)
1.056(1.038, 1.0711)
1.028(1.010, 1.043)
1.035 (1.005, 1.066)
1.025(0.998, 1.053)
1.015 (1.011, 1.019)
Overall 1.010 (1.028, 1.053)
Mobile 1.087(1.042, 1.134)
Coal 1.028 (1.006, 1.050)
Crustal 0.944 (0.863, 1.032)
1.13 (p< 0.01)
1.034 (1.009, 1.062)
1.042 (p< 0.055)
1.031(0.994, 1.069)
1.030 (1.000, 1.076)
(>25 Mg/m3) 2.868 (1.126, 7.250)
(<25 ^g/m3) 0.779 (0.610, 0.995)
1. 06 (NS, from figure)
1.003 (0.992, 1.015)
1.118(1.013, 1.233)
1.053 (1.018, 1.090)

1.043 (1.028, 1.059)
1.057(1.001, 1.115)
1.018 (0.946, 1.095)
1.030(1.011, 1.050)

11.2 (±7.8)
12. 2 (±7.4)
15.7 (±9.2)
18.7 (±10.5)
20.8 (±9.6)
29.6 (±2 1.9)
Median 14.7
Means 11.3-30.5
13 (2, 105)
61.7(0.78,390.5)
nmol/m3
17.28 (-0.6, 72.6)
18 (6, 86)
13.0 (0, 42)
NR
22 (4, 86)
32.5(9.3, 190.1)
16.8 (5, 48)
15.6 (±9.2)

42.1 (±22.0)
39.9 (±18.0)
37.1 (±19.8)
13. 3 (max 86)
April 2002
9-96
DRAFT-DO NOT QUOTE OR CITE

-------
   TABLE 9-14 (cont'd). EFFECT ESTIMATES PER VARIABLE INCREMENTS IN
 24-HOUR CONCENTRATIONS OF FINE PARTICLE INDICATORS (PM2 5, SO^, H+)
                   FROM U.S. AND CANADIAN STUDIES*
Study Location
Toronto, Canada0
Montreal, Canada1"
Indicator
Est. PM2.5
PM25
RR(±CI)**per25-,wg/m3
PM Increase or 15-^g/m3
SO; Increase or 75-nmol/m3 H+ increase
1.048 (1.033, 1.064)
1.058 (1.034, 1.083)
Reported PM
Levels Mean
(Min, Max)***
18.0 (8, 90)
17.4 (2.2, 72.0)
Cause-Specific Mortality
Cardiorespiratorv:
Three New Jersey Cities :M
Newark, NJ
Camden, NJ
Elizabeth, NJ
Total Cardiovascular:
Santa Clara County, CAC
Buffalo, NY)D
Philadelphia, PAF
(seven-county area)
Detroit, MIG
Phoenix, AZH
Los Angeles, CA1
San Bernadino and
Riverside Counties, CA1
Coachella Valley, CAK
Cerebrovascular:
Los Angeles, CA1
Total Respiratory:
Santa Clara County, CAC
Buffalo, NYD
Philadelphia, PAF
(seven-county area)
Detroit, MIG
San Bernadino and
Riverside Counties, CA1


PM25
PM25
PM25

PM25
so;
PM25
PM25
PM25
PM25
Est. PM2 5
PM25

PM25

PM25
so;
PM25
PM25
Est. PM2 5


1.051(1.031, 1.072)
1.062(1.006, 1.121)
1.023 (0.950, 1.101)

1.07 (p> 0.05)
1.040 (0.995, 1.088)
1.028 (p< 0.055)
1.032 (0.977, 1.089)
1.187(1.057, 1.332)
1.027 (1.003, 1.048)
1.007 (0.997, 1.017)
1.086 (0.937, 1.258)

1.036 (0.994, 1.080)

1.13(p>0.05)
1.108(1.007, 1.219)
1.014 (p> 0.055)
1.023 (0.897, 1.166)
1.021 (0.997, 1.045)


42.1 (±22.0)
39.9 (±18.0)
37.1 (±19.8)

13 (2, 105)
61.7(0.78,390.5)
nmol/m3
17.28 (-0.6, 72.6)
18 (6, 86)
13.0 (0, 42)
22 (4, 86)
32.5(9.3, 190.1)
16.8 (5, 48)

22 (4, 86)

13 (2, 105)
61.7(0.78,390.5)
nmol/m3
17.28 (-0.6, 72.6)
18 (6, 86)
32.5(9.3, 190.1)
April 2002
9-97
DRAFT-DO NOT QUOTE OR CITE

-------
   TABLE 9-14 (cont'd). EFFECT ESTIMATES PER VARIABLE INCREMENTS IN
 24-HOUR CONCENTRATIONS OF FINE PARTICLE INDICATORS (PM2 5, SO^, H+)
                   FROM U.S. AND CANADIAN STUDIES*
Study Location
COPD:
Los Angeles, CA1
Increased Hospitalization
Ontario, Canada®
Ontario, Canada1^
NYC/Buffalo, NYS
Toronto, Canada3
Total Respiratory:
King County, WAT
Toronto, Canada"
Buffalo, NYD
Montreal, Canadav
Montreal, Canadaw
St. John, Canadax
Pneumonia:
Detroit, MIF
Respiratory infections:
Toronto, Canada"
COPD:
Atlanta, GAZ
Detroit, MIF
King County WA^
Los Angeles, CABB
Toronto, CanadaY
Indicator

PM25

so=4
so:
03
so=4
H+ (Nmol/m3)
SO^
PM25

PMj
PM25
SOJ
PM25
PM25
PM25

PM25

PM25

PM25
PM25
PM25
PM25
PM25
RR(±CI)**per25-,wg/m3
PM Increase or 15-^g/m3
SOJ Increase or 75-nmol/m3 H+ increase

1.027 (0.966, 1.091)

1.03(1.02, 1.04)
1.03(1.02, 1.04)
1.03(1.02, 1.05)
1.05 (1.01, 1.10)
1.16(1.03, 1.30)'
1.12(1.00, 1.24)
1.15 (1.02, 1.28)

1.058(1.011, 1.110)
1.085 (1.034, 1.138)
1.082(1.042, 1.128)
1.239 (1.048, 1.428)
1.137(0.998, 1.266)
1.057(1.006, 1.110)

1.125 (1.037, 1.220)

1.108(1.072, 1.145)

1.124(0.921, 1.372)
1.055 (0.953, 1.168)
1.064(1.009, 1.121)
1.051(1.009, 1.094) (65+ y.o.)
1.04(0.99, 1.09) ((0-19 y.o.)
1.06(1.02, 1.09) (20-64 y.o.)
1.048(0.998, 1.100)
Reported PM
Levels Mean
(Min, Max)***

22 (4, 86)

R = 3. 1-8.2
R = 2. 0-7.7
NR
28.8 (NR/391)
7. 6 (NR, 48.7)
18.6 (NR, 66.0)

NR
16.8(1,66)
61.7(0.78,390.5)
nmol/m3
Summer 93
12.2 (max 31)
18.6 (SD 9.3)
Summer 93
8.5 (max 53.2)

18 (6, 86)

18.0 (max 90)

19.4 (±9.35)
18 (6, 86)
18.1 (3,96)
Median 22 (4, 86)
18.0 (max 90)
April 2002
9-98
DRAFT-DO NOT QUOTE OR CITE

-------
   TABLE 9-14 (cont'd). EFFECT ESTIMATES PER VARIABLE INCREMENTS IN
 24-HOUR CONCENTRATIONS OF FINE PARTICLE INDICATORS (PM2 5, SO^, H+)
                   FROM U.S. AND CANADIAN STUDIES*
Study Location
Asthma:
Atlanta, GAZ
Seattle, WACC
Seattle, WADD
Toronto, CanadaY
Total Cardiovascular:
Atlanta, GAZ
Buffalo, NYD
Los Angeles, CAEE
St. John, Canadax
Toronto, Canada"
Ischemic Heart Disease:
Detroit, MIF
Toronto, CanadaY
Dvsrhvthmias:
Atlanta, GAZ
Detroit, MIF
Toronto, CanadaY
Heart Failure:
Detroit, MIF
Toronto, CanadaY
Cerebrovascular:
Los Angeles, CAEE
Toronto, CanadaY
Peripheral circulation diseases:
Toronto, CanadaY
Indicator

PM25
PM25
Est. PM2 5
PM25

PM25
SOI
PM25
PM25
PM25

PM25
PM25

PM25
PM25
PM25

PM25
PM25

PM25
PM25

PM25
RR(±CI)**per25-,wg/m3
PM Increase or 15-^g/m3
SOJ Increase or 75-nmol/m3 H+ increase

1.023 (0.852, 1.227)
1.087(1.033, 1.143)
1.445 (1.217, 1.714)
1.064(1.025, 1.106)

1.061 (0.969, 1.162)
1.015 (0.987, 1.043)
(65+) 1.043 (1.025, 1.061)
(<65) 1.035 (1.018, 1.053)
1.151 (0.998, 1.328)
1.072(0.994, 1.156)

1.043 (0.986, 1.104)
1.080(1.054, 1.108)

1.061 (0.874, 1.289)
1.032(0.934, 1.140)
1.061(1.019, 1.104)

1.091 (1.023, 1.162)
1.066(1.025, 1.108)

1.015 (0.992, 1.038)
"NEG" reported

"NEG" reported
Reported PM
Levels Mean
(Min, Max)***

19.4 (±9.35)
16.7 (6, 32)
4.8 (1.2, 32.4)
18.0 (max 90)

19.4 (±9.35)
61.7(0.78,390.5)
nmol/m3
Median 22 (4, 86)
Summer 93
8.5 (max 53.2)
16.8(1,66)

18 (6, 86)
18.0 (max 90)

19.4 (±9.35)
18 (6, 86)
18.0 (max 90)

18 (6, 86)
18.0 (max 90)

Median 22 (4, 86)
18.0 (max 90)

18.0 (max 90)
April 2002
9-99
DRAFT-DO NOT QUOTE OR CITE

-------
   TABLE 9-14 (cont'd). EFFECT ESTIMATES PER VARIABLE INCREMENTS IN
 24-HOUR CONCENTRATIONS OF FINE PARTICLE INDICATORS (PM2 5, SO^, H+)
                   FROM U.S. AND CANADIAN STUDIES*
Study Location
Stroke:
Detroit, MIF
Increased Respiratory Symptoms
Southern California1'1'
Six Cities00
(Cough)
Six Cities00
(Lower Resp. Symp.)
Uniontown, PAm
(Evening Cough)
Six Cities ReanalysesKK
(Lower Resp. Symptoms)
(Cough)
Connecticut summer camp11
State College, PAJJ
(Wheeze)
State College, PAJJ
(Cold)
State College, PAJJ
(Cough)
Decreased Lung Function
Uniontown, PAm
Uniontown, PA1^
(Reanalysis)
State College, PA835
(Reanalysis)
Connecticut summer camp11
Indicator

PM25

SO^
PM25
so^
H+
PM25
so:
H+
PM25
PM25
so;
PM21
PM21
PM21

PM25
PM25
PM25
so;
RR(±CI)**per25-Mg/m3
PM Increase or 15-^g/m3
SO; Increase or 75-nmol/m3 H+ increase

1.018 (0.947, 1.095)
Odd Ratio (95% CI) per 25-^g/m3
PM Increase or 15-^g/m3
SO; Increase or 75-nmol/m3 H+ increase
1.48(1.14, 1.91)
1.24(1.00, 1.54)
1.86(0.86, 4.03)
1.19(0.66,2.15)
1.58(1.18, 2.10)
6.82(2.09, 17.35)
1.16(0.10, 13.73)
1.45 (1.07, 1.97)
1.61 (1.20,2.16)
1.28(0.98, 1.67)
1.71(1.30,2.25)
1.59(0.94,2.71)
1.61 (1.21,2.17)
1.48(1.17, 1.88)
PEFR change (L/min) per 25-^g/m3
PM Increase or 15-^g/m3
SO; Increase or 75-nmol/m3 H+ increase
PEFR -1.38 (-2.77, 0.02)
pmPEFR -1.52, (-2.80, -0.24)
pm PEFR -0.93 (-1.88, 0.01)
PEFR -5.4 (-12.3, 1.52)
Reported PM
Levels Mean
(Min, Max)***

18 (6, 86)

R = 2-37
18.0 (max 86.0)
2.5 (max 15.1)
18.1 (max 37 1.1)
nmol/m3
18.0 (max 86.0)
2.5 (max 15.1)
18.1 (max 37 1.1)
nmol/m3
24.5 (max 88.1)
18.0 (max 86.0)
2.5 (max 15.1)
18.1 (max 37 1.1)
nmol/m3
7.0 (1.1,26.7)
23. 5 (max 85. 8)
23. 5 (max 85. 8)
23. 5 (max 85. 8)

24.5 (max 88.1)
24.5 (max 88.1)
23. 5 (max 85. 8)
7.0(1.1,26.7)
April 2002
9-100
DRAFT-DO NOT QUOTE OR CITE

-------
    TABLE 9-14 (cont'd). EFFECT ESTIMATES PER VARIABLE INCREMENTS IN
  24-HOUR CONCENTRATIONS OF FINE PARTICLE INDICATORS (PM2 5, SO^, H+)
                          FROM U.S. AND CANADIAN STUDIES*
Study Location
Southwest, VALL
State College, PAJJ
Philadelphia, PAMM
Indicator
PM25
PM21
PM25
RR(±CI)**per25-Mg/m3
PM Increase or 15-^g/m3
SOJ Increase or 75-nmol/m3 H+ increase
amPEFR -1.825 (-3.45, -0.21)
pmPEFR -0.63 (-1.73, 0.44)
amPEFR -3.28 (-6.64, 0.07)
pmPEFR -0.91 (-4.04, 2.21)
Reported PM
Levels Mean
(Min, Max)***
21.62 (3.48, 59.65)
23. 5 (max 85. 8)
22.2 (IQR 16.2)
* Studies highlighted in the 1996 CD are in italics; new studies in plain text.  For purposes of comparison across
 studies, results of single-pollutant models are presented in these tables; co-pollutant model results are presented and
 discussed in more detail in Chapter 8.
**Relative Risk (95% Confidence Interval), except for Fairley (1999) and Lipfert et al. (2000), where insufficient
data were available to calculate confidence intervals so p-value is given in parentheses.
***Min, Max 24-h PM indicator level shown in parentheses unless otherwise noted as (±S.D.), NR = not reported,
 or R = range of values from min-max, no mean value reported.
References:

 ASchwartz et al. (1996)
 BLaden et al. (2000)
 GFairley (1999)
 DGwynn et al. (2000)
 ELipfert et al. (2000a)
 FLippmann et al. (2000)
 GMaretal. (2000)
 HSmith et al. (2000)
 'Moolgavkar (2000a)
 JOstro (1995)
 KOstro et al. (2000)
 LSchwartz (2000)
 MTsai et al. (2000)
NBurnett et al. (2000)
°Burnett et al. (1998)
FGoldberg et al. (2000)
QBurnett et al. (1994)
RBurnett et al. (1995)
sThurston et al. (1992, 1994)
TLumley and Heagerty (1999)
"Burnett etal. (1997)
vDelfino et al. (1997)
wDelfino et al. (1998)
xStieb et al. (2000)
YBurnett et al. (1999)
zTolbert et al. (2000)
      ^Moolgavkar et al. (2000)
      BBMoolgavkar (2000b)
      CGSheppard et al. (1999)
      DDNorris et al. (1999)
      EEMoolgavkar (2000c)
      FFOstroetal. (1993)
      GGSchwartz et al. (1994)
      HHNeasetal. (1995)
      "Thurston et al. (1997)
      "Neasetal. (1996)
      ^Schwartz and Neas (2000)
      LLNaeher et al. (1999)
      MMNeasetal. (1999)
April 2002
            9-101
DRAFT-DO NOT QUOTE OR CITE

-------
  TABLE 9-15. EFFECT ESTIMATES PER VARIABLE INCREMENTS IN 24-HOUR
       CONCENTRATIONS OF COARSE-FRACTION PARTICLES (PM10.2.S)
                  FROM U.S. AND CANADIAN STUDIES*
Study Location
Indicator
RR (±CI)** per 25-,wg/m3
Increase
Reported PM
Levels Mean
(Min, Max)***
Acute Mortality
Six Cities:A
Portage, WI
Topeka, KS
Boston, MA
St. Louis, MO
Kingston/Knoxville, TN
Steubenville, OH
Overall Six-City Results
Coachella Valley, CAB
Detroit, MIC
Philadelphia, PAD
Phoenix, AZE
Phoenix, AZF
Santa Clara County, CAG
Eight Canadian Cities11

PM10_2,
PM10_25
PM10_25
PM10_2,
PM10_2,
PM10_25
PM10_25
PM10.2.5
PM10.2.5
PM10.2.5
PM10.,5
PM10.2.5
PM10.2.5
PM1(M,

1.013 (0.970, 1.058)
0.968(0.920, 1.015)
1.005 (0.985, 1.030)
1.005 (0.983, 1.028)
1.025 (0.985, 1.066)
1.061 (1.013, 1.111)
1.004(0.999, 1.010)
1.013 (0.994, 1.032)
1.040 (0.988, 1.094)
1.052 (p> 0.055)
1.030 (0.995, 1.066)
(>25 Mg/m3) 1.185 (1.069, 1.314)
(<25 Mg/m3) 1.020 (1.005, 1.035)
1.03 (p>0.05))
1.018 (0.992, 1.044)

6.6 (±6.8)
14.5 (±12.2)
8. 8 (±7.0)
11.9 (±8.5)
11.2 (±7.4)
16.1 (±13.0)
Median 9.0
17.9 (0, 149)
13 (4, 50)
6.80 (-20.0, 28.3)
33.5 (5, 187)
NR
11(0,45)
12.9 (max 99)
Cause-Specific Mortality
Total Cardiovascular:
Coachella Valley, CAB
Detroit, MIC
Philadelphia, PAD
(seven-county area)
Phoenix, AZE
Santa Clara County, CAG
Total Respiratory:
Coachella Valley, CAB
Detroit, MID
Philadelphia, PAD
(seven-county area)
Santa Clara County, CAG

PM10.2.5
PM10.2.5
PM10.2.5
PM10.2.5
PM10.2.5

PM10.2.5
PM10.2.5
PM10.,5
PM,^<

1.026 (1.006, 1.045)
1.078(1.000, 1.162)
1.034 (p> 0.055)
1.064(1.014, 1.117)
1.03 (p> 0.05)

1.026 (1.006, 1.045)
1.074 (0.910, 1.269)
1.030 (p> 0.055)
1.16(p>0.05)

17.9 (0, 149)
13 (4, 50)
6.80 (-20.0, 28.3)
33.5 (5, 187)
11(0,45)

17.9 (0, 149)
13 (4, 50)
6.80 (-20.0, 28.3)
11 (0,45)
April 2002
9-102
DRAFT-DO NOT QUOTE OR CITE

-------
   TABLE 9-15 (cont'd). EFFECT ESTIMATES PER VARIABLE INCREMENTS IN
   24-HOUR CONCENTRATIONS OF COARSE-FRACTION PARTICLES (PM10.2.S)
                   FROM U.S. AND CANADIAN STUDIES*
Study Location
Indicator
RR(±CI)**per25-,ag/m3
Increase
Reported PM
Levels Mean
(Min, Max)***
Increased Hospitalization
Total Respiratory:
Toronto, Canada1
Pneumonia:
Detroit, MIC
Respiratory infections:
Toronto, Canada1
COPD:
Atlanta, GAK
Detroit, MIC
Los Angeles'3
Toronto, Canada1
Total Cardiovascular:
Atlanta, GAK
Toronto, Canada1
Ischemic Heart Disease:
Detroit, MIC
Toronto, Canada1
Dysrhythmias:
Detroit, MIC
Atlanta, GAK
Toronto, Canada1
Heart Failure:
Detroit, MIC
Toronto, Canada1
Stroke:
Detroit, MIC
Cerebrovascular:
Toronto, Canada1

PM10.,5

PM10.2.5

PM10.,5

PM10.2.5
PM10.2.5
PM10.2.5
PM10.2.5

PM10.2.5
PM10.2.5

PM10.2.5
PM10.2.5

PM10.2.5
PM10.2.5
PM10.2.5

PM10.2.5
PM10.2.5

PM10.2.5

PM10.2.5

1.125 (1.052, 1.20)

1.119(1.006, 1.244)

1.093(1.046, 1.142)

0.770 (0.493, 1.202)
1.093 (0.958, 1.247)
1.17(1.09, 1.26) (0-19 y.o.)
1.09(1.03, 1.15) (20-64 y.o.)
1.05(0.99, 1.11) (65+ y.o.)
1.128(1.049, 1.213)

1.176(0.954, 1.450)
1.205(1.082, 1.341)

1.105(1.027, 1.189
1.037 (1.013, 1.062))

1.002(0.877, 1.144)
1.532(1.021,2.30)
1.051(0.998, 1.108)

1.052(0.967, 1.144)
1.079(1.023, 1.138)

1.049(0.953, 1.155)

"NEG" reported

11.6(1,56)

13 (4, 50)

12.2 (max 68)

9.39 (±4.52)
13 (4, 50)
NR
12.2 (max 68)

9.39 (±4.52)
11.6(1,56)

13 (4, 50)
12.2 (max 68)

13 (4, 50)
9.39 (±4.52)
12.2 (max 68)

13 (4, 50)
12.2 (max 68)

13 (4, 50)

12.2 (max 68)
April 2002
9-103
DRAFT-DO NOT QUOTE OR CITE

-------
    TABLE 9-15 (cont'd). EFFECT ESTIMATES PER VARIABLE INCREMENTS IN
    24-HOUR CONCENTRATIONS OF COARSE-FRACTION PARTICLES (PM10.2.S)
                          FROM U.S. AND CANADIAN STUDIES*
      Study Location
Indicator
RR(±CI)**per25-,ag/m3
        Increase
Reported PM
Levels Mean
 (Min, Max)***
 Peripheral Circulation Diseases:

  Toronto, Canada1              PM10_2.5
 Asthma:
                       1.056(1.003, 1.112)
                                 12.2 (max 68)
Seattle, WAL
Toronto, Canada1
Increased Respiratory Symptoms
Six U.S. CitiesM
(Lower Respiratory
Symptoms)
Six U.S. CitiesM
(Cough)
Southwest VirginiaN
(Runny or Stuffy Nose)
Decreased Lung Function
Southwest Virginia0
Uniontown, PAM
(Reanalysis)
State College, PAM
(Reanalysis)
Philadelphia, PAP
PM10.2.5
PM10.2.5

PM10.2.5
PM10.,5
PM10.2.5

PM10.2.5
PM10.2.5
PM10.2.5
PM10.25
1.111(1.028, 1.201)
1.111 (1.058, 1.166)
Odds Ratio (95% CI) per 25-,wg/m3
PM Increase
1.51(0.66,3.43)
1.77(1.24,2.55)
2.62(1.16,5.87)
PEFR change (L/min) per 25-^g/m3
PM Increase
am PEFR 5. 3 (2.6, 8.0)
pm PEFR +1.73 (5.67, -2.2)
pm PEFR -0.28 (2.86, -3.45)
am PEFR -4.31 (-11.44, 2.75)
16.2 (6, 29)
12.2 (max 68)

NR
NR
NR

27.07 (4.89, 69.07)
NR
NR
9.5(IQR5.1)
* Studies highlighted in the 1996 CD are in italics; new studies in plain text. For purposes of comparison across
 studies, results of single-pollutant models are presented in these tables; co-pollutant model results are presented
 and discussed in more detail in Chapter 8.
** Relative Risk (95% Confidence Interval), except for Fairley (1999) and Lipfert et al. (2000), where insufficient
data were available to calculate confidence intervals so p-value is given in parentheses.
*** Min, Max 24-h PM indicator level shown in parentheses unless otherwise noted as (±S.D.), NR = not reported,
or R = range of values from min-max, no mean value reported.
References:

 ASchwartz et al. (1996)
 BOstro et al. (2000)
 cLippmann et al. (2000)
 DLipfert et al. (2000a)
 EMar et al. (2000)
 FSmith et al. (2000)
    GFairley (1999)
    HBurnett et al. (2000)
    'Burnett etal. (1997)
    'Burnett etal. (1999)
    KTolbert et al. (2000)
    LSheppard et al. (1999)
                MSchwartz and Neas (2000)
                NNaeheretal. (1999)
                °Zhang et al. (2000)
                FNeasetal. (1999)
                QMoolgavkar (2000b)
April 2002
                9-104
          DRAFT-DO NOT QUOTE OR CITE

-------
>
to
o
o
to
H

6
o


o
H

O
                              Percent excess death (total unless otherwise noted) per

                                  25 ug/m3 increase in PM2.5 (•) or PM10.2.5 (o).



Harvard 6 Cities (recomputed)


Steubenville, OH


8 Canadian Cities

Chock et al (2000)

Pittsburgh, PA

Klemm and Mason (2000)

Atlanta, GA
Lipfert et al (2000a)
Philadelphia, PA ~
Lippmann et al (2000)

Detroit, Ml





Santa Clara Co.
Ostro et al. (2000)
Coachella Valley, CA
Castillejos et al (2000)

Mexico City, Mexico

Cifuentes et al. (2000)

5-4-3-2-10 1 234 56 7 8 9 10 11 12 13 14 1
i i i i i i i i i i i i i i i i i i i i


g / 0








w — ,
• } aye .* * 5




	 '-+• 	





n total "
v mortality
	 _ 	





Laq 5 day MA 1 • Q

	 V Q } All year
Laa 2 dav MA } 	 ^ — } Winter

o
HH
H
W
Figure 9-25.  Percent excess risks estimated per 25-jUg/m3 increase in PM2 5 or PM10_2 5 from new studies evaluating both

            PM2 5 and PM10_2 5 data for multiple years. All lags = 1 day, unless indicated otherwise.

-------
 1      molds, endotoxins, etc.) to the observed coarse particle effects in at least some locations.
 2      Overall, for U.S. and Canadian cities, effect size estimates for the coarse fraction (PM10_25) of
 3      those inhalable thoracic particles capable of depositing in TB and A regions of the respiratory
 4      tract generally appear to fall in the range of 0.5 to 6.0% excess total (nonaccidental) deaths per
 5      25 //g/m3 of 24-h PM10_25.  Respective increases for cause-specific mortality are 3.0 to 8.0% for
 6      cardiovascular and 3.0 to 16.0% for respiratory causes per 25-//g/m3 increase in 24-h PM10_25.
 7
 8           Chemical Components of Particulate Matter. Several new studies examined the role of
 9      specific chemical components of PM in relation to mortality risks. Studies of U.S.  and Canadian
10      cities showed mortality associations with one or more of several specific fine particle
11      components of PM, including H+, sulfate, nitrate, as well as COH; but their relative importance
12      varied from city to city, likely depending,  in part, on their concentrations (e.g., no clear
13      associations in those cities where H+ and sulfate levels were very low [i.e., circa nondetection
14      limits]). Figure 9-26 depicts relatively consistent estimates of total mortality excess risk
15      resulting from a 5-//g/m3 increase in sulfate, possibly reflecting impacts of sulfate per se or
16      perhaps sulfate serving as a surrogate for fine particles in general. Sulfate effect size estimates
17      generally fall in the range of 1 to 4% excess total mortality per 5-//g/m3 increase for U.S. and
18      Canadian cities.
19           A significant factor in some western cities is the occasional occurrence of high levels of
20      windblown crustal  particles that constitute the major part of the coarse PM fraction and a
21      substantial fraction of intermodal fine particles (PM^.j).  The small-size tail of the  windblown
22      crustal particles extends into the PM2 54 size range (intermodal), at times contributing
23      significantly to PM2 5. Claiborn et al. (2000) report that in Spokane, WA, PM2 5 constitutes about
24      30% of PM10 on dust event days, but 48% on days preceding the dust event. The intermodal
25      fraction represents  about 51% of PM2 5 during windblown dust events, about 28% on preceding
26      days. However, PMX in Spokane often shows little change during dust events, when coarse
27      particles (presumably crustal particles) are transported into the region.  The lack of increased
28      mortality during periods of time with high wind speeds and presumably high crustal material
29      concentrations was shown by Schwartz et al. (1999) for Spokane, and by Pope et al. (1999b) for
30      three cities  in the Wasatch front region of Utah. Other recent studies suggest that coarse
31      particles, as well as fine particles, may be associated with excess mortality in certain U.S.

        April 2002                                 9-106        DRAFT-DO NOT QUOTE OR CITE

-------
                                 Percent excess death (total mortality, unless otherwise noted)
                                              per 5 |jg/m3 increase in sulfate

Schwartz et al. (1996)
Six Cities
Burnett etal. (1998)
Toronto, Canada
Burnett et al. (2000)
Q 1 -jrnea
-------
                 TABLE 9-16. SUMMARY OF SOURCE-ORIENTED EVALUATIONS OF
                     PARTICULATE MATTER COMPONENTS IN RECENT STUDIES
         Author, City
Source Categories and Species with High Factor
Loadings Used to Suggest the Source Categories
     Source Categories Associated with
     Mortality. Comments.
         Laden et. al.,(2000)
         Harvard Six Cities
         1979-1988
Soil and crustal material: Si

Motor vehicle emissions: Pb

Coal combustion (Regional Sulfate):  Se, S

Fuel oil combustion:  V

Salt: Cl

Note: the trace elements are from PM2 5 samples
     The strongest increase in daily mortality
     was associated with the mobile source
     factor. The coal combustion factor was
     positively associated with mortality in all
     metropolitan areas, with the exception of
     Topeka.  The crustal factor from the fine
     particles was not associated with
     mortality.

     Coal and mobile sources account for the
     majority of fine particles in each city.
         Mar et al. (2000).
         Phoenix, AZ
         1995-1997
PM25 (from DFPSS) trace elements:

Motor vehicle emissions and resuspended road dust:
Mn, Fe, Zn, Pb, OC, EC, CO, and NO2

Soil: Al, Si, and Fe

Vegetative burning: OC and Ks (soil-corrected
potassium)

Local SO, sources: SO2

Regional sulfate:  S
                               PM10_25 (from dichot) trace elements:

                               Soil: Al, Si, K, Ca, Mn, Fe, Sr, and Rb

                               A source of coarse fraction metals: Zn, Pb, and Cu

                               A marine influence:  Cl
     PM, s factors results: Soil factor and local
     SO2 factor were negatively associated with
     total mortality. Regional sulfate was
     positively associated with total mortality
     on the same day, but negatively associated
     on the lag 3 day. Motor vehicle factor,
     vegetative burning factor, and regional
     sulfate factor were significantly positively
     associated with cardiovascular mortality.
                                                Factors from dichot PM10_2 5 trace elements
                                                were not analyzed for their associations
                                                with mortality because of the small sample
                                                size (every-third-day samples from June
                                                1996).
         Ozkaynak et al.
         (1996).
         Toronto, Canada.
Motor vehicle emissions: CO, COH, and NO2
     Motor vehicle factor was a significant
     predictor for total, cancer, cardiovascular,
     respiratory, and pneumonia deaths.
         Tsai et al. (2000).
         Newark, Elizabeth,
         and Camden, NJ.
         1981-1983.
Motor vehicle emissions: Pb and CO

Geological (Soil): Mn and Fe

Oil burning:  V and Ni

Industrial:  Zn, Cu, and Cd (separately)

Sulfate/secondary aerosol: Sulfate

Note:  The trace elements are from PM15 samples.
     Oil burning, industry, secondary aerosol,
     and motor vehicle factors were associated
     with mortality.
1       not positively associated with total mortality, with Mar et al. (2000) reporting a negative

2       association between the crustal component of PM2 5 and cardiovascular mortality.

3             However, these source-category-oriented evaluation results are derived from relatively

4       limited underlying analytic bases for resolving source categories and the identification of souce
        April 2002
                            9-108
DRAFT-DO NOT QUOTE OR CITE

-------
 1      categories must be viewed with caution at this time. For example, whereas Laden et al. (2000)
 2      had 6211 days of every-other-day data from the Harvard Six City Study of eastern/midwest U.S.
 3      cities, they had only elements in PM2 5 analyzed by X-ray fluorescence (XRF) spectroscopy (no
 4      organic PM or gases).  They used factors in the regression analysis and used Pb as a tracer to
 5      identify a motor vehicle source category, Se to identify a coal combustion source category, and Si
 6      as a tracer for soil. Since the "coal combustion" factor had a high loading of S as well as Se, it
 7      could equally as well have been identified as the regional sulfate source category. The "motor
 8      vehicle" and "coal combustion" sources were statistically significant for total  mortality as well as
 9      mortality resulting from ischemic heart disease and respiratory diseases (COPD plus pneumonia).
10      The crustal component had a negative association with total mortality.
11           The Mar et al. (2000) study had 3 years of pollutant data for Phoenix, AZ.  In addition to
12      elements determined by XRF, they had pollutant gases (CO, NO2, SO2, and O3) and total,
13      organic, and elemental carbon. They were able to identify five factors and attributed them to five
14      source categories.  Motor vehicles (plus resuspended road dust), vegetative burning, and regional
15      sulfate all had statistically significant associations with cardiovascular mortality,  but soil
16      (indexed by Si and Al, as crustal markers) had a statistically significant negative  association.
17      Also of importance, Mar et al. (2000) found significant associations between cardiovascular
18      mortality and PM25 and marginally significant (p < 0.10) associations between total mortality and
19      PM10.2.5.
20           Tsai et al. (2000) had only 156 days of data and used measurements of CO, sulfate, and
21      some elements; and they did not have Si, Ca,  Al,  or Mg as soil tracers nor Se as a tracer of coal
22      combustion, although much of the sulfate probably came from coal combustion.  They had three
23      fractions of extractable organic matter, but these did not appear to be useful in determining
24      factors. Statistically significant (p > 0.05) factors for both total daily deaths and combined
25      cardiovascular and respiratory daily deaths in at least one or another of the three New Jersey
26      cities studied (Newark, Camden, and Elizabeth) were attributed to motor vehicles, oil burning,
27      and sulfate. Also, an industrial source containing Zn and Cd was statistically  significant for total
28      deaths in Newark; and an industrial source containing Cd was marginally statistically significant
29      for cardiorespiratory disease in Elizabeth.
        April 2002                                9-109       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Ozkaynak et al. (1996) had only TSP, coefficient of haze (COH), and gases; however, they
 2      reported that a factor with COH, CO, and NO2 (considered to be representative of motor vehicle
 3      emissions) was associated with mortality in Toronto, Canada.
 4           None of these studies had measurements of nitrate or semivolatile organic compounds nor
 5      did they use the newest, and most effective, techniques for source apportionment. For example,
 6      using positive matrix factorization, Ramadan et al. (2000) were able to determine eight factors
 7      using the same data set as Mar et al. (2000).  In spite of these deficiencies, all four studies were
 8      able to associate one or more types of mortality with motor vehicles, several with coal
 9      combustion, and three with sulfate.
10           Factor analyses also were described briefly in a report by Lippmann et al. (2000). In that
11      study, neither sulfate nor acid aerosols were related significantly to morbidity or mortality, but
12      the concentrations were extremely low (with about 70% of the acid measurements below
13      detection limit).
14           It is difficult to compare these source-categories-related  assessments.  They are based on
15      different regions of the country over different periods of time when the sources of particles,
16      marker elements such as Pb, and other urban air pollutants were changing greatly. Also, each of
17      these studies constructed factors based on city-specific data. Thus, the factors in each study are
18      based on the idiosyncrasies of the specific data set for each city in the study, so the factors may
19      indeed represent different sources in different locations. Nevertheless, although somewhat
20      limited at this time, the new factor analysis results appear to implicate ambient PM derived from
21      fossil fuel (oil, coal) combustion and vegetative burning, as well as secondarily formed sulfates,
22      as important contributors to observed mortality effects, but not crustal particles.
23           In summary, the new evidence suggests that exposure to particles from several different
24      source categories, and of different composition and size, may have independent associations with
25      health outcomes. The excess risks from different types of combustion sources (coal, oil,
26      gasoline, wood, and vegetation) may vary from place to place  and from time to time, so that
27      substantial intra-regional and inter-regional heterogeneity would be expected. Likewise,
28      although earlier evaluations in the 1996 PM AQCD  seemed to indicate coarse particles and
29      intermodal particles of crustal composition as not likely being associated with adverse health
30      effects, there are now some reasonably credible studies suggesting that coarse particles (although
31      not necessarily those of crustal composition) may be associated with excess mortality in  at least

        April 2002                                 9-110        DRAFT-DO NOT QUOTE OR CITE

-------
 1      some locations. These notably include areas where past deposition of fine PM metals from
 2      smelter (Phoenix) or steel mills (Steubenville) onto surrounding soils may result in enhanced
 3      toxicity of later resuspended coarse (PM10_25) particles.
 4
 5      Updated Epidemiologic Findings for Long-Term Particulate Matter Exposure
 6      Effects on Mortality
 7           The 1996 PM AQCD indicated that past epidemiologic studies of chronic PM exposures
 8      collectively indicate increases in mortality to be associated with long-term exposure to airborne
 9      particles of ambient origins  (see appendix Table 9A-3).  The PM effect size estimates for total
10      mortality from these studies also indicated that a substantial portion of these deaths reflected
11      cumulative PM impacts above and beyond those exerted by acute exposure events. Table 9-17
12      shows long-term exposure effects estimates (RR values) per variable increments in ambient PM
13      indicators in U.S. and Canadian cities, including results from newer analyses since the 1996 PM
14      AQCD.
15           One of the most important advances since the 1996 PM AQCD is the substantial
16      verification and extension of the findings of the Six City prospective cohort study (Dockery
17      et al., 1993)  and the cohort study relating American Cancer Society (ACS) health data to
18      fine-particle data from 50 cities and sulfate data from 151 cities (Pope et al., 1995).  The
19      reanalyses, sponsored by the Health Effects Institute (HEI), included a data audit, replication of
20      the original investigators' findings, and additional analyses to explore the sensitivity of the
21      original findings to other model specifications. The investigators of the HEI Reanalysis Project
22      (Krewski et al., 2000) first performed a data audit, using random  samples to verify the accuracy
23      of the data sets used in the original  Six City analyses, including death certificate data,  air
24      pollution data, and socioeconomic data.  In general, the air pollution data were reproducible and
25      correlated highly with the original aerometric data in Pope et al. (1995).
26           The reanalyses substantially verified the findings of the original investigators, with PM25 or
27      sulfate relative risk (RR) estimates for total mortality and for cardiopulmonary mortality differing
28      at most by ±0.02 (±2% excess risk) from the least polluted to the most polluted cities in the
29      study. A larger difference was noted for the PM2 5 lung cancer relative risk in the Six  Cities
30      study, 1.37 originally and 1.43 in the reanalysis, neither estimate being statistically significant.
31      The sensitivity analyses for the Six Cities study found generally similar results with other

        April 2002                                9-111        DRAFT-DO NOT QUOTE OR CITE

-------
  TABLE 9-17. EFFECT ESTIMATES PER INCREMENTSA IN LONG-TERM MEAN
  LEVELS OF FINE AND INHALABLE PARTICLE INDICATORS FROM U.S. AND
                         CANADIAN STUDIES
Type of Health
Effect and Location
Increased Total Mortality
Six City11


ACS Study0
(151 U.S. SMSA)

Six City ReanalysisD

ACS Study ReanalysisD

ACS Study Extended
Analyses'3
Southern CaliforniaE



Indicator
in Adults
PM15/10(20^g/m3)
PM25(10^g/m3)
SO=4 (15 /ug/m3)
PM25(10^g/m3)
SO, (15 /ug/m3)
PM15/10 (20 Mg/m3)
PM25 (10 Mg/m3)
PM15/10 (20 ^g/m3)
(SSI)
PM25 (10 Mg/rn3)
PM25 (10 Mg/rn3)
PM10 (50 Mg/m3)
PM10 (cutoff =
30 days/year
>100 Mg/m3)
PM10 (50,wg/m3)
PM10 (cutoff =
30 days/year
>100 Mg/m3)
Increased Bronchitis in Children
Six at/
Six City0
24 City"
24 City"
24 City"
24 City"
Southern California1
12 Southern California
communities1
(all children)
12 Southern California
communitiesK
(children with asthma)
PM15/10(50^g/m3)
TSP (100 ^g/m3)
H+ (WOnmol/m3)
S0=4 (15 ^g/m3)
PM21 (25 ^g/m3)
PM10(50^g/m3)
S0=4(15^ig/m3)
PM10 (25 Mg/m3)
Acid vapor (1.7 ppb)
PM10 (19 Mg/m3)
PM25 (15 Mg/m3)
Acid vapor (1.8 ppb)
Change in Health Indicator per
Increment in PMa
Relative Risk (95% CI)
1.18(1.06-1.32)
1.13 (1.04-1.23)
1.46(1.16-2.16)
1.07 (1.04-1.10)
1.10(1.06-1.16)
1.19(1.06-1.34)
1.13 (1.04-1.23)
1.02 (0.99-1.04)
1.07(1.04-1.10)
1.04(1.01-1.08)
1.242 (0.955-1.616) (males)
1.082 (1.008-1. 162) (males)
0.879 (0.713-1.085) (females)
0.958 (0.899-1.021) (females)
Odds Ratio (95% CI)
3.26(1.13, 10.28)
2.80(1.17, 7.03)
2.65 (1.22, 5.74)
3.02(1.28, 7.03)
1.97 (0.85, 4.51)
3.29(0.81, 13.62)
1.39 (0.99, 1.92)
0.94(0.74, 1.19)
1.16(0.79, 1.68)
1.4(1.1, 1.8)
1.4(0.9,2.3)
1.1 (0.7, 1.6)
Range of City
PM Levels *
Means (^g/m3)

18-47
11-30
5-13
9-34
4-24
18.2-46.5
11.0-29.6
58.7(34-101)
9.0-33.4
21.1 (SD=4.6)
51 (±17)

51 (±17)


20-59
39-114
6.2-41.0
18.1-67.3
9.1-17.3
22.0-28.6
—
28.0-84.9
0.9-3.2 ppb
13.0-70.7
6.7-31.5
1.0-5.0 ppb
April 2002
9-112
DRAFT-DO NOT QUOTE OR CITE

-------
 TABLE 9-17 (cont'd). EFFECT ESTIMATES PER INCREMENTSA IN LONG-TERM
 MEAN LEVELS OF FINE AND INHALABLE PARTICLE INDICATORS FROM U.S.
                       AND CANADIAN STUDIES
Type of Health
Effect and Location
Increased Cough in Children
12 Southern California
communities'
(all children)
12 Southern California
communitiesK
(children with asthma)
Indicator

PM10 (25 Mg/m3)
Acid vapor (1.7 ppb)
PM10 (19 Mg/rn3)
PM25 (15 Mg/rn3)
Acid vapor (1.8 ppb)
Change in Health Indicator per
Increment in PMa
Odds Ratio (95% CI)
1.06 (0.93, 1.21)
1.13 (0.92, 1.38)
1.1 (0.0.8, 1.7)
1.3 (0.7, 2.4)
1.4(0.9,2.1)
Range of City
PM Levels *
Means Cwg/m3)

28.0-84.9
0.9-3.2 ppb
13.0-70.7
6.7-31.5
1.0-5.0 ppb
Increased Obstruction in Adults
Southern California1"
Decreased Lung Function in
Six at/
Six City0
24 City**
24 Cit/4
24 Cit/4
24 City**
12 Southern California
communitiesN
(all children)
12 Southern California
communitiesN
(all children)
12 Southern California
communities0
(4th grade cohort)
12 Southern California
communities0
(4th grade cohort)
PM10 (cutoff of
42 days/year
>100 Mg/m3)
Children
PM15/10(50^g/m3)
TSP (100 ^ig/m3)
H+ (52 nmoles/m3)
PM21(15^g/m3)
SO=(7 ^g/m3)
PM10(17^g/m3)
PM10 (25 Mg/m3)
Acid vapor (1.7 ppb)
PM10 (25 A^g/m3)
Acid vapor (1.7 ppb)
PM10(51.5Mg/m3)
PM25 (25.9 Mg/m3)
PM10_25 (25.6 Mg/m3)
Acid vapor (4.3 ppb)
PM10(51.5Mg/m3)
PM25(25.9,wg/m3)
PM10.2.5(25.6Mg/m3)
Acid vapor (4.3 ppb)
1.09(0.92, 1.30)

NS Changes
NS Changes
-3.45% (-4.87, -2. 01) FVC
-3.21% (-4.98, -1.41) FVC
-3.06% (-4.50, -1.60) FVC
-2.42% (-4.30, -.0.51) FVC
-24.9 (-47.2, -2.6) FVC
-24.9 (-65.08, 15.28) FVC
-32.0 (-58.9, -5.1) MMEF
-7.9 (-60.43, 44.63) MMEF
-0.58 (-1.14, -0.02) FVC growth
-0.47 (-0.94, 0.01) FVC growth
-0.57 (-1.20, 0.06) FVC growth
-0.57 (-1.06, -0.07) FVC growth
-1.32 (-2.43, -0.20) MMEF growth
-1.03 (-1.95, -0.09) MMEF growth
-1.37 (-2.57, -0.15) MMEF growth
-1.03 (-2.09, 0.05) MMEF growth
NR

20-59
39-114
6.2-41.0
18.1-67.3
9.1-17.3
22.0-28.6
28.0-84.9
0.9-3. 2 ppb
28.0-84.9
0.9-3. 2 ppb
NR
NR
April 2002
9-113
DRAFT-DO NOT QUOTE OR CITE

-------
         TABLE 9-17 (cont'd). EFFECT ESTIMATES PER INCREMENTSA IN LONG-TERM
         MEAN LEVELS OF FINE AND INHALABLE PARTICLE INDICATORS FROM U.S.
                                      AND CANADIAN STUDIES
Type of Health
Effect and Location
Decreased Lung Function in
Southern Californiap
(% predicted FEVb
females)
Southern California1"
(% predicted FEVb males)
Southern California1"
(% predicted FEVb males
whose parents had asthma,
bronchitis, emphysema)
Indicator
Adults
PM10 (cutoff of
54.2 days/year
>100 Mg/m3)
PM10 (cutoff of
54.2 days/year
>100 Mg/m3)
PM10 (cutoff of
54.2 days/year
>100 Mg/m3)
Range of City
Change in Health Indicator per PM Levels *
Increment in PMa Means (^g/m3)

+0.9 % (-0.8, 2.5) FEVj 52.7 (21.3, 80.6)
+0.3 % (-2.2, 2.8) FEVj 54.1 (20.0, 80.6)
-7.2 % (-11.5, -2.7) FEVj 54.1 (20.0, 80.6)
        Southern California1"
        (% predicted FEVj,
        females)

        Southern California1"
        (% predicted FEVb males)
          Not reported
      -1.5% (-2.9, -0.1)FEV!
                     7.4(2.7, 10.1)
                     7.3 (2.0, 10.1)
        *Range of mean PM levels given unless, as indicated, studies reported overall study mean (min, max), or mean
         (±SD); NR=not reported.
        AResults calculated using PM increment between the high and low levels in cities, or other PM increments given
         in parentheses; NS Changes = No significant changes.
       References:
       BDockery et al. (1993)
       GPope et al. (1995)
       DKrewski et al. (2000)
       EAbbey et al. (1999)
       FDockery et al. (1989)
       GWareetal. (1986)
       HDockery et al. (1996)
       'Abbey et al. (1995a,b,c)
      JPetersetal. (1999b)
      KMcConnell et al. (1999)
      LBerglund et al. (1999)
      MRaizenne et al. (1996)
      NPeters et al. (1999a)
      °Gauderman et al. (2000)
      FAbbeyetal. (1998)
      QPope et al. (2002)
1      individual covariates included. The time-dependent covariate model for total mortality (taking

2      into account higher postexposures in early years of the study and changes over time to the last

3      years of the study) had a substantially lower RR than the model without time-dependent

4      covariates. Educational level made a large difference, with individuals having less than a high

5      school education at much greater risk for mortality than those with any postsecondary education.
       April 2002
9-114
DRAFT-DO NOT QUOTE OR CITE

-------
 1           Among the ecological covariates, sulfates adjusted for artifact had little effect on the risk
 2      estimates for total mortality compared to that without adjustment, but, in the ACS study, the filter
 3      adjustment actually increased the relative risk for all causes and cardiopulmonary mortality,
 4      while substantially reducing the estimated sulfate effect on lung cancer. Inclusion of SO2 as an
 5      additional ecological covariate greatly reduced the estimated PM2 5 and sulfate effects in the ACS
 6      study, whereas a spatial model including SO2 effects caused only a modest reduction of the
 7      estimated PM2 5 and sulfate effects.  However, the SO2 effects were reduced greatly when sulfates
 8      were included in the model.  Sulfur dioxide and sulfates often are highly correlated, because of
 9      the formation of secondary sulfates.
10           Many model selection issues in the prospective cohort studies are analogous to those in the
11      time series analyses. One issue of particular concern is whether the exposure indices used in the
12      analyses adequately characterize the exposure of the participants in the study during the months
13      or years preceding death.  This question is particularly conspicuous in regard to the Pope et al.
14      (1995)  study, in which PM2 5 and sulfate data were collected in  the 1979 to 1982  period from the
15      EPA AIRS database and the Inhalable Particle Network, largely preceding the collection of the
16      ACS cohort data by only a few years, and so possibly not adequately reflecting exposure to
17      presumably much higher PM concentrations occurring long before the cohort was recruited, nor
18      exposure to presumably lower concentrations during the study.  This issue was raised in the 1996
19      PM AQCD.  However, the Six Cities Study did have air pollution data and repeated survey data
20      over time, with PM2 5 and sulfate data measured every other day and sometimes daily, and so the
21      new investigators were able to use the information about time-dependent cumulative PM
22      concentrations during the course of the study. Changes in smoking status and body mass index
23      over the 10 to!2 years of the study had little effect on risk estimates, but taking into account the
24      decrease in particle concentrations from the earlier years to the  later years reduced the effect size
25      estimate substantially, although it remained statistically significant. Nevertheless, overall, the
26      reanalyses of the ACS and Harvard  Six-Cities studies (Krewski et al., 2000) "replicated the
27      original results,  and tested those results against alternative risk  models  and analytic approaches
28      without substantively altering the original findings of an association between indicators of
29      particulate matter air pollution and mortality."
30           The shape of the relationship of concentration to mortality also was explored.  Preliminary
31      findings suggest some possible  nonlinearity, but further study is needed. Among the most

        April 2002                               9-115       DRAFT-DO NOT QUOTE OR CITE

-------
 1      important new findings of the study are spatial relationships between mortality and air pollution,
 2      discussed later below.
 3           Recently reported extension of the ACS analyses (Pope et al., 2002) to include additional
 4      years of data provides further substantiation of originally reported findings for total, respiratory,
 5      and cardiovascular mortality. Also of great importance, these new analyses provide much
 6      stronger evidence substantiating links between long-term ambient fine PM exposures and lung
 7      cancer.  This is consistent with findings of increased lung cancer risk being associated with
 8      exposure with diesel exhaust particles, an important constituent of PM2 5 in many U.S. urban
 9      areas.
10           With regard to the role of various PM constituents in the PM-mortality association, past
11      cross-sectional studies generally have found that the fine particle component, as indicated either
12      by PM25 or sulfates, was the PM constituent most consistently associated with chronic PM
13      exposure-mortality.  Although the relative measurement errors of the various PM constituents
14      must be further evaluated as a possible source of bias in these estimate comparisons, the Harvard
15      Six-Cities study and the latest reported AHSMOG prospective semi-individual study results
16      (Abbey, et al., 1999; McDonnell et al., 2000) are both indicative of the fine mass components of
17      PM likely being associated more strongly with the mortality effects of PM than coarse PM
18      components.  The ACS study, its reanalyses, and its recent extension all further substantiate
19      ambient fine particle effects, including increased risk not only of cardiopulmonary-related
20      mortality but lung cancer mortality as well.
21           Several  other new studies report epidemiologic  evidence indicating that: (a) PM exposure
22      early in pregnancy (during the first month) may be associated with slowed intrauterine growth
23      leading to low birth weight events (Dejmek et al., 1999); and (b) early postnatal PM exposures
24      may lead to increased infant mortality (Woodruff et al., 1997; Boback and Leon, 1999; Loomis
25      et al., 1999; Lipfert et al., 2000b).
26
27      9.12.2.2 Relationships of Ambient Particulate Matter Concentrations to
28              Morbidity Outcomes
29           New epidemiology studies add greatly to the overall database relating morbidity outcomes
30      to ambient PM levels.  These include much additional evidence for cardiovascular and
31      respiratory diseases being related to ambient PM. The newer epidemiology studies expand the

        April 2002                                9-116       DRAFT-DO NOT QUOTE OR CITE

-------
 1      evidence on cardiovascular (CVD) disease and are discussed first below, followed by discussion
 2      of respiratory disease effects with particular emphasis on newly enhanced evidence for
 3      PM-asthma relationships.
 4
 5      Cardiovascular Effects of Ambient Particulate Matter Exposures
 6           About 75% of all U.S. deaths occur in persons at least 65 years old, and, of these, nearly
 7      40% are for cardiac causes (nearly 45%, if deaths from cerebrovascular causes are also included).
 8      Thus, if ambient PM exposure indeed produces increased total mortality in the elderly, it would
 9      seem possible that cardiovascular (CVD) deaths may be involved.
10
11      Cardiovascular Hospital Admissions.  Just two studies were available for review in the 1996
12      PM AQCD that provided data on acute cardiovascular morbidity outcomes (Schwartz  and
13      Morris, 1995; Burnett et al., 1995). Both studies were of ecologic time series design using
14      standard statistical methods. Analyzing 4 years of data on the >65-year-old Medicare  population
15      in Detroit, MI,  Schwartz and Morris (1995) reported significant associations between ischemic
16      heart disease admissions and PM10, controlling for environmental covariates.  Based on an
17      analysis of admissions data from 168 hospitals throughout Ontario, Canada, Burnett and
18      colleagues (1995) reported significant associations between particle sulfate concentrations, as
19      well as other air pollutants, and daily cardiovascular admissions.  The relative risk because of
20      sulfate particles was slightly larger for respiratory than for cardiovascular hospital admissions.
21      The 1996 PM AQCD concluded on the basis of these  studies that, "There is a suggestion of a
22      relationship to heart disease, but the results are based on only two studies and the estimated
23      effects are smaller than those for other endpoints." The PM AQCD went on to state that acute
24      impacts on CVD  admissions had been demonstrated for elderly populations (i.e., >65), but that
25      insufficient data existed to assess relative impacts on younger populations.
26           Although the literature still remains relatively sparse, an important new body of data now
27      exists that both extends the available  quantitative information on relationships between ambient
28      PM pollution and hospital CVD admissions, and that, more intriguingly,  illuminates some of the
29      physiological changes that may occur on the mechanistic pathway leading from PM exposure to
30      adverse cardiac outcomes. Figure 9-27 depicts excess risk estimates derived from 10 studies of
31      acute PM10 exposure effects on CVD admissions in U.S. cities. Although new studies depicted

        April 2002                               9-117       DRAFT-DO NOT QUOTE OR CITE

-------
            Sametetal. (2000a,b) -
                14US Cities
                Schwartz (1999) -
                 8 US Counties
              Moolgavkar (2000c) -
                Maricopa, AZ

              Moolgavkar (2000c) -
                   LA.CA

              Moolgavkar (2000c) -
                Cook County

                Linn et al. (2000) -
                    LA.CA

                Schwartz (1997) -
                  Tucson,AZ
              Tolbert et al. (2000)  -
                  Atlanta
        Morris and Naumova (1998) -
               Chicago

            Lippmannet al.(2000) -
Total CVD
                                         Period 1 (AIRS Data)
                                    I            <	
                                                            CHF
                                 i    »    i
                                                                    Period 2 (Supersite Data)
                                                                   I	»	1
                                 -15        -10       -5         0         5         10
                                         Reconstructed Excess Risk Percentage
                                                  50 ug/m3 Increase in
       Figure 9-27.   Acute cardiovascular hospitalizations and PM exposure excess risk estimates
                     derived from selected U.S. PM10 studies. CVD = cardiovascular disease and
                     CHF = congestive heart failure.
1      in Figure 9-27 have reported generally consistent associations between daily hospitalizations for

2      cardiovascular disease and measures of PM, the data not only implicate PM, but also CO and

3      NO2 as well, possibly because of covarying of PM and these other gaseous pollutants derived

4      from common emission sources (e.g., motor vehicles). Taken as a whole, this body of evidence

5      suggests that PM is likely an important risk factor for cardiovascular hospitalizations in the

6      United States.
       April 2002
               9-118
DRAFT-DO NOT QUOTE OR CITE

-------
 1           For example, in the recently published NMMAPS 14-city analysis of daily CVD hospital
 2      admissions in persons 65 and older in relation to PM10 (Samet et al., 2000a,b). The mean risk
 3      estimate (for average 0-1 day lag) was a 8.5% increase in CVD admissions per 50 //g/m3 PM10
 4      (95% CI: 1.0 to 33.0%). No relationship was observed between city-specific risk estimates and
 5      measures of socioeconomic status, including percent living in poverty, percent non-white, and
 6      percent with  college educations.  In another study, remarkably consistent PM10 associations with
 7      cardiovascular admissions were observed across eight U.S. metropolitan areas, with a 25 //g/m3
 8      increase in PM10 associated with between 1.8 and 4.2 percent increases in admissions (Schwartz,
 9      1999).  Also, in a study of Los Angeles data from 1992-1995, PM10, CO, and NO2 were all
10      significantly  associated with increased cardiovascular admission in single-pollutant models
11      among persons 30 and  older (Linn et al., 2000). Moolgavkar (2000c) analyzed PM10, CO, NO2,
12      O3, and SO2 in relation to daily total cardiovascular (CVD) and total cerebrovascular admissions
13      for persons 65 and older from three urban counties (Cook, IL; Los Angeles, CA; Maricopa, AZ),
14      and found that, in univariate regressions, PM10 (and PM25 in LA) was associated with CVD
15      admissions in Cook and LA counties but not in Maricopa county.  On the other hand, in
16      two-pollutant models in Cook and LA counties, the PM risk estimates diminished and/or were
17      rendered nonsignificant.
18           The recent NMMAPS study of PM10 concentrations and hospital admissions by persons
19      65 and older  in 14 U.S. cities provides particularly important findings of positive and significant
20      associations,  even when concentrations are below 50 //g/m3 (Samet et al., 2000a,b). As noted in
21      Table 9-18, this study indicates PM10 effects similar to other cities, but with narrower confidence
22      bands, because of its greater power derived by combining multiple cities in the same analysis.
23      This  allows significant associations to be identified, despite the fact that many of the cities
24      considered have relatively small populations and that each of the  14 cities had mean PM10 below
25      50 //g/m3.
26
27      Physiologic Measures of Cardiac Function. Several very recent studies by independent groups
28      of investigators have also reported longitudinal associations between ambient PM concentrations
29      and physiologic measures of cardiovascular function. These studies measure outcomes and most
30      covariates at  the individual level, making it possible to draw conclusions regarding individual
31      risks, as well as to explore mechanistic hypotheses.  For example, several studies recently have

        April 2002                                9-119        DRAFT-DO NOT QUOTE OR CITE

-------
          TABLE 9-18. PERCENT INCREASE IN HOSPITAL ADMISSIONS PER 10-^g/m3
          	INCREASE IN 24-HOUR PMin IN 14 U.S. CITIES	
                                    CVD
          COPD
                   Pneumonia
Increase
(95% CI)
Increase
(95% CI)
Increase
(95% CI)
Constrained Lag Models (Fixed Effect Estimates)
One-day meana
Previous-day mean
Two-day meanb
PM10<50Mg.m3
(2 -day mean)b
Quadratic distributed lag
Unconstrained Distributed
Fixed effects estimate
Random effects estimate
1.07
0.68
1.17
1.47
1.18
Lag
1.19
1.07
(0.93, 1.22)
(0.54,0.81)
(1.01, 1.33)
(1.18, 1.76)
(0.96, 1.39)

(0.97, 1.41)
(0.67, 1.46)
1.44
1.46
1.98
2.63
2.49

2.45
2.88
(1.00, 1.89)
(1.03, 1.88)
(1.49, 2.47)
(1.71,3.55)
(1.78, 3.20)

(1.75,3.17)
(0.19,5.64)
1.57
1.31
1.98
2.84
1.68

1.90
2.07
(1.27, 1.87)
(1.03, 1.58)
(1.65,2.31)
(2.21, 3.48)
(1.25,2.11)

(1.46,2.34)
(0.94, 3.22)
        aLag.
        bMean of lag 0 and lag 1.
        Source: Samet et al. (2000a,b).
 1     reported temporal associations between PM exposures and various electrocardiogram (ECG)
 2     measures of heart beat or rhythm in panels of elderly subj ects. Reduced HR variability is a
 3     predictor of increased cardiovascular morbidity and mortality risks. Three independent studies
 4     reported decreases in HR variability associated with PM in elderly cohorts, although r-MSSD
 5     (one measure of high-frequency HR variability) showed elevations with PM in one study.
 6          Differences in methods used and results obtained across the studies argue for caution in
 7     drawing any strong conclusions yet regarding PM effects from them, especially in light of the
 8     complex intercorrelations that exist among measures of cardiac physiology, meteorology, and air
 9     pollution (Dockery et al., 1999).  Still, the new heart rhythm results, in general, comport well
10     with other findings of cardiovascular mortality and morbidity endpoints being associated with
11     ambient PM.  Chapter 5 discusses available exposure studies of elderly subjects with CVD, such
12     as the Sarnat et al. (2000) Baltimore study. Less active groups tend to have lower exposure to
13     nonambient PM because of reduced personal activity.  However, Williams et al. (2000a,b,c)
       April 2002
9-120
DRAFT-DO NOT QUOTE OR CITE

-------
 1      report a very high pooled correlation coefficient between PM2 5 personal exposure and outdoor
 2      concentrations.  These exposure studies tend to enhance the plausibility of panel study findings
 3      of impacts on HR variability being caused by exposure to ambient-generated PM.
 4
 5      Changes in Blood Characteristics. Additional epidemiologic findings (Peters et al., 1997a)
 6      also provide new evidence for ambient PM exposure effects on blood characteristics (e.g.,
 7      increased c-reactive protein in blood) thought to be associated with increased risk of serious
 8      cardiac outcomes (e.g., heart attacks).
 9
10      Key Conclusions Regarding PM-CVD Morbidity
11           Overall, the newly available studies of PM-CVD relationships appear to support the
12      following conclusions regarding several key issues:
13
14      Temporal Patterns of Response. The evidence from recent time series studies of CVD
15      admissions suggests rather strongly that PM effects are likely maximal  at lag 0, with some
16      carryover to lag 1.
17
18      Physical and Chemical Attributes Related to Particulate Matter Health Effects. The
19      characterization of ambient PM attributes associated with acute CVD is incomplete. Insufficient
20      data exist from the time series CVD hospital admissions literature or from the emerging
21      individual-level studies to provide clear guidance as to which PM attributes, defined either on the
22      basis of size or composition, determine potency.  The epidemiologic studies published to date
23      have been constrained by the limited availability of multiple PM metrics. Where multiple PM
24      metrics exist, they often are of differential quality because of differences in numbers of
25      monitoring sites and in monitoring frequency. Until more extensive and consistent data become
26      available for epidemiologic research, the question of PM size and composition, as they relate to
27      acute CVD impacts, will remain open.
28
29      Susceptible Subpopulations. Because they lack data on individual subject characteristics,
30      ecologic time series studies provide only limited information on susceptibility factors based on
31      stratified analyses.  The relative impact of PM on cardiovascular (and respiratory) admissions

        April 2002                               9-121        DRAFT-DO NOT QUOTE OR CITE

-------
 1      reported in ecologic time series studies is generally somewhat higher than those reported for total
 2      admissions.  This provides some limited support for the hypothesis that acute effects of PM
 3      operate via cardiopulmonary pathways or that persons with preexisting cardiopulmonary disease
 4      have greater susceptibility to PM, or both. Although there is some data from the ecologic time
 5      series studies showing larger relative impacts of PM on cardiovascular admissions in adults 65
 6      and over as compared with younger populations, the differences are neither striking nor
 7      consistent.  Some individual-level studies of cardiophysiologic function suggest that elderly
 8      persons with preexisting cardiopulmonary disease are susceptible to subtle changes in heart rate
 9      variability (HRV) in association with PM exposures. However, because younger and healthier
10      populations have not yet been assessed, it is not possible to say at present whether the elderly
11      have clearly increased  susceptibility compared to other groups, as indexed by cardiac
12      pathophysiological  indices such as HRV.
13
14      Role of Other Environmental Factors. The ecologic time series morbidity studies published  since
15      1996 generally have controlled  adequately for weather influences.  Thus, it is unlikely that
16      residual  confounding by weather accounts for the PM associations observed. With one possible
17      exception (Pope et al.,  1999a), the roles of meteorological factors have not been analyzed
18      extensively as yet in the individual-level studies of cardiac physiologic function.  Thus, the
19      possibility of confounding in such studies as yet cannot be discounted totally or readily.
20      Co-pollutants have  been analyzed rather extensively in many of the recent time series studies of
21      hospital  admissions and PM. In some studies, PM clearly carries an independent association
22      after controlling for gaseous co-pollutants. In others, the "PM effects" are reduced markedly
23      once co-pollutants are  added to the model. Among the gaseous criteria pollutants, CO has
24      emerged as the most consistently associated with cardiovascular (CVD) hospitalizations. The
25      CO effects are generally robust in the multi-pollutant model, sometimes as much  so as PM
26      effects.  However, the typically low levels of ambient CO concentrations in most such  studies
27      and minimal expected  impacts on carboxyhemoglobin levels and consequent associated hypoxic
28      effects thought to underlie CO CVD effects complicate  interpretation of the CO findings and
29      argue for the possibility that CO may be serving as a general surrogate for combustion  products
30      (e.g., PM) in the ambient pollution mix. See the most recent EPA CO Criteria Document (U.S.
31      Environmental Protection Agency, 2000a).

        April 2002                                9-122       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Respiratory Effects of Ambient Particulate Matter Exposures
 2           The number of studies examining hospitalization and emergency department visits for
 3      respiratory-related causes and other respiratory morbidity endpoints has increased markedly since
 4      the 1996 PM AQCD. In addition to evaluating statistical relationships for PM10, quite a few new
 5      studies also evaluated other PM metrics.  Those providing estimates of increased risk in U.S. and
 6      Canadian cities for respiratory-related morbidity measures (hospitalizations, respiratory
 7      symptoms, etc.) in relation to 24-h increments in ambient fine particles (PM2 5) or coarse fraction
 8      (PM10_25) of inhalable thoracic particles are included in Tables 9-12 and 9-13, respectively.
 9
10      Respiratory-Related Hospital Admission/Visits. PM hospital admissions/ visit studies that
11      evaluated excess risks  in relation to PM10 measures are still quite informative. Maximum excess
12      risk estimates for PM10 associations with respiratory-related hospital admissions and visits in
13      U.S. cities are shown in Figure 9-28. Nearly all the studies showed positive, statistically
14      significant relationships between ambient PM10 and increased risk for respiratory-related doctors'
15      visits and hospital admissions.  Overall, the results substantiate well ambient PM10 impacts on
16      respiratory-related hospital admissions/visits. The excess risk estimates fall most consistently in
17      the range of 5 to 25.0% per 50 //g/m3 PM10 increment, with those for asthma hospital admissions
18      and doctor's visits being higher than for COPD and pneumonia hospitalization.  Other, more
19      limited, new evidence  (not depicted in Figure 9-10) shows excess risk estimates for overall
20      respiratory-related or COPD hospital admissions falling in the range of 5 to 15.0% per 24-h
21      25 //g/m3 increment in PM25 or PM10_25. Larger estimates are found for asthma admissions or
22      physician visits, ranging up to ca. 40 to 50% for children <18 yr old in one study.
23           Of particular note in Figure 9-28 are the large effect size estimates now being reported for
24      asthma hospitalizations and visits. Very importantly, these hospital admission/visit studies and
25      other new studies on respiratory symptoms and lung function decrements in asthmatics are
26      emerging as possibly indicative of ambient PM likely being a notable contributor to exacerbation
27      of asthma. Additional evidence for PM-asthma effects is also emerging from panel studies of
28      lung function and respiratory symptoms, as discussed below.
29           New panel studies of lung function and respiratory symptoms in asthmatic subjects have
30      been conducted by more than 10 research teams in various locations world-wide. As a group, the
31      studies examine health outcome effects that are similar, such as pulmonary peak flow rate

        April 2002                                9-123       DRAFT-DO NOT QUOTE OR CITE

-------
               Tolbert et al. (2000) Atlanta  -
                Morris et al. (2000) Seattle  -
              Morris et al. (2000) Spokane  -
                Morris etal. (1999) Seattle  -
          Choudhury et al. (1997) Anchorage  -
         Nauenberg and Basu (1999) LA.CA  -
             Sheppard et al. (1999) Seattle  -
           Zanobetti et al. (2000a) Chicago  -
           Sametetal. (2000a) 14 US Cities  -
              Moolgavkar (2000b) Phoenix  -
               Moolgavkar (2000b) LA.CA  -
              Moolgavkar (2000b) Chicago  -
            Moolgavkar et al. (2000) King C  -
           Moolgavkar et al. (1997) Minn-SP  -
             Moolgavkar et al. (1997) Birm.  -
              Chen et al. (2000) Reno.NV  -
           Zanobetti et al. (2000a) Chicago  -
           Sametetal. (2000a) 14 US Cities  -









,


Asthma Visits




.
Asthma Hospital Admissions

I-H
«_! COPD Hospital Admissions
— '
(-»-!
«H
1 * 1
w Pneumonia Hospital Admissions
                                  -25
25      50       75      100
      Excess Risk, %
                      125
150
        Figure 9-28.  Maximum excess risk in selected studies of U.S. cities relating PM10 estimate
                      of exposure (50 jUg/m3) to respiratory-related hospital admissions and visits.
 1      (PEFR); and the studies typically characterize the clinical-symptomatic aspects in a sample of
 2      mild to moderate asthmatics (mainly children aged 5 to 16 yrs) observed in their natural setting.
 3      Their asthma typically is being treated to keep them symptom free (with "normal" pulmonary
 4      function rates, and activity levels) and to prevent recurrent exacerbations of asthma. Severity of
 5      their asthma is characterized by symptom, pulmonary function, and medication use and would be
 6      classified to include mild intermittent to mild persistent asthma suffers (National Institutes of
 7      Health, 1997).  As a group, they may thusly differ from asthmatics examined in studies of
 8      hospitalization or doctor visits for acute asthmatic episodes, who may have more severe asthma.
 9           Most studies reported ambient PM10 results, but PM2 5 was examined in two studies. Other
10      ambient PM measures (BS and SO4) also were used. For these studies, mean PM10 levels range
        April 2002
9-124
DRAFT-DO NOT QUOTE OR CITE

-------
 1      from a low of 13 //g/m3 in Finland to a high of 167 //g/m3 in Mexico City. The Mexico City
 2      level is over three times more than each of the other levels and is unique compared to the others.
 3      Related 95% CI for these means or ranges show 1-day maximums above  100 //g/m3 in four
 4      studies, with two of these above 150 //g/m3. Hence, these studies mainly evaluated different PM
 5      metrics indexing PM concentrations in the range found in U.S. cities (see Chapter 3). All the
 6      studies controlled for temperature, and several controlled for relative humidity.
 7           Many panel studies are analyzed using a design that takes advantage of the repeated
 8      measures on the same subject. Study subject number (N) varied from 12  to 164, with most
 9      having N >50; and all gathered adequate subject-day data to provide sufficient power for their
10      analyses. Linear models often are used for lung function and logistic models for dichotomous
11      outcomes. Meteorological variables are used as covariates; and medication use is also sometimes
12      evaluated as a dependent variable or treated as an important potential confounder. However,
13      perhaps the most critical choice in the model  is  selection of the lag for the pollution variable.
14      Presenting lag periods with only the strongest associations introduces potential bias,  because the
15      biological basis for lag structure may be related to effect. No biological bases for pertinent lag
16      periods are known, but some hypotheses can be proposed.  Acute asthmatic reactions can occur
17      4 to 6 h after exposure and, thus,  0-day lag may be  more appropriate than 1-day lags  for that
18      acute reaction.  Lag 1 may be more relevant for morning measurement of asthma outcome from
19      PM exposure the day before, and longer term lags (i.e., 2 to 5 days) may represent the outcome of
20      a more prolonged inflammatory mechanism; but too little information is now available to
21      predetermine appropriate lag(s).
22           Chapter 8 noted that people with asthma tend to have greater TB deposition than do healthy
23      people, but this data was not derived from the younger age group studied  in most asthma panel
24      studies.  The Peters et al.  (1997b) study is unique for two reasons:  (1) they  studied the size
25      distribution of the particles in the range 0.01 to  2.5  //m and (2) examined the number of particles.
26      They reported that asthma-related health effects of  5-day means of the number of ultrafine
27      particles were larger than those of the mass of the fine particles.  In contrast, Pekkanen et al.
28      (1997) also examined a range of PM sizes, but PM10 was more consistently  associated with PEF.
29      Delfino et al. (1998) is unique in  that they report larger effects for 1- and  8-h maximum PM10
30      than for the 24-h mean.


        April 2002                                9-125       DRAFT-DO NOT QUOTE OR CITE

-------
 1           The results for the asthma panels of the peak flow analysis consistently show small
 2      decrements for both PM10 and PM2 5.  The effects using 2- to 5-day lags averaged about the same
 3      as did the 0 to 1  day lags. Stronger relationships often were found with ozone. The analyses
 4      were not able to clearly separate co-pollutant effects. The effects on respiratory symptoms in
 5      asthmatics also tended to be positive. Most studies showed increases in cough, phlegm,
 6      difficulty breathing, and bronchodilator use.  The only endpoint more strongly related to longer
 7      lag times was bronchodilator use, which was observed in three studies. The peak flow
 8      decrements and respiratory  symptoms are indicators for asthma episodes.
 9           For PM10, nearly all of the point estimates showed decreases, but most were not statistically
10      significant, as shown in Figure 9-29 as an example of PEF outcomes. Lag 1 may be more
11      relevant for morning measurement of asthma outcome from the previous day. The figure
12      presents studies that provided this data. The results were consistent for both AM and PM peak
13      flow analyses. Similar results were found for the PM25 studies, although there were fewer
14      studies.  Several studies included PM25 and PM10 independently in their analyses of peak flow.
15      Of these, Gold et al. (1999), Naeher et al. (1999), Tiittanen et al. (1999), Pekkanen et al. (1997),
16      and Romieu et al. (1996) all found similar results for PM25 and PM10. The study of Peters et al.
17      (1997b) found slightly larger effects for PM25. The study of Schwartz and Neas (2000) found
18      larger effects for PM2 5 than for PM10_2 5. Naeher et al. (1999) found that FT was related
19      significantly to a decrease in morning PEF. Thus, there is no evidence here for a stronger effect
20      of PM2 5 when compared to PM10.  Also, of studies that provided analyses that attempted to
21      separate out effects of PM10 and PM2  5 from other pollutants, Gold et al. (1999) studied possible
22      interactive effects of PM25 and ozone on PEF; they found independent effects of the two
23      pollutants, but the joint effect was slightly less than the sum of the independent effects.
24           The effects on respiratory symptoms in asthmatics also tended to be positive, although
25      much less consistent than the lung function effects. Most studies showed increases in cough,
26      phlegm, difficulty breathing, and bronchodilator use (although generally not statistically
27      significant), as shown in Figure 9-30  for cough as an example. Three studies  included both PM10
28      and PM2 5 in their analyses.  The studies of Peters et al. (1997c) and Tiittanen  et al. (1999) found
29      comparable effects  for the two measures. Only the Romieu et al.  (1996) found slightly larger
30      effects for PM2 5. These studies also give no good evidence for a stronger effect of PM2 5 when
31      compared to PM10.

        April 2002                                9-126       DRAFT-DO NOT QUOTE OR CITE

-------
           Romieu etal. (1996)
                (Mexico)
         Pekkannen etal. (1997)
              (Finland)
            Gielenetal. (1997)
              (Netherlands)
           Romieu et al. (1997)
                (Mexico)
                           -10              -5              0               5
                                    Change in Pulmonary Function, L/min


Figure 9-29. Selected acute pulmonary function change studies of asthmatic children.
             Effect of 50 Aig/m3 PM10 on morning peak flow lagged 1 day.
            Vedaletal. (1998) -
               (Canada)
          Romieu etal. (1997)
               (Mexico)
           Gielenetal. (1997)
             (Netherlands)
          Peters et al. (1997c)
           (Czech Republic)
                                 -\	1	1	1	1-
                                  12345

                                         Odds Ratios for Cough
Figure 9-30. Odds ratios for cough for a 50-//g/m3 increase in PM10 for selected asthmatic
             children studies, with lag 0 with 95% CI.
April 2002
9-127
DRAFT-DO NOT QUOTE OR CITE

-------
 1           The results of PM10 peak flow analyses for nonasthmatic populations were inconsistent.
 2      Fewer studies reported results in the same manner as the asthmatic studies. Many of the point
 3      estimates showed increases rather than decreases. PM2 5 studies found similar results. The
 4      effects on respiratory symptoms in nonasthmatics were similar to those in asthmatics: most
 5      studies showed that PM10 increases cough, phlegm,  and difficulty breathing, but these increases
 6      were generally not statistically significant.  Schwartz and Neas (2000) found that PM10_2 5 was
 7      significantly related to cough. Tiittanen et al. (1999) found that 1-day lag of PM10_25 was related
 8      to morning PEF, but not evening PEF. Neas et al. (1999) found no association of PM10_25 with
 9      PEF in non-asthmatic subjects.
10
11      Long-Term Particulate Matter Exposure Effects on Lung Function and Respiratory
12      Symptoms
13           In the 1996 PM AQCD, the available respiratory disease studies were limited in terms of
14      conclusions that could be drawn. At that time, three studies based on a similar type of
15      questionnaire administered at three different times as part of the Harvard Six-City and 24-City
16      Studies provided data on the relationship of chronic respiratory disease to PM.  All three  studies
17      suggest a chronic PM exposure effect on respiratory disease.  The analysis of chronic cough,
18      chest illness, and bronchitis tended to be significantly positive for the earlier surveys described
19      by Ware et al. (1986) and Dockery et al. (1989). Using a design similar to the earlier one,
20      Dockery et al. (1996) expanded the analyses to include 24 communities in the United  States and
21      Canada. Bronchitis was found to be higher (odds ratio = 1.66) in the community with highest
22      exposure of strongly acidic particles when  compared with the least polluted community.  Fine
23      PM sulfate was  also associated with higher reporting of bronchitis (OR =  1.65, 95% CI1.12,
24      2.42).
25           The studies by Ware et  al. (1986), Dockery et al. (1989), and Neas et al. (1994) all had
26      good monitoring data and well-conducted standardized pulmonary function testing over many
27      years, but showed no effect on children of PM pollution indexed by TSP, PM15, PM25, or
28      sulfates. In contrast, the later 24-city analyses reported by Raizenne et al. (1996) found
29      significant associations of effects on FEVj or FVC in U.S. and Canadian children with both
30      acidic particles and other PM indicators. Overall, the available studies provided limited evidence
        April 2002                                9-128        DRAFT-DO NOT QUOTE OR CITE

-------
 1      suggestive of pulmonary lung function decrements being associated with chronic exposure to PM
 2      indexed by various measures (TSP, PM10, sulfates, etc.).
 3           A number of studies have been published since 1996 which evaluate the effects of
 4      long-term PM exposure on lung function and respiratory symptoms, as presented in Chapter 8.
 5      The methodology in the long-term studies varies much more than the methodology in the short-
 6      term studies. Some studies reported highly significant results (related to PM), whereas others
 7      reported no significant results.  Of particular note are several studies reporting associations
 8      between long-term PM exposures (indexed by various measures) or changes in such exposures
 9      over time and chronic bronchitis rates, consistent with the findings on bronchitis from the
10      Dockery et al. (1996) study noted above.
11           Unfortunately, the cross-sectional studies often are potentially confounded, in part, by
12      unexplained differences in geographic regions; and it is difficult to separate out results consistent
13      with a PM gradient from any other pollutants or factors having the same gradient.  The studies
14      that looked for a time trend also are confounded by other conditions that changed over time.  The
15      most credible cross-sectional study remains that described by Dockery et al. (1996) and Raizenne
16      et al. (1996). Whereas most studies include two to six communities, this study included 24
17      communities and is considered to provide the most credible estimates  of long-term PM exposure
18      effects on lung function and respiratory symptoms.
19
20      9.12.2.3 Methodological Issues
21           Chapter 8 discussed several still important methodological issues related to assessment of
22      the overall PM epidemiologic database.  These include, especially, issues related to model
23      specifications and consequent adequacy of control for potentially confounding of PM effects by
24      co-pollutants, evaluations of possible source relationships to pollutant effects that may be useful
25      in sorting out better effects attributable to PM versus other co-pollutants or both, and other issues
26      such as lag structure. Key points are discussed concisely below.
27
28      Time Series Studies: Confounding by Co-Pollutants in Individual Cities
29           The co-pollutant issue was discussed at length in the 1996 document and still remains an
30      important issue. It must be recognized that there are large differences in concentrations of
31      measured gaseous co-pollutants (and presumably unmeasured pollutants as well) in different

        April 2002                               9-129       DRAFT-DO NOT QUOTE OR CITE

-------
 1      parts of the United States, as well as the rest of the world; and the concentrations are often
 2      correlated with concentrations of PM and its components because of commonality in source
 3      emissions, wind speed and direction, atmospheric processes, and other human activities and
 4      meteorological conditions. Large sources in the United States include motor vehicle emissions
 5      (gasoline combustion, diesel fuel combustion, evaporation, particles generated by tire wear, etc.),
 6      coal combustion, fuel oil combustion, industrial processes, residential wood burning, solid waste
 7      combustion, and so on.  Thus, one might reasonably expect some large correlations among PM
 8      and co-pollutants, but possibly with substantial differences in relation by season in different
 9      cities or regions. Statistical theory suggests that PM and co-pollutant effect size estimates will be
10      highly unstable and often insignificant in multi-pollutant models when collinearity exists. Many
11      recent studies demonstrate this effect, for both hospital admissions (Moolgavkar, 2000b) and
12      mortality (Moolgavkar, 2000a; Chock et al., 2000). Because the problem seems largely insoluble
13      in studies in single cities, the new multi-city studies (Samet et al., 2000a,b; Schwartz, 1999;
14      Schwartz and Zanobetti, 2000) have provided important new insights. See discussions of
15      NMMAPS  analysis in Chapter 8 and below for discussion of issues related to control for
16      co-pollutant effects. Overall, although such issues may warrant further evaluation, it now
17      appears unlikely that such confounding accounts for the vast array of effects attributed to ambient
18      PM based on the rapidly expanding PM epidemiology database.
19           Numerous new studies have reported associations not only between PM, but also gaseous
20      pollutants (O3, SO2, NO2, and CO), and mortality. In many of these studies, simultaneous
21      inclusion of one or more gaseous pollutants in regression models did  not markedly affect PM
22      effect size estimates, as was generally the case in the NMMAPS analyses for 90 cities (see
23      Figure 9-31). On the other hand, some studies reporting positive and statistically  significant
24      effects for gaseous copollutants (e.g., O3, NO2, SO2, CO) found varying degrees of robustness of
25      their effects estimates or those of PM in multi-pollutant models as discussed in Chapter 8
26      (Section 8.4). Thus, it is likely that there are independent health effects of PM and gaseous
27      pollutants, there is not yet sufficient evidence by which to confidently separate out fully the
28      relative contributions of PM versus those of other gaseous pollutants  or by which  to quantitate
29      modifications of PM effects by other co-pollutants, including possible synergistic interactions
30      that may vary seasonally or from location to location.  Overall, it appears, however, that ambient
31      PM and O3  can be most clearly separated out as likely having independent effects, their

        April 2002                                 9-130       DRAFT-DO NOT QUOTE OR CITE

-------
                        PM10
                        PM10 + 03
                        PM10 + 03 + N02
                        PM10 + 03+SO2
                        PM10 + 03 + CO


_ _ _ _
— . —
_.._.._

1 .UU
1 .UU
1.00
1.00
1.00
                      0.0          0.2          0.4          0.6
                    % Change in Mortality per 10 |jg/m3 Increase in PM10
                                      1.0
      Figure 9-31.  Marginal posterior distributions for effect of PM10 on total mortality at lag 1,
                    with and without control for other pollutants, for the 90 cities. The numbers
                    in the upper right legend are the posterior probabilities that the overall
                    effects are greater than 0.
      Source: Samet et al. (2000a,b).
1     concentrations often not being highly correlated.  More difficulty is encountered, at times, in
2     sorting out whether NO2, CO, or SO2 are exerting independent effects in cities where they tend to
3     be highly correlated with ambient PM concentrations, possibly because of derivation of
4     important PM constituents from the same source  (e.g., NO2, CO, PM from mobile sources) or a
5     gaseous pollutant (e.g., SO2) serving as a precursor for a significant PM component (e.g.,
6     sulfate).  However, other information discussed in Section 8.4 on conceptual frameworks for
7     evaluating possible confounding makes it clear that diagnostic evaluations of inflation or
8     deflation of PM effect size estimates by addition  of gaseous co-pollutants into multiple pollutant
      April 2002
9-131
DRAFT-DO NOT QUOTE OR CITE

-------
 1      models, at best, may indicate potential confounding of PM effects in a given analysis. Other
 2      independently-derived exposure analyses, i.e., Sarnat et al. (2000, 2001), however, strongly
 3      suggest a very low probability of observed PM effects being due to confounding with gaseous
 4      criteria pollutants (CO, NO2, SO2, O3).
 5
 6      Time Series Studies: Model Selection for Lags, Moving Averages, and Distributed Lags
 1           A number of different approaches have been used to evaluate the temporal dependence of
 8      mortality or morbidity on time-lagged PM concentrations, including unweighted moving
 9      averages of PM concentrations over one or more days, general weighted moving averages, and
10      polynomial distributed moving averages. Unless there are nearly complete daily data, each
11      different lag will be using a different set of mortality data corresponding to spaced PM
12      measurement; for example, for lag 0 with every-sixth-day PM measurements, the mortality data
13      are on the same day as the PM data, for lag 1 the mortality data are on the next day after the PM
14      data, and so on. Although this effect is likely to be small, it should nonetheless be kept in mind.
15           The issue of dealing with lag structure, which may not necessarily be the same for all cities
16      or for all regions, can be illustrated by NMMAPS findings. As shown in Table 9-19, the rank
17      ordering of effects by lag days differs somewhat among NMMAPS regions. The combined data
18      set suggests that lag  1 provides the best fit, but with some regional differences. This raises the
19      question as to whether a single lag model should be assumed to characterize a diverse set of
20      regional findings. Because the particle constituents,  co-pollutants, susceptible subpopulations,
21      and meteorological covariates are likely to differ substantially from one region to another, the
22      timing of the largest mortality effects also may be presumed to differ in at least some cases. This
23      undoubtedly contributes to the variance of the estimated effects.
24           The distributed lag models used in the NMMAPS II morbidity studies are a noteworthy
25      methodological advance. The fitted distributed lag models showed significant heterogeneity
26      across cities for COPD and pneumonia, however (see Table 15 therein), again raising the
27      question of how heterogeneous  effects can best be combined so as not to obscure potentially real
28      city-specific or region-specific differences.
29           Only three cities with nearly complete daily PM10 data were used to evaluate more general
30      multi-day lag models (Chicago, Minneapolis/St. Paul, Pittsburgh), and these show somewhat
31      different patterns of effect, with lag 0 < lag 1 and lag 1 »lag 2 for Chicago, lag 0 = lag 1 > lag 2

        April 2002                               9-132        DRAFT-DO NOT QUOTE OR CITE

-------
              TABLE 9-19. PERCENT INCREASE IN MORTALITY PER 10 //g/m3 PM10
                      IN SEVEN U.S. REGIONS (from Figure 23 in NMMAPS II)
        Region                                 Rank Order of Effects by Lags
        Northwest                              lag 0 < lag 1 = lag 2
        Southwest                              lag 0 < lag 1< lag 2
        Southern California                      lag 0 < lag 1, lag 1 > lag 2, lag 0 < lag 2
        Upper Midwest                          lag 0 > lag 1, lag 0 > lag 2, lag 1 < lag 2
        Industrial Midwest                       lag 0 < lag 1, lag 1 > lag 2
        Northeast                               lag 0 < lag 1, lag 1 » lag 2
        Southeast                               lag 0 « lag  1, lag 1 > lag 2
        Combined	lag 0 < lag 1, lag 1 > lag 2	
 1     for Minneapolis, and lag 0 < lag 1 = lag 2 for Pittsburgh. The 7-day distributed lag model is
 2     significant for Pittsburgh, but less so in the other cities. The remaining data are limited
 3     intrinsically in what they can reveal about temporal structure.
 4
 5     Time Series Studies: Model Selection for Concentration-Response Functions
 6           Given the number of analyses that needed to be performed, it is not surprising that most of
 7     the NMMAPS studies focused on linear concentration-response models. More recent studies
 8     (Daniels et al., 2000) for the 20 largest U.S. cities have found posterior mean effects of 2 to 2.7%
 9     excess risk of total daily mortality per 50 //g/m3 24-h PM10 at lags 0, 1, 0+1 days; 2.4 to 3.5%
10     excess risk of cardiovascular and respiratory mortality; and 1.2 to 1.7% for other causes of
11     mortality. The posterior 95% credible regions are all significantly greater than 0. However, the
12     threshold models gave distinctly different estimates of 95% credible regions for the threshold for
13     total mortality (15 //g/m3 at lag 1, range 10 to 20), cardiovascular and respiratory mortality
14     (15 //g/m3 at lag 0+1, range 0 to 20), and other causes of mortality (65 //g/m3 at lag 0+1, range
15     50 to 75 //g/m3).
16           Another problem is that the shape of the relationship between mortality and PM10 may
17     depend, to some extent, on the associations of PM10 with gaseous co-pollutants.  The association

       April 2002                                9-133        DRAFT-DO NOT QUOTE OR CITE

-------
 1      is not necessarily linear, and is indeed likely to have both seasonal and secular components that
 2      depend on the city location.  Thus, further elaborations of these models may be desirable.
 3
 4      Effects of Exposure Error in Daily Time Series Epidemiology
 5           There has been considerable controversy over how to deal with the nonambient component
 6      of personal exposure. Recent biostatistical analyses of exposure error have indicated that the
 7      nonambient component will not bias the statistically calculated risk in community time-series
 8      epidemiology, provided that the nonambient component of personal  exposure is independent of
 9      the ambient concentration.  Consideration of the random nature of nonambient sources and recent
10      studies, in which estimates of a, ambient-generated PM divided by ambient PM concentrations,
11      have been used to estimate separately the ambient-generated and nonambient components of
12      personal exposure, support the assumption that the nonambient exposure is independent of the
13      ambient concentration. Therefore, it is reasonable to conclude that community time series
14      epidemiology describes statistical associations between health effects and exposure to ambient-
15      generated PM,  but does not provide any information on possible health effects resulting from
16      exposure to nonambient PM (e.g., indoor-generated PM).
17           From the point of view of exposure error, it is also significant to note that, although
18      ambient concentrations of a number of gaseous pollutants (O3, NO2,  SO2) often are found to be
19      highly correlated with various PM parameters, personal exposures to these gases are not
20      correlated highly with personal exposure to PM indicators.  The correlations of the ambient
21      concentrations  of these gases also are not correlated highly with the personal exposure to these
22      gases. Therefore, when significant statistical associations are found  between these gases and
23      health effects, it could be that these gases may, at times, be serving as surrogates for PM rather
24      than being causal themselves. Pertinent information on CO has not been reported.
25           The attenuation factor, a, is a useful variable.  For relatively constant a, the risk because of
26      a personal exposure to  10 //g/m3 of ambient PM is equal I/a times the risk from a concentration
27      of 10 //g/m3 of ambient PM, where a varies from a low of 0.1 to 0.2  to a maximum of 1.0. (The
28      health risk for an interquartile change in ambient concentration of PM is the same as that for an
29      interquartile change in  exposure to ambient PM). Differences in a among cities, reflecting
30      differences in air-exchange rates (e.g., because of variation in seasonal temperatures and in extent
31      of use of air conditioners) and differences in indoor/outdoor time ratios, may, in part, account for

        April 2002                                9-134        DRAFT-DO NOT QUOTE OR CITE

-------
 1      any differences in risk estimates based on statical associations between ambient concentrations
 2      and health effects for different cities or regions. If a were 0.3 in city A, but 0.6 in city B, and the
 3      risks for an increase in personal exposure of 10 //g/m3 were identical, then a regression of health
 4      effects on ambient concentrations would yield a health risk for city B that would be twice that
 5      obtained for city A.
 6           A number of exposure analysts have discussed the PM exposure paradox (i.e., that
 7      epidemiology yields statistically significant associations between ambient concentrations and
 8      health effects even though there is a near zero correlation between ambient concentrations and
 9      personal exposure in many studies).  Several explanations have been advanced to resolve this
10      paradox.  First, personal exposure contains both an ambient-generated and a nonambient
11      component.  Community time series epidemiology yields information only on the ambient-
12      generated component of exposure. Therefore, the appropriate correlation to investigate is the
13      correlation between ambient concentration and personal exposure to ambient-generated PM, not
14      between ambient concentrations and total personal exposure (i.e., the sum of ambient-generated
15      and nonambient PM).  Second, biostatistical analysis of exposure error indicates that if the risk
16      function is linear in the PM indicator, the average of the sum of the  individual risks (risk function
17      times individual exposure) may be replaced by the risk function times the community average
18      exposure. Thus, the appropriate correlation (of ambient concentrations and ambient-generated
19      exposure) is not the pooled correlation of different days and different people but the correlation
20      between the daily ambient concentrations and the community average daily personal exposure to
21      ambient-generated PM. Because the nonambient component is not  a function of the ambient
22      concentration, its average will tend to be similar each day.  Therefore, the correlation coefficient
23      will depend on a but not on the nonambient exposure. These types  of correlation yield high
24      correlation coefficients.
25           A few studies have conducted simulation analyses of effects of measurement errors on the
26      estimated PM mortality effects. These studies suggest that ambient PM excess risk effects are
27      more likely underestimated than overestimated, and that spurious PM effects (i.e., qualitative
28      bias such as change in the sign of the coefficient) because of transferring of effects from other
29      covariates require extreme conditions and  are therefore  very unlikely. The error because the
30      difference between the average personal exposure and the ambient concentration is likely the


        April 2002                                 9-135       DRAFT-DO NOT QUOTE OR CITE

-------
 1      major source of bias in the estimated relative risk.  One study also suggested that apparent linear
 2      exposure-response curves are unlikely to be artifacts of measurement error.
 3           In conclusion, for time-series epidemiology, ambient concentration is a useful surrogate for
 4      personal exposure to ambient-generated PM, although the risk per unit ambient PM
 5      concentration is biased low by the factor a compared to the risk per unit exposure to ambient-
 6      generated PM.  Epidemiologic studies of statistical associations between long-term effects and
 7      long term ambient concentrations compare health outcome rates across cities with different
 8      ambient concentrations. Ordinarily, PM exposure measurement errors are not expected to
 9      influence the interpretation of findings from either the community time-series or long-term
10      epidemiologic studies that have used ambient concentration data if they include sufficient
11      adjustments for seasonality and key personal and geographic confounders. When individual level
12      health outcomes are measured in small cohorts, to reduce exposure misclassification errors, it is
13      essential that better real-time exposure monitoring techniques be used and that further speciation
14      of indoor-generated, ambient, and personal PM mass be accomplished. This should enable
15      measurement (or estimation) of both ambient and nonambient components of personal exposure
16      and evaluation  of the extent to which personal exposure to ambient-generated PM, personal
17      exposure to nonambient PM, or total personal exposure (to ambient-generated plus nonambient
18      PM) contribute to observed health effects.
19
20      9.12.3 Coherence of Reported Epidemic logic Findings
21           Interrelationships Between Health Endpoints. Considerable coherence exists across
22      newly available epidemiologic study findings. For example, it was earlier noted that effects
23      estimates for total (nonaccidental) mortality generally fall in the range of 2.5 to 5.0% excess
24      deaths per 50 //g/m3 24-h PM10 increment.  These estimates comport well with those found for
25      cause-specific cardiovascular- and respiratory-related mortality. Furthermore, larger effect sizes
26      for cardiovascular (in the range of 3 to 6% per 50 //g/m3 24-h PM10 increment) and respiratory (in
27      the range of 5 to 25% per 50 //g/m3 24-h PM10) hospital admissions and visits are found, as
28      would be expected versus those for PM10-related mortality.  Also, several independent panel
29      studies, evaluating temporal associations between PM exposures and measures of heart beat
30      rhythm in elderly subjects, provide generally consistent indications of decreased heart rate (HR)
31      variability being associated with ambient PM exposure (decreased HR variability being an
        April 2002                                9-136       DRAFT-DO NOT QUOTE OR CITE

-------
 1      indicator of increased risk for serious cardiovascular outcomes, e.g., heart attacks). Other studies
 2      point toward changes in blood characteristics (e.g., increased C-reactive protein levels) related to
 3      increased risk of ischemic heart disease as also being associated with ambient PM exposures.
 4
 5           Spatial Interrelationships. Both the NMMAPS and Cohort Reanalyses studies had a
 6      sufficiently large number of cities to allow considerable resolution of regional PM effects within
 7      the "lower 48" states, but this approach was taken much farther in the Cohort Reanalysis studies
 8      than in NMMAPS. There were 88 cities with PM10 effect size estimates in NMMAPS; 50 cities
 9      with PM25 and 151 cities with sulfates in Pope et al. (1995) and in the reanalyses using the
10      original data; and, in the additional analyses by the cohort study reanalysis team, 63 cities with
11      PM25 data and 144 cities with sulfate data.  The relatively large number of data points allowed
12      estimation of surfaces for elevated long-term concentrations of PM2 5, sulfates, and SO2 with
13      resolution on a scale of a few tens to hundreds of kilometers. Information drawn from the maps
14      presented in Figures 16-21  in Krewski et al. (2000) is summarized below.
15           The patterns are similar, but not identical. In particular, the modeled PM25 surface
16      (Krewski, Figure 18) has peak levels in the industrial midwest, including the Chicago and
17      Cleveland areas, the upper  Ohio River Valley, and around Birmingham, AL. Lower, but
18      elevated, PM25 is found almost everywhere else east of the Mississippi, as well as in southern
19      California. This is rather similar to the modeled sulfate surface (Krewski, Figure 16), with the
20      absence of a peak in Birmingham and an emerging sulfate peak in Atlanta.  The only region with
21      elevated SO2 concentrations is the Cleveland-Pittsburgh area.  A preliminary evaluation is that
22      secondary sulfates in particles derived from local SO2 is more likely to  be important in the
23      industrial midwest, south from the Chicago-Gary region and along the upper Ohio River region.
24      This intriguing pattern may be related to the combustion of high-sulfur fuels in the subject areas.
25           The overlay of mortality and air pollution is also of interest. The spatial overlay of long-
26      term PM25 and mortality (Krewksi, Figure 21) is highest for the upper Ohio River region, but
27      also includes a significant association over most of the industrial midwest from Illinois to the
28      eastern noncoastal parts of North Carolina, Virginia, Pennsylvania, and New York. This is
29      reflected, in diminished form, by the sulfates map (Krewski, Figure 19) where the peak sulfate-
30      mortality associations occur somewhat east of the peak PM2 5-mortality associations.  The SO2
31      map (Krewski, Figure 20) shows peak associations similar to, but slightly east of, the peak

        April 2002                                9-137        DRAFT-DO NOT QUOTE OR CITE

-------
 1      sulfate associations. This suggests that, although SO2 may be an important precursor of sulfates
 2      in this region, there may be other considerations (e.g., metals) in the association between PM2 5
 3      and long-term mortality, embracing a wide area of the midwest and northeast (especially
 4      noncoastal areas).
 5           It should be noticed that, although a variety of spatial modeling approaches were discussed
 6      in the NMMAPS methodology report (NMMAPS Part I, pp. 66-71), the primary spatial analyses
 7      in the 90-city study (NMMAPS, Part II) were based on a simpler seven-region breakdown of the
 8      contiguous 48 states. The 20-city results reported for the spatial model in NMMAPS I show a
 9      much smaller posterior probability of a PM10 excess risk of short-term mortality, with a spatial
10      posterior probability versus a nonspatial probability of a PM10 effect of 0.89 versus 0.98 at lag 0,
11      of 0.92 versus 0.99 at lag 1, and of 0.85 versus 0.97 at lag 2.  The evidence that PM10 is
12      associated with an  excess short-term mortality risk is still moderately strong with a spatial model,
13      but much less strong than with a nonspatial model. In view of the sensitivity of the strength of
14      evidence to the spatial model, the model assumptions warrant additional study. Even so, there is
15      a considerable degree of coherence between the long-term and short-term mortality findings of
16      the studies, with stronger evidence of a modest but significant short-term PM10 effect and a larger
17      long-term fine particle (PM25 or sulfate) effect in the industrial midwest. The short-term effects
18      are larger but less certain in southern California and the northeast, whereas the long-term  effects
19      seem less certain there.
20
21
22      9.13 EVALUATION OF STATISTICAL AND MEASUREMENT
23           ERROR ISSUES
24      9.13.1  Errors Related to Concentration, Exposure, and Dose
25      What Is the Effect of Measurement Error and Misclassification on Estimates of the
26      Association Between Air Pollution and Health?
27           In PM epidemiology, statistical models are developed that relate heath effects to some
28      measurement of ambient PM.  However, if PM is toxic, the most direct relationship should be
29      between health effects and PM dose.  Therefore, in PM epidemiology, ambient PM
30      concentrations must be considered a surrogate for PM dose. In going from ambient PM
31      concentrations to PM dose, there are many possibilities for introducing error or variability. This

        April 2002                               9-138       DRAFT-DO NOT QUOTE OR CITE

-------
 1      section will discuss such possibilities and, to the extent information is available, the influence of
 2      such errors on the variability in epidemiologic results.
 3          Figure 9-32 shows an expanded version of the Risk Assessment Framework giving in more
 4      detail the various processes involved in going from PM sources to PM dose. In Figure 9-32,
 5      variables that can be measured directly are enclosed in hexagons; variables that cannot be
 6      measured directly but can be estimated are enclosed in diamonds; and processes that influence
 7      the relationship between ambient PM concentrations and PM dose are enclosed in ovals. There
 8      are many opportunities for error in going from ambient concentrations to dose.
 9
10      9.13.1.1  Opportunities for Error in the Use of Ambient PM Concentration as a
11              Surrogate for PM Dose in Epidemiologic Studies
12      Measurement ofPM Concentrations
13          As discussed in Chapter 2, Section 2.2.2.6, since there is no standard reference material that
14      can represent suspended PM, there cannot be any real determination of the accuracy with which
15      the concentration of suspended PM is measured. The precision of the measurement can be
16      determined by comparison of results from several collocated monitors.  The mass of PM,
17      collected on a filter and equilibrated for 24 hours at 25 C and 40% relative humidity according to
18      the Federal Reference Method, can be measured with high precision.  The precision of a
19      measurement of PM10_2 5 is normally less than that of PM2 5 but can be nearly as high if special
20      care is taken.  The measurement of ultrafme PM (PM01) presents special problems and little is
21      known about the accuracy or precision of such measurements.
22          As discussed in Chapter 2, Section 2.2.2.1, a major problem in the measurement of PM10
23      and especially PM2 5, is the variable loss of semivolatile components of PM. The most important
24      are particle-bound water (PBW), ammonium nitrate, and semivolatile organic compounds.
25      During equilibration, much of the PBW is lost and the remainder is stable at the low, constant
26      relative humidity of equilibration. However, variable fractions of the other semivolatile
27      components are also lost during  sampling or equilibration (Figure 9-33). For continuous
28      monitors, the collection surface must be changed at least every hour or the PM must be dried
29      in situ. Otherwise, changes in relative humidity will cause changes in the amount of PBW which
30      will cause large changes in perceived mass. Techniques which use heating to remove PBW may
31      also remove portions of the ammonium nitrate and  semivolatile organic compounds which in

        April 2002                                9-139       DRAFT-DO NOT QUOTE OR CITE

-------
                  Source to Exposure
                                                                 Cone, in Other
                                                                  Ambient ues
                                              Transport
                                            Transformation
                                             and removal
Sources of
Ambient PM
                         Gaseous Pre-
                         coursors of
                         Secondary PM
                                                            Ambient
                                                           Community
                                                             Cone.
                                              Transport
                                            Transformation
                                             and removal
      ue Outdoor at
        Home, i.e.
       Backyard ue
         Cone.
                                       ndoor Cone.
                                       of Ambient-
                                        Infiltrated
Modifications by  	
Indoor Chemistry
        Sources of
         Indoor-
      Generated PM
                                      Indoor Cone.
                                       of Indoor-
                                       Generated
                                         PM
                        Secondary Indoor
                                                                        Time-Activity Patterns, i.e
                                                                         Time in Various Ambient
                                                                          and Nonambient ues
      Personal Cloud
       or Personal
       Activity PM
                                                                                           Total
                                                                                          Personal
                                                                                         Respiratory
                                                                                          Exposure
                      Exposure to Dose
Figure 9-32.  An expanded version of the Risk Assessment Framework: (a) PM sources to
                PM exposure, (b) PM exposure to PM dose.
April 2002
                                          9-140
DRAFT-DO NOT QUOTE OR CITE

-------
                            Should be
                            retained
                                                           Particle-bound water
                                                            should be removed
                                         (NH4)XS04
                                          X= 0 to 2
                                         Aerodynamic Diameter, |jm
                            ::!: Semivolatile components subject to evaporation during or after sampling
       Figure 9-33.  Schematic showing major nonvolatile and semivolatile components of PM25.
                     Semivolatile components are subject to partial to complete loss during
                     equilibration or heating. The optimal technique would be to remove all
                     particle-bound water but no ammonium nitrate or semivolatile organic PM.
 1     some cases could be on the order of 50% of the total suspended PM mass. Thus, current
 2     measurements of PM mass have uncertainties and variability relative to the mass of PM
 3     suspended in the atmosphere.  Since most of the semivolatile components are in the
 4     accumulation mode, this is a more serious problem for PM2 5 measurements than for PM10_2 5
 5     measurements. As discussed in Chapter 2, Section 2.3, several new techniques are being tested
 6     which may allow removal of PBW without loss of the semivolatile components of PM.
 7     However, no such measurements of PM mass have been used in epidemiologic studies. Most
 8     currently available epidemiologic studies used PM indicators which measure only the relatively
 9     nonvolatile components of PM.  Likewise, except for the new studies using in-situ concentrated
10     ambient air particles, most toxicologic studies of ambient air particles have used filter-collected
11     material which contains only the relatively nonvolatile components of PM. Therefore, little
12     information is available on the possible health effects of the semivolatile components of PM.
       April 2002
9-141
DRAFT-DO NOT QUOTE OR CITE

-------
 1      Errors Due to Inadequate Resolution of PMMeasurements by Size, Composition, and Source
 2           Another source of error, with implications for epidemiology, is lumping together PM
 3      components that behave differently with respect to processes that influence the relationship
 4      between concentration and exposure, dose, and toxicity. This includes use of PM10
 5      measurements instead of separate measurements of PM2 5 and PM10_2 5, measurement of PM mass
 6      rather than individual chemical components, and measurement of mass instead of contributions
 7      from specific source categories. Examples of the results obtained form improved resolution are
 8      shown in Tables 9-20 through 9-22. As shown in Table 9-20, only two studies found a
 9      statistically significant relationship (t > 1.96) for both PM25 and PM10_25. In most cases, only one
10      or the other size fraction was significant.  However, in each case the most significant fraction
11      showed a higher % excess risk per//g/m3 and a greater t-statistic than was found for PM10. When
12      chemical components are treated separately (Table 9-21) the statistical significance may be
13      reduced compared to PM2 5, but the % excess risk per //g/m3 is higher than for PM2 5. Similarly,
14      when PM25 is split into orthogonal factors, representative of different source categories (Table
15      9-22), the % excess risk increases even though the t values  are slightly smaller. The inclusion of
16      some fraction of coarse mode particles in PM2 5 may be a source of error in locations with high
17      and variable concentrations of thoracic coarse PM. A related error may occur from the use of 24-
18      hour average concentrations if the health effect is more closely related to peak dose than to
19      integrated dose.
20
21      Frequency of PM Measurements
22           Most epidemiologic studies have relied on monitoring data from existing networks, some
23      of which provide only every-6th-day monitoring data. This represents a loss of information
24      compared to having every day monitoring data. In addition, a level of uncertainty is introduced
25      into the estimation of the lag structure (the variation of PM effect as a function the number of
26      days between exposure and the observation of the health effect) since each lag day is based on a
27      different day of health effects.  If everyday measurements of PM are available, each lag day is
28      based on health effects measured on the same day. The use of every-sixth day measurements
29      may also lead to errors in estimating annual and seasonal averages and distributions.
30
31

        April 2002                                9-142        DRAFT-DO NOT QUOTE OR CITE

-------
         TABLE 9-20. PERCENT EXCESS RISK (t-statistic) PER 10 //g/m3
     INCREASE IN PM FOR THE RELATIONSHIP OF VARIOUS INDICATORS
      OF PM WITH VARIOUS TYPES OF MORTALITY (CV = cardiovascular)
      IN SEVERAL DIFFERENT LOCATIONS. IN ONLY ONE CASE WERE
      BOTH PM2 5 AND PM10 2 5 SIGNIFICANT. IN MOST CASES, THE MORE
         SIGNIFICANT OF THE PM2 5 OR PM10 2 5 SIZE FRACTION HAD
          A LARGER % EXCESS RISK AND T-STATISTIC THAN PM10.
Location
Mortality
PM10
PM25
PMin.,,
Phoenix1
CV
1.9(2.5)
7.1(2.9)
2.3 (2.5)
Mexico2
Total
1.8 (4.2)
1.5(1.9)
4.1 (5.0)
Mexico2 Santa Clara, Co3 Boston4
CV
2.0 (2.4)
1.6(1.1)
4.8 (4.4)
Total
1.6*
3.1**
I 5***
Total
1.3 (4.9)
2.2 (6.3)
0.2 (0.6)
Steubenville4
Total
0.9 (2.2)
1.0(1.8)
2.4 (2.4)
6-Cities4
Total
0.8(5.8)
1.5 (7.4)
0.4(1.5)
 1. Mar et al., 2000. 2. Castillejos etal., 2000. 3. Fairley, 1999. 4. Swartzetal., 1996.
 Significant at: *, p = 0.05; **, p = 0.01; ***, not significant.
  TABLE 9-21. EXAMPLES OF HOW % EXCESS RISK PER 10 //g/m3 INCREASE IN
 PM INDICATOR INCREASES FOR SPECIFIC CHEMICAL COMPONENTS OF PM.
           IN THIS CASE, THE T-STATISTICS TEND TO BE LOWER.

 Location                       Santa Clara, Co1                6-Cities2

 PM25                               3.1                       1.5

 Sulfate                             17.4                       2.2

 Nitrate                              8.8                       —

 1. Fairley, 1999. 2. Swartzetal., 1996.
     TABLE 9-22. PERCENT EXCESS RISK (t-statistic) PER INTERQUARTILE
     INCREASE IN PM INDICATOR FOR THE RELATIONSHIP OF VARIOUS
  INDICATORS OF PM WITH CARDIOVASCULAR MORTALITY FOR PHOENIX
 (Mar et al., 2000). FACTORS, ESTIMATED USING A FACTOR ANALYSIS SOURCE
   APPORTIONMENT MODEL, ARE VEHICLE EXHAUST AND RESUSPENDED
     ROAD DUST (vehicle), VEGETATIVE BURNING (wood), AND REGIONAL
      SULFATE (R. SO4=). SOURCE CATEGORIES WERE SIGNIFICANT ON
                          DIFFERENT LAG DAYS.

PM25
Vehicle
Wood
R. SO4
Lag Day
1
1
O
0
%ER (t)
6.0 (2.9)
5.8 (2.6)
5.0 (2.7)
5.7 (2.0)
April 2002                           9-143      DRAFT-DO NOT QUOTE OR CITE

-------
 1      Spatial Variation
 2           Most epidemiologic analyses assume that the PM concentration is uniform across the
 3      spatial area in which health effects are measured or that the temporal variations in various parts
 4      of the spatial area are highly correlated.  As discussed in Chapter 3, Sections 3.2.5 and 3A, a lack
 5      of uniformity may lead to error due to low site-to-site correlations between daily concentrations
 6      or to spatial differences in long-term average concentrations. The site-to-site correlation is most
 7      important for acute epidemiologic studies that relate daily concentrations to daily health effects.
 8      Data from the PM2 5 monitoring network in 1999 and 2000 indicates relative high site-to-site
 9      correlations in many cities. However, site-to-site correlations may not be as high for chemical
10      components or source category contributions. The small amount of data available suggest lower
11      site-to-site correlations for PM10_2 5. In some cities, where PM air pollution is heavily influenced
12      by local point sources, site-to-site correlations of PM25 may be low. Such cities may not provide
13      the best data for time-series epidemiology.  Spatial differences  in average concentration may be
14      more important for studies of the effects of long term exposure to PM on longevity or rates of
15      disease. Spatial inhomogeneity, as found in cities with local sources of primary PM25, may be
16      more important for health effects that are nonlinear with PM dose in the range of PM dose
17      experienced.
18
19      The Difference Between Ambient PM Concentration and Ambient PM Exposure
20           As discussed in Chapter 5, Section 5.3, there are two sources of variability in the
21      relationship between ambient concentrations and exposure to ambient PM (also called ambient
22      PM exposure). The indoor environment is protective, in that the concentration of ambient PM
23      indoors is generally less than the concentration of ambient PM  outdoors. The relationship
24      depends on the particle size and on the air exchange rate.  Thus, there will be differences between
25      ultra fine, fine, and thoracic coarse PM and between air-conditioned and un-air-conditioned
26      homes.  The second source of variability is the fraction of time spent outdoors.  These two
27      sources of variability are combined into the attenuation factor, a, the ratio of ambient exposure to
28      ambient concentration. The product of a and the ambient PM concentration, C, yields the
29      ambient PM exposure, A (i.e., A = a  C) where a is different for the  different particle-size
30      fractions.  Since a may vary from person-to-person and time-to-time,  due to variations in  the air
31      exchange rate, the relationship between A and C may vary  across the population and across

        April 2002                                9-144       DRAFT-DO NOT QUOTE OR CITE

-------
 1      seasons. In spite of this variability, the correlation between A and C was high in the one study in
 2      which individual daily values of A were estimated. Variations in a, as indexed by air
 3      conditioning use, may explain some of the heterogeneity in excess rates of health effects
 4      observed in epidemiologic studies of PM10 in different cities.
 5
 6      Seasonal Variations
 1           Few epidemiological studies have had enough data for season-by-season analyses. Such
 8      differences might be expected due to seasonal variations in the relative concentrations of
 9      pollutants and PM components, the average a, correlations between ambient concentrations and
10      exposure, and correlations among potential surrogates and confounders. Most recent studies do
11      attempt to adjust for seasonal influences in their statistical models.
12
13      The Difference Between Ambient PM Exposure and Total PM Exposure
14           Total exposure to PM,  as measured by a monitor worn by a person, is composed of an
15      ambient exposure component and a nonambient exposure component.  The former includes
16      exposure to ambient pollution while outdoors and exposure to a fraction of the ambient pollution
17      while indoors. The latter is composed of primary and secondary indoor-generated PM and
18      personal cloud PM.  The nonambient exposure is found to be variable from day-to-day for a
19      given subject and to be variable from subject-to-subject on a given day. However, the average
20      daily nonambient exposure may be relatively constant not only within a given city, but from city
21      to city within developed countries.
22
23      Ambient Concentration—Personal Exposure Relationships for Gaseous Co-Pollutants
24           Ambient concentrations of gaseous co-pollutants, such as CO, NO2, SO2 and O3, are
25      sometimes used  in epidemiologic analyses.   Only a few studies have examined the correlations of
26      ambient copollutant concentrations with (1) ambient PM concentrations, (2) personal exposures
27      to co-pollutants, and (3) either ambient or total personal PM exposure.  These studies find that
28      the ambient concentrations of NO2, SO2, and O3 are not well correlated with personal exposure to
29      these gaseous  co-pollutants.  Rather, the concentrations of these gaseous co-pollutants and CO
30      are well correlated with the ambient concentrations of PM25,  the ambient exposures to PM25,  and
31      the total exposures to PM2 5.  Therefore, these studies conclude that ambient concentrations of

        April 2002                                9-145       DRAFT-DO NOT QUOTE OR CITE

-------
 1      NO2, SO2, O3, and likely CO, are not confounders but rather surrogates for ambient PM exposure
 2      or more likely of ambient exposure to source categories with which the gases are correlated, i.e.,
 3      NO2 and CO with motor vehicle associated PM and SO2 and O3 with regional sulfate.
 4
 5      The Difference Between Exposure and Dose
 6           As discussed in Chapter 6 there are several  causes of variability between exposure and
 7      dose.  The relationship between exposure and dose is highly dependent on particle size. Not only
 8      total deposition, but also the location of deposition, varies with particle size, as shown in
 9      Figure 9-34. The deposition fraction and location also depends on the size of the lung and the
10      breathing rate and is different for nose breathing and mouth  breathing. Thus, deposition is higher
11      during exercise than normal activity. Also, children, with smaller lungs and higher breathing
12      rates than adults, will have higher deposition fractions than adults.  Deposition fraction and
13      deposition location may be different in people with compromised lungs.  Very importantly,
14      deposition per unit surface area may be higher in  the healthy sections of their lungs.  It is not
15      currently known which of the various deposition parameters are most important, i.e., deposition
16      could be estimated as mass per body mass, particle surface area per lung surface area, or number
17      of particles per number of alveolar cells. The importance of exercise in influencing dose was
18      demonstrated in a recent study of asthma.  Exposure to O3 was related to increased prevalence of
19      asthma but only among children who participated in outdoor activities which involved exercise.
20
21      9.13.2  Possible Errors Related to  Health and Epidemiology
22      Resolution of Health Effects
23           As of 2001, the majority of PM epidemiology data was based on the relationship between
24      PM10 mass and total mortality. However, it is possible that different kinds of particles may cause
25      different kinds of health effects and  with different times between exposure and death.  Thus,
26      lumping all nonaccidental deaths together may obscure useful information.  Similar arguments
27      apply to morbidity. Some studies that consider classes of mortality tend to find higher excess
28      risks for cardiovascular and respiratory mortality  than for total mortality.
29
30


        April 2002                               9-146        DRAFT-DO NOT QUOTE OR CITE

-------
           .001   .002
                      .005  .01
                                       .05    .1    .2     .512
                                          Particle Diameter (|jm)
                                                                            10    20
                                                                                       50   100
      Figure 9-34.  The percent deposition of inhaled particles in the tracheobronchial (TB) and
                    alveolar (A) regions of the lung as a function of particle size.  The graph is
                    based on calculations using the ICRP model for a young adult with an
                    inhalation volume of 500 ml and a breathing frequency of 15 breaths a
                    second for spherical particles with a density of 1 g/ce.
1      Variation in Time Between Exposure and Appearance of Health Effects
2           Variations in toxicity of various types of PM or variations in the health status of members
3      of the exposed population may lead to variations in the time lag between exposure and
4      appearance of a response.  Therefore, studies need to account for responses that may lag exposure
5      by several days.  A dose of PM may also cause a health effect on more than one lag day.  If so,
6      and if day to day concentrations are correlated, as is the usual case for PM, the use of only one
7      lag day will overestimate the risk on the lag day selected, due to autocorrelation, but will
8      underestimate the total risk.  To obtain the total risk it is necessary to integrate the risk over
9      several days by a multiple regression model that accounts for health effects persisting for several
      April 2002
9-147
DRAFT-DO NOT QUOTE OR CITE

-------
 1
 2
 3
 4
 5
 9
10
11
12
13
14
15
16
17
18
19
20
21
22
days.  This technique, with either a constrained or unconstrained lag structure, or using a running
average of daily concentrations, accounts for autocorrelation in the day-to-day concentrations and
leads to higher estimated excess risks.

9.13.3  Apportioning Health Effects to PM (by size, chemical component,
        or source category) and Gaseous Co-Pollutants
     One of the important technical problems in air pollution epidemiology is properly
apportioning health effects related to air pollution to the proper PM size fraction, chemical
component, or source component or to one or more gaseous co-pollutants. A major problem in
epidemiology is that a study may attribute an effect to a measured variable used as a regressor
when another measured or unmeasured variable is really the causal agent.  The incorrect
attribution of effect (or part of the effect) to a variable used as a regressor, to another variable is
known as confounding.  The potential for confounding exists anytime the concentration of a
causal agent is significantly correlated with the measured concentration of the regressor. The
proper apportionment of effect in air pollution epidemiology is  difficult because PM and the
gaseous co-pollutants, NO2, CO, SO2, and O3 are often significantly correlated with each other.
The concepts of confounding; over-, under-, and mis-filtering, and effects modification are
discussed in Sections 8.1 and 8.4.  Consider the relationship diagramed in Figure 9-35 for two air
pollutants, A and B. Lines with double arrows indicate a statistically significant association
between the two variables. There are many possible variations  in the relationships among these
variables.
A
True Ambient
Concentration
1

B
True Ambient
Exposure
2
8

A
Measured
Ambient
Concentration

B
Measured
Ambient
Concentration
3
7

A
Exposure

B
Exposure
4
6

A
Outcome

5
B
Outcome
        Figure 9-35.  Diagram showing relationships (correlations) between A and B and between
                     various concentration, exposure, and outcome measures.
        April 2002
                                         9-148
DRAFT-DO NOT QUOTE OR CITE

-------
 1             1. All relationships, paths 1-8, are significant.
 2           la. In a multiple regression, A and B will share the health effects due to A and B. The
 3      split will depend on the differential error in the paths between concentrations and outcomes.
 4      Depending on the error structure and the relative strengths of the relationships, a portion of the
 5      health effect due to the pollutant with the higher error will be transferred to the pollutant  with
 6      the lower error. It will not be possible to accurately apportion the health effects between A and B
 7      (confounding).
 8           Ib. If A is used as a single regressor, some of the effect of A will be transferred to B and
 9      the effect of A will be overestimated (A is confounded by B, under-fitting).
10
11            2. Pollutant A does not cause the outcome of interest at the exposure level encountered.
12      Pathways 4 and 5 and outcome A disappear.
13           2a. Using B as a single regressor, the correct value is obtained for the association of B with
14      the outcomes due to B.
15           2b. Using A as a single regressor,  a false positive value is obtained for the effect of A due
16      to the correlation of A with B (A is a surrogate for B, mis-fitting).
17           2c. If A and B are used in a multiple regression, some of the effect of B will be transferred
18      to A and the true effect of B will be underestimated (over-fitting).
19
20            3. Pollutants A and B are independent and cause independent health effects. Pathways 1
21      and 5 disappear.  Since the concentration-outcome relationships are independently, an analysis
22      with either single or multiple regressors  will give the correct association  for each pollutant.
23      Situation 3 is the desirable situation.
24
25           Unfortunately, it is not possible, on the basis of one epidemiologic  study in one community
26      during  one time period, to tell whether A or B is responsible for the health effects; or if both are
27      responsible, to correctly apportion the effects to each. To correctly apportion the health effects
28      between A and B  it is necessary to seek other sources of information.
29
30           Toxicity. If we know that the potential confounder, A in situation 2,  does not cause
31      outcome A at the levels of exposure,  we know that a single regression with B as the regressor

        April 2002                                9-149       DRAFT-DO NOT QUOTE OR CITE

-------
 1      will yield the correct value for the association of B with outcome B (Situation 2a). However, a
 2      multiple regression using A and B would underestimate the association of outcome B with B
 3      because some of the health effects due to B would be incorrectly transferred to A (Situation 2c,
 4      over-fitting). Thus, if we know that A does not cause the outcome at the exposure level of A,
 5      then the health effect  may confidently be assigned to B (realizing that there could always be an
 6      unknown variable C,  correlated with B, that is really the causal agent).
 7
 8           Exposure. Useful information can also be obtained from concentration-exposure
 9      relationships. Suppose that A and B are significantly correlated but that the ambient
10      concentration of A is  not significantly correlated with either the ambient, nonambient, or total
11      personal exposure to A.  This can occur, as discussed in Chapter 5 and Section 9.6.4,  if the
12      spatial distribution of A is inhomogeneous of if A is very rapidly removed once it is penetrates
13      indoors. In this case pathway 4 disappears. Even is A is capable of causing the outcome at the
14      levels of exposure, a regression using A will not show any association because the exposure
15      causing the outcomes are not correlated with the concentration used as the regressor.
16
17           Lag structure.  If the effect due to A peaks on the day of exposure (lag day 0) but the
18      effect of B peaks on the day after exposure (lag day 1), we can conclude that both A and B cause
19      independent effects. This conclusion holds only if the concentration of A is not correlated with
20      the concentration of B on the prior day.
21
22           Multiple regression. In a number of studies the effects  attributed to A (PA, the  slope of the
23      regressions of A on outcomes) and B (PB) in a multiple regression are compared to those in single
24      regressions to obtain information on possible confounding. An increase in the variance of PAM
25      and PBM, from the multiple regression, over those PAS and PBS,  obtained from single regression; a
26      reduction in the values of either PAM and PBM compared to PAS  and PBS, or a decrease in the value
27      of P for the combined action of A and B relative to sum of PAS and PBS is evidence for potential
28      confounding. However, this information is not sufficient to determine whether A is a confounder
29      of B or whether A is a surrogate for B.  If, PBM equals PBS and  PAM = 0, it may be assumed that B
30      is causal and A is not. In order for this to happen, the correlation between A and B would have
31      to be non-significant.  Thus, if A and B are uncorrelated, PA is zero or non-significant, and PB is
        April 2002                                 9-150        DRAFT-DO NOT QUOTE OR CITE

-------
 1      positive and significant, a multiple regression, yielding similar values of PA and PB, will confirm
 2      that A is noncausal and B is causal. If the correlation is between 0 and 1, the multiple regression
 3      may provide some idea of how much of the effect of A could be due to potential confounding by
 4      B, but it cannot determine whether the confounding is real or only potential. If a community is
 5      studied where the correlation of A and B is low or nonsignificant, a comparison of single and
 6      multiple regression can demonstrate that significant confounding is not occurring if that is the
 7      case.
 8
 9           Orthogonal regressors.  Various types of source apportion models have been developed to
10      assist implementation of PM standards by identifying the sources  of PM in a given airshed. The
11      process involves application of statistical techniques such a factor analysis to daily concentration
12      values of PM components to generate orthogonal factors, containing various loadings of PM
13      components and in some cases, also containing gaseous co-pollutants. In some case, these
14      factors can be identified with specific source categories.  This source category factor (SCF) can
15      be used to determine the daily contributions of these source factors to the PM concentration.
16      Since the SCF are orthogonal (i.e., independent and uncorrelated) we have situation 3 and either
17      single regressions with each source factor or a multiple regression with all SCF should give
18      correct values of the relationship  of each SCF to the health outcomes with which it is related.
19      There will be no potential for confounding since the SCF are uncorrelated.
20           The concept of SCF can help explain some of the results from multiple regression of PM
21      and a gaseous co-pollutant.  Consider the situation shown in Figure 9-36. The vehicular
22      traffic-related (VTR) SCF and the regional sulfate (RSO4) SCF are uncorrelated and we will
23      assume they are causal.  The VTR SCF contains contributions from CO, NO2, and PM2 5. The
24      RSO4 SCF contain contributions from regional sulfate in the form of (NH4)2SO4, NH4H2SO4, and
25      H2SO4, but not local sulfate in the form of CaSO4. PM2 5 is correlated with both SCFs but CO
26      and NO2 are only correlated with the VTR SCF.
27           Now considered a multiple  regression using PM10 and CO (or NO2) as variables.  Since
28      PM2 5 contains both RSO4 and VTR components, it is possible that the ambient concentrations of
29      CO (or NO2) will be more highly correlated than PM2 5 with the ambient concentrations of the
30      VTR SCF. Thus, in a multiple regression, the effects of the VTR SCF would be transferred
31      largely to CO (or NO2) and only the effects of RSO4 would be transferred to the PM2 5.  Thus, CO

        April 2002                                9-151        DRAFT-DO NOT  QUOTE OR CITE

-------
True Ambient
Concentration
N<~>2 nr 4

\ CO *

TT
\ \ Motor
\ \ Vehicle
\ Related PM
"/ (MVR)
\ /
X Regional
/ Sulfate (RS)

	 * PM *












Measured
Ambient
Concentration
* * \




?
/



Exposure
/\ b 4
\
\ * *











Outcomes
Outcome NO2 or





i
CO


Outcome MVR

Outcome RS

i






       Figure 9-36.  Diagram showing concentrations—exposure—outcome relationships
                    (correlations for CO or NO2, PM2 5, and source category factors for vehicular
                    traffic related PM and regional sulfate.
 1
 2
 3
 4
 5
 6
 1
 8
 9
10
11
12
13
(or NO2) might show a higher relative risk than PM2 5 (or possibly than PM2 5 in a single
regression) and the relative risk associated with PM2 5 would be reduced in the multiple
regression. In this case, CO (or NO2) would be confounded by the VTR SCF.
     While a regression with SCFs will give excess risk values, unconfounded by other SCFs,
it will not be possible to tell from epidemiology alone whether the CO, the NO2, or the PM
component of the VTR pollution is truly causal. However, in view of the anticipated low
correlation between CO and NO2 with their respective personal exposures, and the unlikelihood
that CO or NO2 cause acute mortality as the very low values of personal exposures for most of
the population, the identification of the PM component of VTR pollution as the most likely
causal agent is reasonable.
       April 2002
                                         9-152
DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.14 IMPLICATIONS OF HEALTH EFFECTS OF LONG-TERM
 2           EXPOSURES TO PARTICULATE MATTER
 3           What are the implications of observed effects of long-term exposure to paniculate matter
 4      and other pollutants for life expectancy?
 5
 6      9.14.1  Methodological Issues
 7           Closed-cohort studies of ambient air pollutants are methodologically similar to typical
 8      epidemiological studies of occupational cohorts and, in some respects, to experimental trials.
 9      Subjects are enrolled, characterized as to their exposures and other relevant health factors, and
10      followed over time as they experience adverse health outcomes.  Methodological issues
11      regarding the loss of subjects to follow-up, the movement of subjects between exposure groups
12      or levels, and the characterization of exposure are well-understood and are adequately handled by
13      standard epidemiologic methods.
14           The assignment of exposure in both environmental and occupational studies is generally
15      based on area rather than personal sampling and any consequential exposure misclassification
16      will generally bias effect estimates towards the null. With appropriate individual-level
17      assessment and analysis of other risk factors, the assignment of a common exposure to a group
18      does not give raise to an ecological fallacy (Kunzli and Tager,  1997). The current PM AQCD
19      has avoided a reliance on purely ecological analyses of county-level data that lack
20      individual-level data on non-environmental determinants of mortality.
21           A key difference between epidemiologic closed-cohorts studies and experimental trials is
22      the lack of randomization of subjects to exposure. In observational studies, randomization is
23      replaced by a careful consideration and analytic correction for differences in other salient health
24      factors other than the exposure of interest. A potential confounder must be (a) an independent
25      determinant of the outcome of interest among unexposed subjects, (b) non-causally associated
26      with the exposure of interest, and (c) not a part of the causal pathway linking the exposure and
27      outcome. Once these potential confounders have been controlled, differences in survival, that is,
28      in the relative rates of mortality, are attributed to differences in subjects' exposure histories.
29
30

        April 2002                               9-153        DRAFT-DO NOT QUOTE OR CITE

-------
 1      9.14.2  Overall Survival and Life Expectancy
 2           Our current knowledge of the adverse health effects of long-term exposures to ambient
 3      particulate matter is based on a small number of epidemiological studies that compare differences
 4      in the survival of well-characterized closed-cohorts of free-living human subjects with air
 5      pollution levels in their cities of residence (AQCD Section 8.2.3).  Compared with the more
 6      intricate methodological aspects of epidemiological studies of short-term particulate matter
 7      exposures using time-series methods to examine non-enumerated open-cohorts, the design,
 8      conduct and analysis of closed-cohorts is straightforward.  However, such survival studies are
 9      much less common due to the difficulty and expense of enrolling and maintaining follow-up of
10      an enumerated cohort.
11           At the time of the 1996 PM AQCD, three closed-cohort (survival) studies of particulate
12      matter had been published in the peer-review literature. Two of these survival  studies were
13      national in scope, the Harvard Six-Cities Adult Cohort Study (Dockery et al., 1993) and the
14      American Cancer Society Cohort Study (Pope et al., 1995), and one focused solely on California,
15      the Adventist Health Study of Smog or AHSMOG (Abbey et al., 1991). The American Cancer
16      Society Cohort Study was  a secondary analysis of a very extensive cohort of 552,138  subjects in
17      151 cities whose exposures were characterized by routinely collected air quality data and who
18      were followed for seven years. The Harvard Six-Cities Adult Cohort Study enrolled 8,111
19      subjects in six cities, characterized their exposures with investigator-conducted measurements of
20      size-fractionated particulate matter, and followed these subjects for 14 to 16 years.  The
21      Adventist Health  Study on Smog enrolled 6,340 non-smoking subjects, grouped into three major
22      urban areas and the remainder of California, whose exposures were characterized by routinely
23      collected air quality data, and who were followed for an average of 10 years.
24           The two national studies found strong associations between higher particulate matter levels
25      and decreased survival.  For non-external causes of mortality, a 25 //g/m3 increment in PM25 was
26      associated with increases in the rate of mortality: 36 percent in the Harvard  Six-Cities Adult
27      Study and 18 percent in the American Cancer Society Cohort Study. The California study did
28      not initially find any statistically significant overall mortality effects.
29           After the 1996 PM AQCD was completed, concerns were expressed regarding the adequacy
30      of the conduct and analysis of these survival studies of particulate matter (Gamble,  1998). Many

        April 2002                                9-154        DRAFT-DO NOT  QUOTE OR CITE

-------
 1      concerns related to standard methodological issues regarding the assessment of exposure,
 2      geographic mobility, and adequacy of control for potential confounders.
 3
 4      9.14.3  Verification and Sensitivity Analyses
 5           Since the 1996 PM AQCD, the two national studies have been critically reanalyzed by
 6      independent researchers under the auspices of the Health Effects Institute (Krewski et al., 2000).
 7      In addition to the replication and validation of the original findings of the Harvard Six-Cities
 8      Adult Study, Krewski et al. considered the sensitivity of the original findings to alternative risk
 9      models and analytic approaches. Generally this sensitivity analysis found that the original results
10      were robust to changes in model specification and the inclusion of other community-level
11      covariates. Both the original and the reanalyses found a 13% increase in risk of mortality per
12      10 //g/m3 increment in PM25. The HEI reanalysis project both confirmed and extended the
13      results of the American Cancer Society Cohort Study. Both the reanalysis and the extension
14      found a 7% increase in risk of mortality per 10 //g/m3 in PM2 5.
15           Since the conclusion of the reanalysis project, these three survival cohorts have been
16      extended by the original investigators to  additional years of follow-up and alternative exposure
17      measures.  Using airport visibility records to estimate exposures to PM25, the Adventist Health
18      Study on Smog reported an 8.5  percent increase in the rate of non-external mortality associated
19      with a 10 //g/m3 increment in PM25 (McDonnell et al., 2000).
20           Thus, the relative risk  estimates for these three survival cohorts have converged in the
21      range of 7 to 13 percent increase in the non-external mortality rate associated with a 10 //g/m3
22      increment in a long-term average of PM2 5. Methodological criticisms of these studies have been
23      largely resolved in favor of the validity of their original findings of a strong  association between
24      long-term exposures to particulate matter and decreased survival (Bates, 2000).
25
26      9.14.4  Impact on Life-Expectancy
27           The increased rate of non-external  mortality found in these three survival cohorts is greater
28      than the mere accumulation of the adverse effects of short-term exposures for a few days.
29      Conceptually, particulate matter may be  associated with both the long-term development of
30      underlying health problems ("Frailty") and with the short-term variations in  timing of mortality

        April 2002                               9-155        DRAFT-DO NOT QUOTE OR CITE

-------
 1      among a susceptible population with some underlying health condition (Kunzli et al. 2001).
 2      Epidemiologic studies of the mortality effects of short-term exposure to particulate matter using
 3      unenumerated open-cohorts ("time-series studies") can only capture particulate matter's
 4      association with short-term variations in mortality and, therefore, must systematically
 5      underestimate the proportion of total mortality attributable to particulate matter. A recent
 6      time-series study that examined the contribution of daily particulate matter levels over an
 7      extended lag period (42 days) could only partially bridge the gap between the effects of
 8      short-term and long-term exposures to particulate matter (Zanobetti et al., 2002).
 9           Recent investigations of the public health implications of effect estimates for long-term PM
10      exposures also were reviewed in Chapter 8.  Life table calculations by Brunekreef (1997) found
11      that relatively small differences in long-term exposure to airborne PM of ambient origin can have
12      substantial effects on life expectancy. For example, a calculation for the 1969 to 1971 life table
13      for U.S. white males indicated that a chronic exposure increase of 10 //g/m3 PM was associated
14      with a reduction of 1.31 years for the entire population's life expectancy at age 25. The new
15      evidence noted above of infant mortality associations with PM exposure suggests that life
16      shortening in the entire population from long-term PM exposure could well be significantly
17      larger than estimated by Brunekreef (1997).
18
19      9.14.5 Specific Causes of Death
20           The increase in non-external mortality cannot be explained by increases in chronic
21      respiratory diseases since chronic non-malignant lower respiratory disease accounts for only
22      5.6 percent and lung cancer for only another 6.9 percent of all deaths over age 24 years due to
23      non-external causes.  Cardiovascular diseases, which account for 43  percent of non-external
24      mortality, must play the leading role in the decreased survival associated with  exposure to
25      ambient PM.  It is nevertheless useful to highlight the newer results of the extension of the ACS
26      study analyses (that include more years of participant follow-up and  address previous criticisms
27      of the earlier ACS analyses), which provide the strongest evidence to date that long-term ambient
28      PM exposures are associated with increased risk of lung cancer. That increased risk appears to
29      be in about the same range as that seen for a non-smoker residing with a smoker and, therefore,
30      passively exposed chronically to tobacco smoke, with any  consequent life-shortening impacts
31      due to lung cancer.
        April 2002                                9-156        DRAFT-DO NOT QUOTE OR CITE

-------
REFERENCES

Abbey, D. E.; Ostro, B. E.; Petersen, F.; Burchette, R. J. (1995) Chronic respiratory symptoms associated with
      estimated long-term ambient concentrations of fine particulates less than 2.5 microns in aerodynamic
      diameter (PM2 5) and other air pollutants. J. Exposure Anal. Environ. Epidemiol. 5: 137-159.
Abbey, D. E.; Burchette, R. J.; Knutsen, S. F.; McDonnell, W. F.; Lebowitz, M. D.; Enright, P. L. (1998) Long-term
      paniculate and other air pollutants and lung function in nonsmokers. Am. J. Respir. Crit. Care Med.
      158: 289-298.
Abbey, D. E.; Nishino, N.; McDonnell, W. F.; Burchette, R. J.; Knutsen, S. F.; Beeson, W. L.; Yang, J. X. (1999)
      Long-term inhalable particles and other air pollutants related to mortality in nonsmokers. Am. J. Respir. Crit.
      Care Med. 159: 373-382.
Adams, P. F.; Hendershot, G. E.; Marano, M. A. (1999) Current estimates  from the National Health Interview
      Survey, 1996. Vital Health Stat. 10(200).
Bigger, J. T., Jr.; Fleiss, J. L.; Steinman, R. C.; Rolnitzky, L. M.; Kleiger, R. E.; Rottman, J. N. (1992) Frequency
      domain measures of heart period variability and mortality after myocardial infarction.  Circulation
      85: 164-171.
Bobak, M.; Leon, D. A. (1999) Pregnancy outcomes and outdoor air pollution: an ecological study in districts of the
      Czech Republic 1986-8. Occup. Environ. Med. 56: 539-543.
Broday, D. M.; Georgopoulos, P. G. (2000) Growth and deposition of hygroscopic paniculate matter in the human
      lungs. Aerosol Sci. Technol.: submitted.
Brunekreef, B. (1997) Air pollution and life expectancy: is there a relation? Occup. Environ. Med. 54: 781-784.
Burnett, R. T.; Dales, R. E.; Raizenne, M. E.; Krewski, D.; Summers, P. W.; Roberts, G. R.; Raad-Young, M.;
      Dann, T.; Brook, J. (1994) Effects of low ambient levels of ozone and sulfates on the frequency of
      respiratory admissions to Ontario hospitals. Environ. Res. 65:  172-194.
Burnett, R. T.; Dales, R.; Krewski, D.; Vincent, R.; Dann, T.; Brook, J. R.  (1995) Associations between ambient
      paniculate sulfate and admissions to Ontario hospitals for cardiac and respiratory diseases. Am. J. Epidemiol.
      142: 15-22.
Burnett, R. T.; Cakmak, S.; Brook, J. R.; Krewski, D. (1997) The role of paniculate size and chemistry in the
      association between summertime ambient air pollution and hospitalization for cardiorespiratory diseases.
      Environ. Health Perspect. 105: 614-620.
Burnett, R. T.; Cakmak, S.; Raizenne, M. E.; Stieb, D.; Vincent, R.; Krewski, D.; Brook, J. R.; Philips, O.;
      Ozkaynak, H. (1998) The association between ambient carbon monoxide levels and daily mortality in
      Toronto, Canada. J. Air Waste Manage. Assoc. 48: 689-700.
Burnett, R. T.; Smith-Doiron, M.; Stieb, D.; Cakmak, S.; Brook, J. R. (1999) Effects of paniculate and gaseous air
      pollution on cardiorespiratory hospitalizations. Arch. Environ. Health 54: 130-139.
Burnett, R. T.; Brook, J.; Dann, T.; Delocla, C.; Philips, O.; Cakmak, S.; Vincent, R.; Goldberg, M. S.; Krewski, D.
      (2000) Association between paniculate- and gas-phase components of urban air pollution and daily mortality
      in eight Canadian cities. In: Grant, L. D., ed. PM2000: paniculate matter and health. Inhalation Toxicol.
      12(suppl. 4): 15-39.
Castillejos, M.; Borja-Aburto, V. H.; Dockery, D. W.; Gold, D. R.; Loomis, D. (2000) Airborne coarse particles and
      mortality. In: Inhalation Toxicology: proceedings of the third colloquium on paniculate air pollution and
      human health; June, 1999; Durham, NC. Inhalation Toxicology 12(suppl. 1): 61-72.
Chen, L.; Yang, W.; Jennison, B. L.; Omaye, S. T. (2000) Air paniculate pollution and hospital admissions for
      chronic obstructive pulmonary disease in Reno, Nevada. Inhalation  Toxicol. 12: 281-298.
Chock, D. P.; Winkler, S.; Chen, C.  (2000) A study of the association between daily  mortality and ambient air
      pollutant concentrations in Pittsburgh, Pennsylvania. J. Air Waste Manage. Assoc. 50: 1481-1500.
Choudhury, A. H.; Gordian, M. E.; Morris, S.  S. (1997) Associations between respiratory illness and PM10 air
      pollution. Arch. Environ. Health 52: 113-117.
Cifuentes, L. A.; Vega, J.; Kopfer, K.; Lave, L. B. (2000) Effect of the fine fraction of paniculate matter versus the
      coarse mass and other pollutants on daily mortality in Santiago, Chile. J. Air Waste Manage. Assoc.
      50: 1287-1298.
Claiborn, C. S.; Finn, D.; Larson, T. V.; Koenig, J. Q. (2000) Windblown dust contributes to high PM2 5
      concentrations. J. Air Waste Manage. Assoc. 50: 1440-1445.
April 2002                                      9-157         DRAFT-DO NOT QUOTE OR CITE

-------
Clark, W. E.; Whitby, K. T. (1975) Measurements of aerosols produced by the photochemical oxidation of SO2 in
      air. J. Colloid Interface Sci. 51: 477-490.
Clyde, M. A.; Guttorp, P.; Sullivan, E. (2000) Effects of ambient fine and coarse particles on mortality in Phoenix,
      Arizona. J. Exposure Anal. Environ. Epidemiol.: submitted.
Code of Federal Regulations. (1991a) Appendix J to Part 50-reference method for the determination of paniculate
      matter as PM10 in the atmosphere. C. F. R. 40: §50.
Code of Federal Regulations. (1991b) Ambient air monitoring reference and equivalent methods. C. F. R. 40: §53.
Costa, D. L.; Dreher, K. L. (1999) What do we need to know about airborne particles to make effective risk
      management decisions? A toxicology perspective.  Hum. Ecol. Risk Assess. 5: 481-491.
Daniels, M.; Dominici, F.; Samet, J. M.; Zeger, S. L. (2000) Estimating paniculate matter-mortality dose-response
      curves and threshold levels: an analysis of daily time-series for the 20 largest US cities. Am. J. Epidemiol.
      152: 397-406.
Dejmek, J.; Selevan, S. G.; Benes, L; Solansky, L; Sram,  R. J. (1999) Fetal growth and maternal exposure to
      paniculate matter during pregnancy. Environ. Health Perspect. 107: 475-480.
Delfino, R. J.; Murphy-Moulton, A. M.; Burnett, R. T.; Brook, J. R.; Becklake, M. R. (1997) Effects of air pollution
      on emergency room visits for respiratory illnesses  in Montreal, Quebec. Am.  J. Respir. Crit. Care Med.
      155: 568-576.
Delfino, R. J.; Zeiger, R. S.; Seltzer, J. M.; Street, D. H. (1998) Symptoms inpediatric asthmatics and air pollution:
      differences in effects by symptom severity, anti-inflammatory medication use and particulate averaging time.
      Environ. Health Perspect. 106: 751-761.
Dockery, D. W.;  Speizer, F. E.; Stram, D. O.; Ware, J. H.; Spengler, J. D.; Ferris, B. G., Jr. (1989) Effects of
      inhalable particles on respiratory health of children. Am. Rev. Respir. Dis. 139: 587-594.
Dockery, D. W.;  Pope, C. A., Ill; Xu, X.;  Spengler, J. D.; Ware, J. H.; Fay, M. E.; Ferris, B. G., Jr.; Speizer, F. E.
      (1993) An association between air pollution and mortality in six U.S. cities. N. Engl. J. Med.
      329: 1753-1759.
Dockery, D. W.;  Cunningham, J.; Damokosh, A. L; Neas, L. M.; Spengler, J. D.; Koutrakis, P.; Ware, J. H.;
      Raizenne,  M.; Speizer, F. E. (1996) Health effects of acid aerosols on North American children: respiratory
      symptoms. Environ. Health Perspect. 104: 500-505.
Dockery, D. W.;  Pope, C. A., Ill; Kanner, R. E.; Villegas, G. M.; Schwartz, J. (1999) Daily changes in oxygen
      saturation and pulse rate associated with particulate air pollution and barometric pressure. Cambridge, MA:
      Health Effects Institute; research report no. 83.
Driscoll, K. E.; Costa, D. L.; Hatch, G.; Henderson, R.; Oberdorster, G.; Salem, H.;  Schlesinger, R. B. (2000)
      Intratracheal instillation as an exposure technique for the evaluation of respiratory tract toxicity: uses and
      limitations. Toxicol. Sci. 55: 24-35.
Elder, A. C. P.; Gelein, R.; Finkelstein J. N.; Cox, C.; Oberdorster,  G. (2000) Endotoxin priming affects the lung
      response to ultrafine particles and ozone in young  and old rats. In: Inhalation Toxicology: proceedings of the
      third colloquium on particulate air pollution and human health; June, 1999; Durham, NC. Inhalation
      Toxicology 12(suppl. 1): 85-98.
Fairley, D. (1999) Daily mortality and air pollution in Santa Clara County, California:  1989-1996. Environ. Health
      Perspect. 107: 637-641.
Fanucchi, M. V.; Plopper, C. G. (1997) Pulmonary developmental responses to toxicants. In: Roth, R. A., ed.
      Toxicology of the Respiratory System, v. 8 of Comprehensive Toxicology. New York, NY: Pergamon,
      p. 203-220.
Fanucchi, M. V.; Wong, V. J.; Hinds, D.;  Tarkington, B.  K.; Van Winkle, L. S.; Evans, M. J.; Plopper, C. G. (2000)
      Repeated episodes of exposure to ozone alters postnatal development of distal conducting airways  in infant
      rhesus monkeys. Am. J. Respir. Crit. Care Med.  161: A615.
Federal Register. (1987) Revisions to the national ambient air quality standards for particulate matter. F. R. (July 1)
      52: 24,634-24,669.
Fitz-Simons, T. S.; Mathias, M.; Rizzo, M. (2000) Analyses of 1999 PM data for the PM NAAQS review.
      U.S. Environmental Protection Agency, Office of Air Quality Planning and Standards; September 29.
Gardner, S. Y.; Lehmann, J. R.; Costa, D. L. (2000) Oil fly ash-induced elevation of plasma fibrinogen levels in
      rats.  Toxicol. Sci. 56: 175-180.
Gauderman, W. J.; Mcconnell, R.; Gilliland, F.; London,  S.; Thomas, D.; Avol, E.; Vora, H.; Berhane, K.;
      Rappaport, E. B.; Lurmann, F.; Margolis, H. G.; Peters, J. (2000) Association between air pollution and lung
      function growth in southern California children. Am. J. Respir. Crit. Care Med. 162: 1383-1390.


April 2002                                     9-158         DRAFT-DO NOT QUOTE OR CITE

-------
Gielen, M. H.; VanDer Zee, S. C.; Van Wijnen, J. H.; Van Steen, C. I; Brunekreef, B. (1997) Acute effects of
      summer air pollution on respiratory health of asthmatic children. Am. J. Respir. Crit. Care Med.
      155:2105-2108.
Gilmour, P. S.; Brown, D. M.; Lindsay, T. G.; Beswick, P. H.; MacNee, W.; Donaldson, K. (1996) Adverse health
      effects of PM10 particles: involvement of iron in generation of hydroxyl radical. Occup. Environ. Med.
      53: 817-822.
Godleski, J. J.; Sioutas, C.; Katler, M.; Koutrakis, P. (1996) Death from inhalation of concentrated ambient air
      particles in animal models of pulmonary disease. Am. J. Respir. Crit. Care Med. 153: A15.
Godleski, J. J.; et al. (1998) Increased cardiac vulnerability during exposure to inhaled environmental particles
      [abstract]. Presented at: Health Effects Institute annual meeting; May; Boston, MA. Boston, MA: Health
      Effects Institute.
Godleski, J. J.; Verrier, R. L.; Koutrakis, P.; Catalano, P. (2000) Mechanisms of morbidity and mortality from
      exposure to ambient air particles. Cambridge, MA: Health Effects Institute;  research report no. 91.
Gold, A.; Litonjua, J.; Schwartz, M.; Verrier, R.; Milstein, A.; Larson, E.; Lovett, B. (1998) Cardiovascular
      vulnerability to paniculate pollution. Presented at: 1998 International conference [American Thoracic
      Society]; April; Chicago, IL.  Am. J. Respir. Crit. Care Med. 157: A261.
Gold, D. R.; Damokosh, A. I.; Pope, C. A., Ill; Dockery, D. W.; McDonnell, W. F.; Serrano, P.; Retama, A.;
      Castillejos, M. (1999) Particulate and ozone pollutant effects on the respiratory function of children in
      southwest Mexico City. Epidemiology 10: 8-16.
Gold, D. R.; Litonjua, A.; Schwartz, J.; Lovett, E.; Larson, A.; Nearing, B.; Allen,  G.; Verrier, M.; Cherry, R.;
      Verrier, R. (2000) Ambient pollution and heart rate variability. Circulation 101: 1267-1273.
Goldberg, M. S.; Bailar,  J. C., Ill; Burnett, R. T.; Brook, J. R.; Tamblyn, R.; Bonvalot, Y.; Ernst, P.; Flegel, K. M.;
      Singh,  R. K.; Valois, M.-F. (2000) Identifying  subgroups of the general population that may be susceptible to
      short-term increases in paniculate air pollution: a time-series study in Montreal, Quebec. Cambridge, MA:
      Health Effects Institute; research report 97. Available: http://www.healtheffects.org/pubs-research.htm [15
      February, 2001].
Goldsmith, C.-A. W.; Hamada, K.; Ning,  Y. Y.; Qin,  G.; Catalano, P.; Murthy, G. G. K.; Lawrence, J.; Kobzik, L.
      (1999) The effects of environmental aerosols on airway hyperresponsiveness in a murine model of asthma.
      Inhalation Toxicol. 11:  981-998.
Gordon, T.; Nadziejko, C.; Chen, L. C.; Schlesinger, R. (2000) Effects of concentrated ambient particles in rats and
      hamsters: an exploratory study. Cambridge, MA: Health Effects Institute; research report no. 93.
Gwynn, R. C.; Burnett, R. T.; Thurston, G. D. (2000)  A time-series analysis of acidic paniculate matter and daily
      mortality and morbidity in the Buffalo, New York, region. Environ. Health Perspect. 108: 125-133.
Hayano, J.; Sakakibara, Y.; Yamada, M.;  Ohte, N.; Fujinami, T.; Yokoyama, K.; Watanabe, Y.; Takata, K. (1990)
      Decreased magnitude of heart rate  spectral components in coronary artery disease. Its relation to
      angiographic severity. Circulation 81:  1217-1224.
Hofmann, W.; Balashazy, I.; Heistracher, T.; Koblinger, L. (1996) The significance of particle deposition patterns in
      bronchial airway bifurcations for extrapolation modeling. Aerosol Sci. Technol. 25: 305-327.
Hoyert, D. L.; Kochanek, K. D.; Murphy, S. L. (1999) Deaths: final data for 1997.  Atlanta, GA: U.S. Centers for
      Disease Control and Prevention, National Center for Health Statistics. (National vital statistics reports: v. 47,
      no. 19).
Hsieh, T. H.; Yu, C. P. (1998) Two-phase pulmonary  clearance of insoluble particles in mammalian species.
      Inhalation Toxicol. 10:  121-130.
Kamens, R.; Jang, M.; Chien, C.-J.; Leach, K. (1999)  Aerosol formation from the reaction of a-pinene and ozone
      using a gas-phase kinetics-aerosol partitioning  model. Environ. Sci. Technol. 33: 1430-1438.
Keywood, M. D.; Ayers, G. P.; Gras, J. L.; Gillett, R.  W.; Cohen, D. D. (1999) Relationships between size
      segregated mass concentration data and ultrafine particle number concentrations in urban areas.
      Atmos. Environ. 33: 2907-2913.
Kim, Y. J.; Boatman, J. F.; Gunter, R. L.; Wellman, D. L.; Wilkison, S. W. (1993)  Vertical distribution of
      atmospheric aerosol size distribution over south-central New Mexico. Atmos. Environ. Part A
      27: 1351-1362.
Kittelson, D. B. (1998) Engines and nanoparticles: a review. J. Aerosol Sci. 29: 575-588.
Kleiger, R. E.; Miller, J.  P.; Bigger J. T. Jr.; Moss, A.  J.; Multicenter Post-infarction Research Group. (1987)
      Decreased heart rate variability and its association with increased mortality after acute myocardial infarction.
      Am. J.  Cardiol. 59: 256-262.


April 2002                                      9-159         DRAFT-DO NOT QUOTE OR CITE

-------
Kleinman, M. T.; Leaf, D. A.; Kelly, E.; Caiozzo, V.; Osann, K.; O'Niell, T. (1998) Urban angina in the mountains:
      effects of carbon monoxide and mild hypoxemia on subjects with chronic stable angina. Arch. Environ.
      Health 53: 388-397.
Klemm, R. J.; Mason, R. M., Jr. (2000) Aerosol research and inhalation epidemiological study (ARIES): air quality
      and daily mortality statistical modeling-interim results. J. Air. Waste Manage. Assoc. 50: 1433-1439.
Klemm, R. J.; Mason, R. M., Jr.; Heilig, C. M.; Neas, L. M.; Dockery, D. W. (2000) Is daily mortality associated
      specifically with fine particles? Data reconstruction and replication of analyses. J. Air Waste Manage. Assoc.
      50: 1215-1222.
Krewski, D.; Burnett, R. T.; Goldberg, M. S.; Hoover, K.; Siemiatycki, J.; Jerrett, M.; Abrahamowicz, M.; White,
      W. H.. (2000) Reanalysis of the Harvard Six Cities study and the American Cancer Society study of
      paniculate air pollution and mortality. A special report of the Institute's Particle Epidemiology Reanalysis
      Project. Cambridge, MA: Health Effects Institute.
Kuschner, W. G.; Wong, H.; D'Alessandro, A.; Quinlan, P.; Blanc, P. D. (1997) Human pulmonary responses to
      experimental inhalation of high concentration fine and ultrafine magnesium oxide particles. Environ. Health
      Perspect. 105:  1234-1237.
Laden, F.; Neas, L. M.; Dockery, D. W.; Schwartz, J. (2000) Association of fine paniculate matter from different
      sources with daily mortality in six U.S. cities. Environ. Health Perspect. 108: 941-947.
Lawrence, L.; Hall, M. J. (1999) 1997 summary: national hospital discharge survey. Hyattsville, MD: U.S.
      Department  of Health &  Human Resources, Centers for Disease Control and Prevention, National Center for
      Health Statistics. (Advance data from vital and health statistics: no. 308).
Layton, D. W. (1993) Metabolically consistent breathing rates for use in dose assessments. Health Phys. 64: 23-36.
Li, X. Y.; Gilmour, P. S.; Donaldson, K.; MacNee, W.  (1997) In vivo and in vitro proinflammatory effects of
      paniculate air pollution (PM10). In: Driscoll, K. E.; Oberdorster, G., eds. Proceedings of the sixth
      international meeting on the toxicology of natural and man-made fibrous and non-fibrous particles;
      September 1996; Lake Placid, NY. Environ. Health Perspect. Suppl. 105(5): 1279-1283.
Liao, D.; Creason, J.; Shy, C.; Williams, R.; Watts, R.; Zweidinger, R. (1999) Daily variation of paniculate air
      pollution and poor cardiac autonomic control in the elderly. Environ. Health Perspect. 107: 521-525.
Linn, W. S.; Gong, H., Jr.; Shamoo, D. A.; Anderson, K.  R.; Avol, E. L. (1997) Chamber exposures of children to
      mixed ozone, sulfur dioxide, and sulfuric acid. Arch. Environ. Health 52: 179-187.
Linn, W. S.; Szlachcic, Y.; Gong, H., Jr.; Kinney, P. L.; Berhane, K. T. (2000) Air pollution and daily hospital
      admissions in metropolitan Los Angeles. Environ.  Health Perspect. 108: 427-434.
Lipfert, F. W.; Morris, S. C.; Wyzga, R.  E. (2000a) Daily mortality in the Philadelphia metropolitan area and
      size-classified paniculate matter. J. Air Waste Manage. Assoc.: 1501-1513.
Lipfert, F. W.; Zhang, J.; Wyzga, R. E. (2000b) Infant mortality and air pollution: a comprehensive analysis of U.S.
      data for 1990. J. Air Waste Manage. Assoc. 50:  1350-1366.
Lippmann, M.; Ito, K.; Nadas, A.; Burnett, R. T. (2000) Association of paniculate matter components with daily
      mortality and morbidity in urban populations. Cambndge, MA: Health Effects Institute; research report
      no. 95.
Loomis, D.; Castillejos, M.; Gold, D. R.; McDonnell, W.; Borja-Aburto, V. H. (1999) Air pollution and infant
      mortality in Mexico City. Epidemiology 10: 118-123.
Lumley, T.; Heagerty, P. (1999) Weighted empirical adaptive variance estimators for correlated data regression.
      J. R. Stat. Soc. B 61(part 2): 459-477.
Lundgren, D. A.; Burton, R. M. (1995) Effect of particle  size distribution on the cut point between fine and coarse
      ambient mass fractions. In: Phalen, R. F.; Bates, D. V., eds. Proceedings of the colloquium on paniculate air
      pollution and human mortality and morbidity; January 1994; Irvine, CA.  Inhalation Toxicol. 7:  131-148.
Mar, T. F.; Norris,  G. A.; Koenig, J. Q.;  Larson, T. V. (2000) Associations between air pollution and mortality in
      Phoenix, 1995-1997. Environ. Health Perspect. 108:  347-353.
Martin, G. J.; Magid,  N. M.; Myers, G.; Barnett, P. S.;  Schaad, J. W.; Weiss, J.  S.; Lesch, M.;  Singer, D. H. (1987)
      Heart rate variability and sudden death secondary to coronary artery disease  durtng ambulatory
      electrocardiographic monitoring. Am. J.  Cardiol. 60: 86-89.
McConnell, R.; Berhane, K.; Gilliland, F.; London, S. J.; Vora, H.; Avol, E.; Gauderman, W. J.; Margolis, H. G.;
      Lurmann, F.; Thomas, D. C.; Peters, J. M. (1999) Air pollution and bronchitic symptoms in southern
      California children with asthma. Environ. Health Perspect.  107: 757-760.
Michel, O.; Nagy, A.-M.; Schroeven, M.; Duchateau, J.; Neve, J.; Fondu, P.; Sergysels, R. (1997) Dose-response
      relationship  to  inhaled endotoxin in normal subjects. Am. J. Respir. Crit. Care Med. 156: 1157-1164.


April 2002                                     9-160         DRAFT-DO NOT  QUOTE OR CITE

-------
Monn, C.; Becker, S. (1999) Cytotoxicity and induction of proinflammatory cytokines from human monocytes
      exposed to fine (PM2 5) and coarse particles (PM10.2 5) in outdoor and indoor air. Toxicol. Appl. Pharmacol.
      155: 245-252.
Moolgavkar, S. H. (2000a) Air pollution and daily mortality in three U.S. counties. Environ. Health Perspect.
      108: 777-784.
Moolgavkar, S. H. (2000b) Air pollution and hospital admissions for chronic obstructive pulmonary disease in three
      metropolitan areas in the United States. In: Grant, L. D., ed. PM2000: paniculate matter and health.
      Inhalation Toxicol. 12(suppl. 4): 75-90.
Moolgavkar, S. H. (2000c) Air pollution and hospital admissions for diseases of the circulatory system in three U.S.
      metropolitan areas. J. Air Waste Manage Assoc. 50:  1199-1206.
Moolgavkar, S. H.; Luebeck, E. G.; Anderson, E. L. (1997) Air pollution and hospital admissions for respiratory
      causes in Minneapolis-St. Paul and Birmingham. Epidemiology 8: 364-370.
Moolgavkar, S. H.; Hazelton, W.; Luebeck, G.; Levy, D.; Sheppard, L. (2000) Air pollution, pollens, and
      admissions for chronic respiratory  disease in King County, Washington. In: Inhalation Toxicology:
      proceedings of the third colloquium on paniculate air pollution and human health; June, 1999; Durham, NC.
      Inhalation Toxicology 12(suppl. 1): 157-171.
Morawska, L.; Thomas, S.; Bofinger, N.;  Wainwright, D.; Neale, D. (1998) Comprehensive characterization of
      aerosols in a subtropical urban atmosphere: particle size distribution and correlation with gaseous pollutants.
      Atmos. Environ. 32: 2467-2478.
Morris, R. D.; Naumova, E. N. (1998) Carbon monoxide  and hospital admissions for congestive heart failure:
      evidence of an increased effect at low temperatures. Environ. Health Perspect. 106: 649-653.
Morris, R. D.; Naumova, E. N.; Munasinghe, R. L. (1995) Ambient air pollution and hospitalization for congestive
      heart failure among elderly people  in seven large US cities. Am. J. Public Health 85: 1361-1365.
Naeher, L. P.; Holford, T. R.; Beckett, W. S.; Belanger, K.; Triche, E. W.; Bracken,  M. B.; Leaderer, B. P. (1999)
      Healthy women's PEF variations with ambient summer concentrations of PM10, PN2 5, SO42_, H+, and O3. Am.
      J. Respir. Crit. Care Med. 160: 117-125.
Nakamura, T.; Hayashida, Y. (1992) Autonomic cardiovascular responses to smoke  exposure in conscious rats. Am.
      J. Physiol. 262(5 pt. 2): R73 8-745.
National Institutes of Health. (1997) Guidelines for the diagnosis and management of asthma: expert panel report 2.
      Bethesda, MD: U.S. Department of Health and Human Services, National Heart, Lung, and Blood Institute;
      publication no. 97-4051.
Nauenberg, E.; Basu, K.  (1999) Effect of insurance coverage on the relationship between asthma hospitalizations
      and exposure to air pollution. Public Health Rep. 114: 135-148.
Neas, L. M.; Dockery, D. W.; Ware, J. H.; Spengler, J.  D.; Ferris, B. G., Jr.; Speizer, F. E.  (1994) Concentration of
      indoor paniculate  matter as a determinant of respiratory health in children. Am. J. Epidemiol.
      139: 1088-1099.
Neas, L. M.; Dockery, D. W.; Koutrakis, P.; Tollerud, D. J.; Speizer, F. E. (1995) The association of ambient air
      pollution with twice daily peak expiratory flow rate measurements in children. Am. J. Epidemiol.
      141: 111-122.
Neas, L. M.; Dockery, D. W.; Burge, H.; Koutrakis, P.; Speizer, F. E. (1996) Fungus spores, air pollutants, and
      other determinants of peak expiratory flow rate in children. Am. J. Epidemiol. 143: 797-807.
Neas, L. M.; Schwartz, J.; Dockery, D. (1999) A case-crossover analysis of air pollution and mortality in
      Philadelphia.  Environ. Health Perspect. 107: 629-631.
Nikula, K. J.; Vallyathan, V.; Green, F. H. Y.; Hahn, F. F. (2000) Influence of dose  on the  distribution of retained
      paniculate material in rat and human lungs. Presented at: PM2000: paniculate matter and health—the
      scientific basis for regulatory decision-making, specialty conference & exhibition; January; Charleston, SC.
      Pittsburgh, PA: Air & Waste Management Association.
Norris, G.; Young-Pong, S. N.;  Koenig, J. Q.; Larson, T.  V.; Sheppard, L.; Stout, J.  W. (1999) An association
      between fine particles and asthma emergency department visits for children in Seattle.  Environ. Health
      Perspect. 107: 489-493.
Norris, G.; Larson, T.; Koenig, J.; Claiborn, C.; Sheppard, L.; Finn, D. (2000) Asthma aggravation, combustion, and
      stagnant air. Thorax 55: 466-470.
Ohtsuka, Y.; Clarke, R. W.; Mitzner, W.; Branson, K.;  Jakab, G.  J.; Kleeberger, S. R. (2000) Interstrain variation in
      murine susceptibility to inhaled acid-coated particles. Am. J. Physiol. L469-L476.
April 2002                                     9-161         DRAFT-DO NOT QUOTE OR CITE

-------
Ostro, B. D.; Lipsett, M. J.; Wiener, M. B.; Seiner, J. C. (1991) Asthmatic responses to airborne acid aerosols. Am.
      J. Public Health 81: 694-702.
Ostro, B. D.; Lipsett, M. J.; Mann, J. K.; Krupnick, A.; Harrington, W. (1993) Air pollution and respiratory
      morbidity among adults in Southern California. Am. J. Epidemiol.  137: 691-700.
Ostro, B. D.; Lipsett, M. J.; Mann, J. K.; Braxton-Owens, H.; White, M. C. (1995) Air pollution and asthma
      exacerbations among African-American children in Los Angeles. In: Phalen, R. F.; Bates, D. V., eds.
      Proceedings of the colloquium on paniculate air pollution and human mortality and morbidity, part II;
      January 1994; Irvine, CA. Inhalation Toxicol. 7: 711-722.
Ostro, B. D.; Broadwin, R.; Lipsett, M. J. (2000) Coarse and fine particles and daily mortality in the Coachella
      Valley, California: a follow-up study. J. Exposure Anal. Environ. Epidemiol. 10: 412-419.
Ozkaynak, H.; Xue, J.; Zhou, H.; Raizenne, M. (1996) Associations between daily mortality and motor vehicle
      pollution in Toronto, Canada. Boston, MA: Harvard University School of Public Health, Department of
      Environmental Health; March 45.
Pekkanen, J.; Timonen, K. L.; Ruuskanen, J.; Reponen, A.; Mirme, A. (1997) Effects of ultrafine and fine particles
      in urban air on peak expiratory flow among children with asthmatic symptoms. Environ. Res. 74: 24-33.
Perry, K. D.; Cahill, T. A.; Eldred, R. A.; Dutcher, D. D.; Gill, T. E. (1997) Long-range transport of North African
      dust to the eastern United States. J. Geophys. Res. [Atmos] 102: 11,225-11,238.
Peters, A.; Doring, A.; Wichmann, H.-E.; Koenig, W. (1997a) Increased plasma viscosity during an air pollution
      episode: a link to mortality? Lancet 349: 1582-1587.
Peters, A.; Wichmann, H. E.; Tuch, T.; Heinrich, J.; Heyder, J. (1997b) Respiratory effects are associated with the
      number of ultrafine particles. Am.  J. Respir. Crit. Care Med.  155: 1376-1383.
Peters, A.; Dockery, D. W.; Heinrich, J.; Wichmann, H. E. (1997c) Short-term effects of paniculate air pollution on
      respiratory  morbidity in asthmatic children. Eur. Respir. J. 10: 872-879.
Peters, J. M.; Avol, E.; Navidi, W.; London, S. J.; Gauderman, W. J.; Lurmann, F.; Linn, W. S.; Margolis, H.;
      Rappaport, E.; Gong, H., Jr.; Thomas, D. C. (1999a) A study of twelve southern California communities with
      differing levels and types of air pollution. I. Prevalence of respiratory morbidity. Am. J. Respir.  Crit. Care
      Med.  159: 760-767.
Peters, J. M.; Avol, E.; Gauderman, W. J.; Linn, W. S.; Navidi, W.; London,  S. J.; Margolis, H.; Rappaport, E.;
      Vora, H.; Gong, H., Jr.; Thomas, D. C. (1999b) A study of twelve southern California communities with
      differing levels and types of air pollution. II. Effects on pulmonary function. Am. J. Respir. Crit. Care Med.
      159: 768-775.
Plopper, C. G.; Fanucchi, M. V. (2000) Do urban environmental pollutants exacerbate childhood lung diseases?
      Environ. Health Perspect. 108: A252-A253.
Pope, C. A.,  Ill; Thun, M. J.; Namboodiri, M. M.; Dockery, D. W.; Evans, J.  S.; Speizer, F. E.; Heath,  C. W., Jr.
      (1995) Paniculate air pollution as a predictor of mortality in a prospective study of U.S. adults. Am. J.
      Respir. Crit. Care Med. 151: 669-674.
Pope, C. A.,  Ill; Hill, R. W.; Villegas, G.  M. (1999a) Paniculate air pollution and daily mortality on Utah's Wasatch
      Front. Environ. Health Perspect.: 107: 567-573.
Pope, C. A.;  Verrier, R. L.; Lovett, E. G.; Larson, A. C.; Raizenne, M. E.; Kanner, R. E.; Schwartz, J.;  Villegas,
      G. M.; Gold, D. R.; Dockery, D. W. (1999b) Heart rate variability associated with paniculate air pollution.
      Am. Heart J. 138: 890-899.
Raizenne, M.; Neas, L.  M.; Damokosh, A. L; Dockery, D. W.; Spengler, J. D.; Koutrakis, P.; Ware, J. H.; Speizer,
      F. E. (1996) Health effects of acid aerosols on North American children: pulmonary function. Environ.
      Health Perspect.  104: 506-514.
Ramadan, Z.; Song, X.-H.; Hopke, P. K. (2000) Identification of sources of Phoenix aerosol by positive matrix
      factorization. J. Air Waste Manage. Assoc. 50: 1308-1320.
Romieu, I.; Meneses, F.; Ruiz, S.; Sienra, J. J.; Huerta, J.; White, M. C.; Etzel, R. A. (1996) Effects of air pollution
      on the respiratory health of asthmatic children living in Mexico City. Am. J. Respir. Crit. Care Med.
      154: 300-307.
Romieu, I.; Meneses, F.; Ruiz, S.; Huerta, J.; Sienra, J. J.; White, M.; Etzel, R.; Hernandez, M. (1997)  Effects of
      intermittent ozone exposure on peak expiratory flow and respiratory symptoms among asthmatic children in
      Mexico City. Arch. Environ. Health 52: 368-376.
Salvi, S.; Blomberg, A.; Rudell, B.; Kelly, F.; Sandstrom, T.;  Holgate, S. T.; Frew, A.  (1999) Acute inflammatory
      responses in the airways and peripheral blood after short-term exposure to diesel exhaust in healthy human
      volunteers.  Am. J. Respir. Crit. Care Med. 159: 702-709.


April 2002                                     9-162        DRAFT-DO NOT QUOTE OR CITE

-------
Samet, J. M; Dominici, F.; Zeger, S. L.; Schwartz, J.; Dockery, D. W. (2000a) National morbidity, mortality, and
      air pollution study. Part I: methods and methodologic issues. Cambridge, MA: Health Effects Institute;
      research report no. 94.
Samet, J. M.; Zeger, S. L.; Dominici, F.; Curriero, F.; Coursac, L; Dockery, D. W.; Schwartz, J.; Zanobetti, A.
      (2000b) The national morbidity, mortality, and air pollution study. Part II: morbidity, mortality, and air
      pollution in the United States. Cambridge, MA: Health Effects Institute; research report no. 94.
Sarnat, J. A.; Koutrakis, P.; Suh, H. H. (2000) Assessing the relationship between personal paniculate and gaseous
      exposures of senior citizens living in Baltimore, MD. J. Air Waste Manage. Assoc. 50: 1184-1198.
Schlesinger, R. B. (1988) Biological disposition of airborne particles: basic principles and application to vehicular
      emissions. In: Watson, A. Y.; Bates, R. R.; Kennedy, D., eds. Air pollution, the automobile, and public
      health. Washington, DC: National Academy Press; pp. 239-298.
Schlesinger, R. B.; Ben-Jebria, A.; Dahl, A. R.; Snipes, M. B.; Ultman, J. (1997) Disposition of inhaled toxicants.
      In: Massaro, E. J., ed.  Handbook of human toxicology. Boca Raton, FL: CRC Press; pp. 493-550.
Schwartz, J. (1997) Air pollution and hospital admissions for cardiovascular disease in Tucson. Epidemiology
      8: 371-377.
Schwartz, J. (1999) Air pollution and hospital admissions for heart disease in eight U.S. counties. Epidemiology
      10: 17-22.
Schwartz, J. (2000) Harvesting and long term exposure effects in the relation between air pollution and mortality.
      Am. J. Epidemiol.  151: 440-448.
Schwartz, J.; Morris, R. (1995) Air pollution and hospital admissions for cardiovascular disease in Detroit,
      Michigan. Am. J. Epidemiol. 142: 23-35.
Schwartz, J.; Neas, L. M. (2000) Fine particles are more strongly associated than coarse particles with acute
      respiratory health effects in schoolchildren. Epidemiology. 11: 6-10.
Schwartz, J.; Zanobetti, A. (2000) Using meta-smoothing to estimate dose-response trends across multiple studies,
      with application to air pollution and daily death. Epidemiology 11: 666-672.
Schwartz, J.; Dockery, D. W.; Neas, L. M.; Wypij,  D.; Ware, J. H.; Spengler, J. D.; Koutrakis, P.; Speizer, F. E.;
      Ferris, E.G., Jr. (1994) Acute effects of summer air pollution on respiratory symptom reporting in children.
      Am. J. Respir. Crit. Care Med. 150: 1234-1242.
Schwartz, J.; Dockery, D. W.; Neas, L. M. (1996) Is daily mortality associated specifically with fine particles?
      J. Air Waste Manage.  Assoc. 46: 927-939.
Schwartz, J.; Norris, G.; Larson, T.; Sheppard, L.; Claiborne, C.; Koenig, J. (1999) Episodes of high coarse particle
      concentrations are not associated with increased mortality. Environ. Health Perspect. 107: 339-342.
Sheppard, L.; Levy, D.; Norris, G.; Larson, T. V.; Koenig, J. Q. (1999) Effects of ambient air pollution on
      nonelderly asthma hospital admissions in Seattle, Washington, 1987-1994. Epidemiology 10: 23-30.
Singer, D. H.; Martin, G. J.; Magid, N.; Weiss, J. S.; Schaad, J. W.; Kehoe, R.; Zheutiin, T.; Fintel, D. J.; Hsieh,
      A. M.; Lesch, M. (1988) Low heart rate variability and sudden cardiac death.  J. Electrocardiol.
      21(suppl.): S46-S55.
Smiley-Jewell, S. M.; Liu, F. J.; Weir, A. J.; Plopper, C. G. (2000) Acute injury to differentiating Clara cells in
      neonatal rabbits results in age-related failure of bronchiolar epithelial repair. Toxicol. Pathol. 28: 267-276.
Smith, R. L.; Spitzner, D.; Kim, Y.; Fuentes, M. (2000) Threshold dependence of mortality effects for fine and
      coarse particles in Phoenix, Arizona. J. Air Waste Manage.  Assoc. 50: 1367-1379.
Stieb, D. M.; Beveridge, R. C.; Brook, J. R.; Smith-Doiron, M.; Burnett, R. T.; Dales, R. E.; Beaulieu, S.; Judek,  S.;
      Mamedov, A. (2000) Air pollution, aeroallergens and cardiorespiratory emergency department visits in Saint
      John, Canada. J. Exposure Anal. Environ. Epidemiol. 10: 461-477.
Task Force of the European Society of Cardiology and the North American Society of Pacing and
      Electrophysiology. (1996)  Heart rate variability: standards of measurement, physiological interpretation and
      clinical use. Circulation 93: 1043-1065.
Terashima, T.; Wiggs, B.; English, D.; Hogg, J. C.; VanEeden, S.  F. (1997) Phagocytosis of small carbon particles
      (PM10) by alveolar macrophages stimulates the release of polymorphonuclear leukocytes from bone  marrow.
      Am. J. Respir. Crit. Care Med. 155: 1441-1447.
Thurston, G. D.; Ito, K.; Kinney,  P. L.; Lippmann, M. (1992) A multi-year study of air pollution and respiratory
      hospital admissions in three New York State metropolitan areas: results for 1988 and 1989 summers.
      J. Exposure Anal. Environ. Epidemiol. 2: 429-450.
April 2002                                      9-163          DRAFT-DO NOT QUOTE OR CITE

-------
Thurston, G. D.; Ito, K.; Hayes, C. G.; Bates, D. V.; Lippmann, M. (1994) Respiratory hospital admissions and
      summertime haze air pollution in Toronto, Ontario: consideration of the role of acid aerosols. Environ. Res.
      65: 271-290.
Thurston, G. D.; Lippmann, M.; Scott, M. B.; Fine, J. M. (1997) Summertime haze air pollution and children with
      asthma. Am. J. Respir. Crit. Care Med. 155: 654-660.
Tiittanen, P.; Timonen, K. L.; Ruuskanen, J.; Mirme, A.; Pekkanen, J. (1999) Fine paniculate air pollution,
      resuspended road dust and respiratory health among symptomatic children. Eur. Respir. J. 13: 266-273.
Tolbert, P. E.; Klein, M.; Metzger, K. B.; Peel, J.; Flanders, W. D.; Todd, K.; Mulholland, J. A.; Ryan, P. B.;
      Frumkin, H. (2000) Interim results of the study of particulates and health in Atlanta (SOPHIA). J. Exposure
      Anal. Environ. Epidemiol. 10: 446-460.
Tsai, F. C.; Apte, M. G.; Daisey, J. M. (2000) An exploratory analysis of the relationship between mortality and the
      chemical composition of airborne paniculate matter. Inhalation Toxicol. 12(suppl.): 121-135.
Turpin, B. J. (1999) Options for characterizing organic paniculate matter. Environ. Sci. Technol. 33: 76A-79A.
U.S. Environmental Protection Agency. (1996) Air quality criteria for paniculate matter. Research Triangle Park,
      NC: National Center for Environmental Assessment-RTF Office; report nos. EPA/600/P-95/00laF-cF. 3v.
U.S. Environmental Protection Agency. (2000a) Air quality criteria for carbon monoxide. Research Triangle Park,
      NC: National Center for Environmental Assessment; report no. EPA/600/P-99/001F. Available:
      www.epa.gov/ncea/co/ [2000, October 6].
U.S. Environmental Protection Agency. (2000b) Health assessment document for diesel exhaust [external review
      draft]. Research Triangle Park, NC: Office of Health and Environmental Assessment, Environmental Criteria
      and Assessment Office; report nos. EPA/600/8-90/057E.
Vedal, S.; Petkau, J.; White, R.; Blair, J. (1998) Acute effects of ambient inhalable particles in asthmatic and
      nonasthmatic children. Am. J. Respir. Crit. Care Med. 157: 1034-1043.
Ware, J. H.; Ferris, B. G., Jr.; Dockery, D. W.; Spengler, J. D.;  Stram, D. O.; Speizer, F. E.  (1986) Effects of
      ambient sulfur oxides and suspended particles on respiratory health of preadolescent children. Am. Rev.
      Respir. Dis. 133:834-842.
Watkinson, W. P.; Campen, M. J.; Costa, D. L. (1998) Cardiac arrhythmia induction after exposure to residual oil
      fly ash particles in a rodent model of pulmonary hypertension. Toxicol. Sci. 41: 209-216.
Watkinson, W. P.; Campen, M. J.; Dreher, K. L.; Su, W.-Y.; Kodavanti, U.P.; Highfill, J. W.; Costa, D. L. (2000)
      Thermoregulatory effects following exposure to paniculate matter in healthy and
      cardiopulmonary-compromisedrats. J. Therm. Biol. 25: 131-137.
Weschler, C. J.; Shields, H. C. (1999) Indoor ozone/terpene reactions as a source of indoor particles. Atmos.
      Environ. 33:2301-2312.
Whitby, K.  T. (1978) The physical characteristics of sulfur aerosols. Atmos. Environ. 12: 135-159.
Whitby, K.  T.; Sverdrup, G. M. (1980) California aerosols: their physical and chemical characteristics. In: Hidy,
      G. M.; Mueller, P. K.; Grosjean, D.; Appel, B. R.; Wesolowski, J. J., eds. The character and origins of smog
      aerosols: a digest of results from the California Aerosol Characterization Experiment (ACHEX). New York,
      NY:  John Wiley & Sons, Inc.; pp. 477-517. (Advances in environmental science and technology: v. 9).
Wichmann, H.-E.; Spix, C.; Tuch, T.; Wolke, G.; Peters, A.; Heinrich, J.; Kreyling, W. G.; Heyder, J. (2000) Daily
      mortality and fine and ultrafine particles in Erfurt, Germany. Part I: role of particle number and particle mass.
      Cambridge, MA: Health Effects Institute; Research Report no. 98.
Williams, R.; Suggs, J.; Zweidinger, R.; Evans, G.; Creason, J.; Kwok, R.; Rodes,  C.; Lawless, P.; Sheldon, L.
      (2000a) The 1998 Baltimore Paniculate Matter Epidemiology-Exposure Study: part 1. Comparison of
      ambient, residential outdoor, indoor and apartment paniculate matter monitoring. J. Exposure Anal. Environ.
      Epidemiol.  10: 518-532.
Williams, R.; Suggs, J.; Creason, J.;  Rodes, C.; Lawless, P.; Kwok, R.; Zweidinger, R.; Sheldon, L. (2000b)  The
      1998 Baltimore Paniculate  Matter Epidemiology-Exposure Study: part 2. Personal exposure assessment
      associated with an elderly study population. J. Exposure  Anal. Environ. Epidemiol. 10: 533-543.
Williams, R. W.;  Creason, J.; Zweidinger, R.; Watts, R. R.; Shy, C. (2000c) Indoor, outdoor, and personal exposure
      monitoring of paniculate air pollution: the Baltimore elderly epidemiology-exposure  pilot study.
      Atmos. Environ. 34: 4193-4204.
Wilson, W. E.; Suh, H. H. (1997) Fine particles and coarse particles: concentration relationships relevant to
      epidemiologic studies. J. Air Waste Manage. Assoc. 47:  1238-1249.
Woodruff, T. J.; Grille, J.; Schoendorf, K. C. (1997) The relationship between selected causes of postneonatal
      infant mortality and paniculate air pollution in the United States. Environ. Health Perspect. 105: 608-612.


April 2002                                     9-164         DRAFT-DO NOT QUOTE OR CITE

-------
Zanobetti, A.; Schwartz, J. (2000) Race, gender, and social status as modifiers of the effects of PM10 on mortality.
      J. Occup. Environ. Med. 42: 469-474.
Zhang, T.; Huang, C.; Johns, E. J. (1997) Neural regulation of kidney function by the somatosensory system in
      normotensive and hypertensive rats. Am. J. Physiol. 273: R1749-R1757.
Zhang, H.; Triche, E.; Leaderer, B. (2000) Model for the analysis of binary time series of respiratory symptoms.
      Am. J. Epidemiol. 151:  1206-1215.
April 2002                                    9-165        DRAFT-DO NOT QUOTE OR CITE

-------
                          APPENDIX 9A
 Key Quantitative Estimates of Relative Risk for Particulate Matter-Related
 Health Effects Based on Epidemiologic Studies of U.S. and Canadian Cities
   Assessed in the 1996 Particulate Matter Air Quality Criteria Document
April 2002                         9A-1      DRAFT-DO NOT QUOTE OR CITE

-------
         TABLE 9A-1. EFFECT ESTIMATES PER 50-^g/m3 INCREASE
   IN 24-HOUR PM,n CONCENTRATIONS FROM U.S. AND CANADIAN STUDIES
Study Location
RR (±CI)
Only PM
in Model
RR (±CI) Reported
Other Pollutants PM10 Levels
in Model Mean (Min/Max)*
Increased Total Acute Mortality
Six Cities"
Portage, WI
Boston, MA
Topeka, KS
St. Louis, MO
Kingston/Knoxville, TN
Steubenville, OH
St. Louis, MOC
Kingston, TNC
Chicago, ILh
Chicago, ILg
Utah Valley, UTb
Birmingham, ALd
Los Angeles, CAf
Increased Hospital Admissions (for Elderly
Respiratory Disease
Toronto, Canada1
Tacoma, WAJ
New Haven, CTJ
Cleveland, OHk
Spokane, WA1
COPD
Minneapolis, MNn
Birmingham, ALm
Spokane, WA1
Detroit, MI°

1.04 (0.98, 1.09)
1.06(1.04, 1.09)
0.98 (0.90, 1.05)
1.03 (1.00, 1.05)
1.05 (1.00, 1.09)
1.05 (1.00, 1.08)
1.08(1.01, 1.12)
1.09 (0.94, 1.25)
1.04 (1.00, 1.08)
1.03 (1.02, 1.04)
1.08(1.05, 1.11)
1.05(1.01, 1.10)
1.03 (1.00, 1.055)
> 65 years)

1.23 (1.02, 1.43)*
1.10(1.03, 1.17)
1.06(1.00, 1.13)
1.06(1.00, 1.11)
1.08(1.04, 1.14)

1.25(1.10, 1.44)
1.13(1.04, 1.22)
1.17(1.08, 1.27)
1.10(1.02, 1.17)
—
— 18 (±11.7)
— 24 (±12.8)
— 27 (±16.1)
— 31 (±16.2)
— 32 (±14.5)
— 46 (±32.3)
1.06(0.98,1.15) 28(1/97)
1.09 (0.94, 1.26 30 (4/67)
— 37 (4/365)
1.02 (1.01, 1.04) 38 (NR/128)
1.19(0.96,1.47) 47(11/297)
— 48 (21, 80)
1.02 (0.99, 1.036) 58( 15/177)


1.12(0.88,1.36)* 30-39*
1.11(1.02,1.20) 37(14,67)
1.07(1.01,1.14) 41(19,67)
— 43 (19, 72)
— 46 (16, 83)

— 36 (18, 58)
— 45 (19, 77)
— 46 (16, 83)
— 48 (22, 82)
April 2002
9A-2
DRAFT-DO NOT QUOTE OR CITE

-------
          TABLE 9A-1 (cont'd). EFFECT ESTIMATES PER 50-^g/m3 INCREASE
    IN 24-HOUR PMnCONCENTRATIONS FROM U.S. AND CANADIAN STUDIES
Study Location
Pneumonia
Minneapolis, MNn
Birmingham, ALm
Spokane, WA1
Detroit, MI°
Ischemic HP
Detroit, MP
RR (±CI)
OnlyPM
in Model

1.08(1.01, 1.15)
1.09(1.03, 1.15)
1.06(0.98, 1.13)
—

1.02(1.01, 1.03)
RR (±CI) Reported
Other Pollutants PM10 Levels
in Model Mean (Min/Max)1

— 36 (18,58)
— 45 (19, 77)
— 46 (16, 83)
1.06(1.02,1.10) 48(22,82)

1.02 (1.00, 1.03) 48 (22, 82)
Increased Respiratory Symptoms
Lower Respiratory
Six Cities5
Utah Valley, UTr

Utah Valley, UTS
Cough
Denver, COX
Six Cities5
Utah Valley, UTS
Decrease in Lung Function
Utah Valley, UTr
Utah Valley, UTS
Utah Valley, UTW

2.03 (1.36, 3.04)
1.28(1.06, 1.56)T
1.01 (0.81, 1.27)71
1.27 (1.08, 1.49)

1.09(0.57,2.10)
1.51(1.12,2.05)
1.29(1.12, 1.48)

55 (24, 86)**
30 (10, 50)**
29(7,51)***

Similar RR 30(13,53)
— 46(11/195)

— 76(7/251)

— 22 (0.5/73)
Similar RR 30(13,53)
— 76(7/251)

— 46(11/195)
— 76(7/251)
— 55(1,181)
References:

"Schwartz etal.(1996a).
bPopeetal. (1992, 1994)/O3.
'Dockery et al. (1992)/O3.
•"Schwartz (1993).
8Ito and Thurston (1996)/O3.
fKinney et al. (1995)/O3, CO.
hStyeretal. (1995).
'Thurston et al. (1994)/O3.
jSchwartz (1995)/SO2.
"•Schwartz et al. (1996b).
'Schwartz (1996).
"Schwartz (1994a).
"Schwartz (1994b).
"Schwartz (1994c).
"Schwartz and Morris (1995)/O3, CO, SO2.
"Schwartz etal. (1994).
Tope etal. (1991).
Tope and Dockery (1992).
'Schwartz (1994d).
Tope and Kanner (1993).
 "Ostro etal. (1991).
 fMin/Max 24-h PM10 in parentheses unless noted
  otherwise as standard deviation (±SD), 10 and
  90 percentile (10, 90). NR = not reported.
 "Children.
 "Asthmatic children and adults.
 'Means of several cities.
 "PEFR decrease in mlVs.
 "FEV! decrease.
 *RR refers to total population, not just >65 years.
April 2002
                       9A-3
DRAFT-DO NOT QUOTE OR CITE

-------
  TABLE 9A-2. EFFECT ESTIMATES PER VARIABLE INCREMENTS IN 24-HOUR
       CONCENTRATIONS OF FINE PARTICLE INDICATORS (PM2 5, SO^, H+)
                         FROM U.S. AND CANADIAN STUDIES
Acute Mortality
Six City3
Portage, WI
Topeka, KS
Boston, MA
St. Louis, MO
Kingston/Knoxville, TN
Steubenville, OH
Increased Hospitalization
Ontario, Canadab
Ontario, Canada0
NYC/Buffalo, NYd
Torontod I
Increased Respiratory Symptoms
Southern Californiaf
Six Cities8
(Cough)
Six Cities8
(Lower Resp. Symp.)
Decreased Lung Function
Uniontown, PAe
Indicator

PM25
PM25
PM25
PM25
PM25
PM,,

so;
so;
03
so;
f (Nmol/m3)
so;
PM,,

so;
PM25
PM2 5 Sulfur
H+
PM25
PM2 5 Sulfur
H+

PM25
RR (±CI) per 25 i^g/m3
PM Increase

1.030(0.993, 1.071)
1.020(0.951, 1.092)
1.056(1.038, 1.0711)
1.028(1.010, 1.043)
1.035 (1.005, 1.066)
1.025 (0.998, 1.053)

1.03 (1.02, 1.04)
1.03 (1.02, 1.04)
1.03 (1.02, 1.05)
1.05(1.01, 1.10)
1.16(1.03, 1.30)*
1.12(1.00, 1.24)
1.15 (1.02, 1.78)

1.48(1.14, 1.91)
.19(1.01, 1.42)"
.23 (0.95, 1.59)"
.06 (0.87, 1.29)"
.44(1.15-1.82)**
.82 (1.28-2.59)**
.05 (0.25-1.30)**

PEFR 23.1 (-0.3, 36.9) (per 25 ,wg/m3)
Reported PM
Levels Mean
(Mm/Max)1

11.2 (±7.8)
12.2 (±7.4)
15.7 (±9.2)
18.7 (±10.5)
20.8 (±9.6)
29.6 (±21.9)

R = 3. 1-8.2
R = 2.0-7.7
NR
28.8 (NR/391)
7.6 (NR, 48.7)
18.6 (NR, 66.0)

R = 2-37
18.0 (7.2, 37)"*
2.5(3.1,61)*"
18.1 (0.8,5.9)*"
18.0 (7.2, 37)"*
2.5 (0.8, 5.9)"*
18.1(3.1,61)"*

25/88 (NR/88)
References:
"Schwartz etal. (1996a).
bBurnettetal. (1994).
cBurnettetal. (1995) O3.
dThurston et al. (1992, 1994).
dNeas et al.  (1995).
fOstro etal.  (1993).
BSchwartzetal. (1994).
fMin/Max 24-h PM indicator level shown in parentheses unless
otherwise noted as (±SD), 10 and 90 percentile (10,90) or
R = range of values from min-max, no mean value reported.
'Change per 100 nmoles/m3
"Change per 20 ^g/m3 for PM2 5; per 5 ^g/m3 for PM2 5 sulfur;
 per 25 nmoles/m3 for H+.
***50th percentile value (10,90 percentile).
April 2002
   9A-4
DRAFT-DO NOT QUOTE OR CITE

-------
             TABLE 9A-3. EFFECT ESTIMATES PER INCREMENTS3 IN
         ANNUAL MEAN LEVELS OF FINE PARTICLE INDICATORS FROM
                             U.S. AND CANADIAN STUDIES
Type of Health
Effect and Location
Increased Total Chronic
Six Cityb


ACS Study0
(151 U.S. SMSA)

Indicator
Mortality in Adults
PM15/10
PM25
so;
PM25
so:
Increased Bronchitis in Children
Six Cityd
Six City6
24 Cityf
24 Cityf
24 Cityf
24 Cityf
Southern California8
PM15/10
TSP
H+
so;
PM21
PM10
so:
Change in Health Indicator per
Increment in PMa
Relative Risk (95% CI)
1.42(1.16-2.01)
1.31 (1.11-1.68)
1.46(1.16-2.16)
1.17(1.09-1.26)
1.10(1.06-1.16)
Odds Ratio (95% CI)
3.26(1.13, 10.28)
2.80(1.17,7.03)
2.65 (1.22, 5.74)
3.02(1.28,7.03)
1.97(0.85,4.51)
3.29(0.81, 13.62)
1.39 (0.99, 1.92)
Range of City
PM Levels
Means (,wg/m3)

18-47
11-30
5-13
9-34
4-24

20-59
39-114
6.2-41.0
18.1-67.3
9.1-17.3
22.0-28.6
—
Decreased Lung Function in Children
Six City4h
Six City6
24 Citylj
24 City1
24 City1
24 City1
PM15/10
TSP
H+ (52 nmoles/m3)
PM2 ! (15 Mg/m3)
SOI (7 Mg/m3)
PM10(17Mg/m3)
NS Changes
NS Changes
-3. 45% (-4.87, -2.01)FVC
-3.21% (-4.98, -1.41) FVC
-3.06% (-4.50, -1.60) FVC
-2.42% (-4.30, -.0.51) FVC
20-59
39-114
—
—
—
—
 "Estimates calculated annual-average PM increments assume: a 100-^g/m3 increase for TSP; a 50-^g/m3
  increase for PM10 and PM15; a 25-^g/m3 increase for PM2 5; and a 15-^g/m3 increase for SO:, except where
  noted otherwise; a 100-nmole/m3 increase forH+.
 bDockeryetal. (1993).
 Tope etal. (1995).
 dDockery et al. (1989).
 6Wareetal. (1986).
 TJockery et al. (1996).
 BAbbey et al. (1995).
 hNS Changes = No significant changes.
 'Raizenne et al. (1996).
 JPollutant data same as for Dockery et al. (1996).
April 2002
9A-5
DRAFT-DO NOT QUOTE OR CITE

-------
REFERENCES

Abbey, D. E.; Mills, P. K.; Petersen, F. F.; Beeson, W. L. (1991) Long-term ambient concentrations of total
      suspended particulates and oxidants as related to incidence of chronic disease in California Seventh-Day
      Adventists. Environ. Health Perspect. 94: 43-50.
Abbey, D. E.; Lebowitz, M. D.; Mills, P. K.; Petersen, F. F.; Beeson, W. L.; Burchette, R. J. (1995a) Long-term
      ambient concentrations of particulates and oxidants and development of chronic disease in a cohort of
      nonsmoking California residents. In: Phalen, R. F.; Bates, D. V., eds. Proceedings of the colloquium on
      paniculate air pollution and human mortality and morbidity; January 1994; Irvine, CA. Inhalation Toxicol.
      7: 19-34.
Abbey, D. E.; Ostro, B. E.; Fraser, G.; Vancuren, T.;  Burchette,  R. J. (1995b) Estimating fine particulates less that
      2.5 microns in aerodynamic diameter (PM2.5) from airport visibility data in California. J. Exposure Anal.
      Environ. Epidemiol. 5: 161-180.
Abbey, D. E.; Ostro, B. E.; Petersen, F.; Burchette, R. J. (1995c) Chronic respiratory symptoms associated with
      estimated long-term ambient concentrations of fine particulates less than 2.5 microns in aerodynamic
      diameter (PM2 5) and other air pollutants. J. Exposure Anal. Environ. Epidemiol. 5: 137-159.
Abbey, D. E.; Burchette, R. J.; Knutsen, S. F.; McDonnell, W. F.; Lebowitz, M. D.; Enright, P. L. (1998) Long-term
      paniculate and other air pollutants and lung function in nonsmokers. Am. J. Respir. Crit. Care Med.
      158: 289-298.
Abbey, D. E.; Nishino, N.; McDonnell, W. F.; Burchette, R. J.; Knutsen, S. F.; Beeson, W. L.; Yang, J. X.  (1999)
      Long-term inhalable particles and other air pollutants related to mortality  in nonsmokers. Am. J. Respir. Crit.
      Care Med.  159: 373-382.
Adams, P. F.; Hendershot, G. E.; Marano, M. A. (1999) Current estimates from the National Health Interview
      Survey, 1996. Vital Health Stat. 10(200).
American Conference of Governmental Industrial Hygienists.  (1994) Appendix D: particle size-selective sampling
      criteria for airborne paniculate matter. In: 1994-1995 Threshold limit values for chemical substances and
      physical agents and biological exposure indices.  Cincinnati, OH: American Conference of Governmental
      Industrial Hygienists; pp. 43-46.
Bates, D. V. (2000) Lines  that connect: assessing the causality inference in the case of paniculate pollution.
      Environ. Health Perspect. 108: 91-92.
Becker, S.; Soukup, J. M.; Gilmour, M. I.; Devlin, R. B. (1996)  Stimulation of human and rat alveolar macrophages
      by urban air particulates: effects on oxidant radical generation and cytokine production. Toxicol. Appl.
      Pharmacol. 141:637-648.
Berglund, D. J.; Abbey, D. E.; Lebowitz, M. D.; Knutsen, S. F.;  McDonnell, W. F. (1999) Respiratory symptoms
      and pulmonary function in an elderly nonsmoking population. Chest 115: 49-59.
Binkova, B.; Vesely, D.; Vesela, D.; Jelinek, R.; Sram, R. J. (1999) Genetoxicity and embryotoxicity of urban air
      paniculate matter collected during winter and summer period in two different districts of the Czech Republic.
      Mutat. Res. 440: 45-58.
Bobak, M.; Leon, D. A. (1999) Pregnancy outcomes and outdoor air pollution: an ecological study in districts of the
      Czech Republic 1986-8. Occup. Environ. Med. 56: 539-543.
Brook, R. D.; Brook, J. R.; Urch, B.; Vincent, R.; Rajagopalan, S.; Silverman, F. (2002) Inhalation of fine
      paniculate air pollution and ozone causes acute arterial vasoconstriction in health adults. Circulation 105:
      1534-1536.
Brunekreef, B. (1997) Air pollution and life expectancy: is there a relation? Occup. Environ. Med. 54:  781-784.
Burnett, R. T.; Dales, R. E.; Raizenne, M. E.; Krewski, D.; Summers, P. W.; Roberts, G. R.; Raad-Young, M.;
      Dann, T.; Brook, J. (1994) Effects of low ambient levels of ozone and sulfates on the frequency of
      respiratory admissions to Ontario hospitals. Environ. Res. 65: 172-194.
Burnett, R. T.; Dales, R.; Krewski, D.; Vincent, R.; Dann, T.;  Brook, J. R. (1995) Associations between ambient
      paniculate sulfate and admissions to Ontario hospitals for cardiac and respiratory diseases.  Am. J. Epidemiol.
      142: 15-22.
Burnett, R. T.; Cakmak, S.; Brook, J. R.; Krewski, D. (1997) The role of paniculate size and chemistry in the
      association between summertime ambient air pollution and hospitalization for cardiorespiratory diseases.
      Environ. Health Perspect. 105: 614-620.
April 2002                                      9A-6          DRAFT-DO NOT QUOTE OR CITE

-------
Burnett, R. T.; Cakmak, S.; Raizenne, M. E.; Stieb, D.; Vincent, R.; Krewski, D.; Brook, J. R.; Philips, O.;
      Ozkaynak, H. (1998) The association between ambient carbon monoxide levels and daily mortality in
      Toronto, Canada. J. Air Waste Manage. Assoc. 48: 689-700.
Burnett, R. T.; Smith-Doiron, M.; Stieb, D.; Cakmak, S.; Brook, J. R. (1999) Effects of paniculate and gaseous air
      pollution on cardiorespiratory hospitalizations. Arch. Environ. Health 54: 130-139.
Burnett, R. T.; Brook, J.; Dann, T.; Delocla, C.; Philips, O.; Cakmak, S.; Vincent, R.; Goldberg, M. S.; Krewski, D.
      (2000) Association between paniculate- and gas-phase components of urban air pollution and daily mortality
      in eight Canadian cities. In: Grant, L. D., ed. PM2000: paniculate matter and health. Inhalation Toxicol.
      12(suppl. 4): 15-39.
Calvert, J. G.; Heywood, J. B.; Sawyer, R. F.; Seinfeld, J. H. (1993)  Achieving acceptable air quality:  some
      reflections on controlling vehicle emissions. Science (Washington, DC) 261: 37-45.
Castillejos, M.; Borja-Aburto, V. H.; Dockery, D.  W.; Gold, D. R.; Loomis, D. (2000) Airborne coarse particles and
      mortality. In: Phalen, R. F., ed. Inhalation Toxicology: proceedings of the third colloquium on paniculate air
      pollution and human health (first special issue); June, 1999; Durham, NC. Inhalation Toxicol.
      12(suppl. 1): 61-72.
Chapman, R.  S.; Watkinson, W. P.; Dreher, K. L.;  Costa, D. L. (1997) Ambient paniculate matter and respiratory
      and cardiovascular illness in adults: particle-borne transition metals and the heart-lung axis. Environ.
      Toxicol. Pharmacol. 4: 331-338.
Chen, L.; Yang, W.; Jennison, B. L.; Omaye, S. T. (2000) Air paniculate pollution and hospital admissions for
      chronic obstructive pulmonary disease in Reno, Nevada. Inhalation Toxicol. 12: 281-298.
Chock, D. P.; Winkler, S.; Chen, C. (2000) A study of the association between daily mortality and ambient air
      pollutant concentrations in Pittsburgh, Pennsylvania. J. Air Waste Manage. Assoc. 50: 1481-1500.
Chor^zy, M.;  Szeliga, J.; Strozyk, M.; Cimander, B. (1994) Ambient air pollutants in Upper Silesia: partial chemical
      composition and biological activity. In: Symposium of risk assessment of urban air: emissions, exposure, risk
      identification,  and risk quantitation; May-June 1992; Stockholm, Sweden. Environ. Health Perspect.
      102(suppl4):61-66.
Choudhury, A. H.;  Gordian, M. E.; Morris, S. S. (1997) Associations between respiratory illness and PM10 air
      pollution. Arch. Environ. Health 52: 113-117.
Cifuentes, L.  A.; Vega, J.; Kopfer, K.; Lave, L. B.  (2000) Effect of the fine fraction of paniculate matter versus the
      coarse  mass  and other pollutants on daily mortality in Santiago, Chile. J. Air Waste Manage. Assoc.
      50:  1287-1298.
Claiborn, C. S.; Finn, D.; Larson, T. V.; Koenig, J. Q. (2000) Windblown dust contributes to  high PM25
      concentrations. J. Air Waste Manage. Assoc. 50:  1440-1445.
Clarke, R. W.; Antonini, J. M.; Hemenway, D. R.;  Frank, R.; Kleeberger, S. R.; Jakab, G. J. (2000) Inhaled
      particle-bound sulfate: effects on pulmonary inflammatory responses and alveolar macrophage function.
      Inhalation Toxicol. 12: 169-186.
Clyde, M.  A.; Guttorp, P.; Sullivan, E. (2000) Effects of ambient fine and coarse particles on mortality in Phoenix,
      Arizona. J. Exposure Anal. Environ. Epidemiol.:  submitted.
Code of Federal Regulations. (200 la) Appendix J to Part 50~reference method for the determination of paniculate
      matter  as PM10 in the atmosphere. C. F. R.  40: §50.
Code of Federal Regulations. (200Ib) Appendix L to Part 50~reference method for the determination  of fine
      paniculate matter as PM2.5 in the atmosphere. C. F. R. 40: §50.
Costa, D. L.; Dreher,  K. L. (1999) What do we need to know about airborne particles to make effective risk
      management decisions? A toxicology perspective. Hum. Ecol. Risk Assess. 5: 481-491.
Daniels, M.; Dominici, F.; Samet, J. M.; Zeger, S.  L. (2000) Estimating paniculate matter-mortality dose-response
      curves  and threshold levels: an analysis of daily time-series for the 20 largest US cities. Am. J. Epidemiol.
      152: 397-406.
Dejmek, J.; Selevan, S. G.; Benes, I.; Solansky, I.;  Sram, R. J. (1999) Fetal growth and maternal exposure to
      paniculate matter during pregnancy. Environ. Health Perspect. 107: 475-480.
Delfino, R. J.; Murphy-Moulton, A. M.; Burnett, R. T.; Brook, J. R.; Becklake, M. R. (1997)  Effects of air pollution
      on emergency room visits for respiratory illnesses in Montreal, Quebec. Am. J. Respir. Crit. Care Med.
      155: 568-576.
Delfino, R. J.; Zeiger, R. S.; Seltzer, J. M.; Street, D. H. (1998) Symptoms inpediatric asthmatics and air pollution:
      differences in effects by symptom severity, anti-inflammatory medication use and paniculate averaging time.
      Environ. Health Perspect. 106: 751-761.


April 2002                                      9A-7         DRAFT-DO NOT QUOTE OR CITE

-------
Dockery, D. W.; Spengler, J. D. (1981) Personal exposure to respirable particulates and sulfates. J. AirPollut.
      Control Assoc. 31: 153-159.
Dockery, D. W.; Speizer, F. E.; Stram, D. O.; Ware, J. H.; Spengler, J. D.; Ferris, B. G., Jr. (1989) Effects of
      inhalable particles on respiratory health of children. Am. Rev. Respir. Dis. 139: 587-594.
Dockery, D. W.; Pope, C. A., Ill; Xu, X.; Spengler, J. D.; Ware, J. H.; Fay, M. E.; Ferris, B. G., Jr.; Speizer, F. E.
      (1993) An association between air pollution and mortality in six U.S. cities. N. Engl. J. Med.
      329: 1753-1759.
Dockery, D. W.; Cunningham, J.; Damokosh, A. I.; Neas, L. M.; Spengler, J. D.; Koutrakis, P.; Ware, J. H.;
      Raizenne, M.; Speizer, F. E. (1996) Health effects of acid aerosols on North American children: respiratory
      symptoms. Environ. Health Perspect. 104: 500-505.
Dockery, D. W.; Pope, C. A., Ill; Kanner, R. E.; Villegas, G. M.; Schwartz, J. (1999) Daily changes in oxygen
      saturation and pulse rate associated with paniculate air pollution and barometric pressure. Cambridge, MA:
      Health Effects Institute; research report no. 83.
Dong, W.; Lewtas, J.; Luster, M. I. (1996) Role of endotoxin in tumor necrosis factor "alpha" expression from
      alveolar macrophages treated with urban air particles. Exp. Lung Res. 22: 577-592.
Dreher, K. L. (2000) Paniculate matter physicochemistry and toxicology: in search of causality—a critical
      perspective. Inhalation Toxicol. 12(suppl. 3): 45-57.
Driscoll, K. E.; Costa, D. L.; Hatch, G.; Henderson, R.; Oberdorster, G.; Salem, H.; Schlesinger, R. B. (2000)
      Intratracheal instillation as an exposure technique for the evaluation of respiratory  tract toxicity: uses and
      limitations. Toxicol. Sci. 55: 24-35.
Dye, J. A.; Lehmann, J. R.; McGee, J. K.; Winsett, D. W.; Ledbetter, A. D.; Everitt, J. I.;  Ohio, A. J.; Costa, D. L.
      (2001) Acute pulmonary toxicity of paniculate matter filter extracts in rats: coherence with epidemiological
      studies in Utah Valley residents. Environ. Health Perspect. 109(suppl. 3): 395-403.
Ebelt, S. T.; Petkau, A. J.; Vedal, S.; Fisher, T. V.; Brauer, M.  (2000) Exposure of chronic obstructive pulmonary
      disease patients to paniculate matter: relationships between personal and ambient air concentrations. J. Air
      Waste Manage. Assoc. 50: 1081-1094.
Elder, A. C. P.; Gelein, R.; Finkelstein J. N.; Cox, C.; Oberdorster, G. (2000) Endotoxin priming affects the lung
      response to ultrafine particles and ozone in young and old rats. In: Inhalation Toxicology: proceedings of the
      third colloquium on paniculate air pollution and human health; June, 1999; Durham, NC. Inhalation
      Toxicology 12(suppl. 1): 85-98.
Fairley, D. (1999) Daily  mortality and air pollution in Santa Clara County, California: 1989-1996. Environ. Health
      Perspect. 107: 637-641.
Fanucchi, M. V.; Wong, V. J.; Hinds, D.; Tarkington, B. K.; Van Winkle, L. S.; Evans, M. J.; Plopper, C. G. (2000)
      Repeated episodes of exposure to ozone alters postnatal development of distal conducting airways in infant
      rhesus monkeys. Am. J. Respir. Crit. Care Med. 161: A615.
Federal Register. (1987) Revisions to the national ambient air quality standards for paniculate matter. F. R. (July 1)
      52: 24,634-24,669.
Federal Register. (1997) National ambient air quality standards for paniculate matter; final rule. F. R. (July 18)
      62: 38,652-38,752.
Fitz-Simons, T. S.; Mathias, M.; Rizzo, M. (2000) Analyses of 1999 PM data for the PM NAAQS review.
      U.S. Environmental Protection Agency, Office of Air Quality Planning and Standards; September 29.
Fortoul, T. I.; Osorio, L. S.; Tovar, A. T.; Salazar, D.; Castilla, M. E.; Olaiz-Fernandez, G. (1996)  Metals in lung
      tissue from autopsy cases in Mexico City residents: comparison of cases from the 1950s and the 1980s.
      Environ. Health Perspect. 104: 630-632.
Frampton, M. W.; Ohio, A. J.; Samet, J. M.; Carson, J. L.; Carter, J. D.; Devlin, R. B. (1999) Effects of aqueous
      extracts of PM10 filters from the Utah Valley on human airway epithelial cells. Am. J. Physiol. 277:
      L960-L967.
Gamble, J. L. (1998) Effects of ambient air pollution on daily mortality: a time series analysis of Dallas, Texas,
      1990-1994. Presented at: 91st annual meeting and exhibition of the Air & Waste Management Association;
      June; San Diego, CA. Pittsburgh, PA: Air & Waste Management Association; paper no. 98-MP26.03.
Gauderman, W. J.; Mcconnell, R.; Gilliland, F.; London, S.; Thomas, D.; Avol, E.; Vora, H.; Berhane, K.;
      Rappaport, E. B.;  Lurmann, F.; Margolis, H. G.; Peters, J. (2000) Association between air pollution and lung
      function growth in southern California children. Am. J. Respir. Crit. Care Med. 162: 1383-1390.
Ohio, A. J.; Devlin, R. B. (2001) Inflammatory lung injury after bronchial instillation of air pollution particles.
      Am. J. Respir. Crit. Care Med. 164: 704-708.


April 2002                                     9A-8         DRAFT-DO NOT QUOTE OR CITE

-------
Ohio, A. J.; Kim, C.; Devlin, R. B. (2000) Concentrated ambient air particles induce mild pulmonary inflammation
      in healthy human volunteers. Am. J. Respir. Crit. Care Med. 162: 981-988.
Ohio, A. J.; Gilbey, J. G.; Roggli, V. L.; Richards, J. H.; McGee, J. K.; Carson, J. L.; Devlin, R. B.; Cascio, W. E.
      (2001) Diffuse alveolar damage after exposure to an oil fly ash. Am. J. Respir. Crit. Care Med. 164:
      1514-1518.
Gielen, M. H.; VanDer Zee, S. C.; Van Wijnen, J. H.; Van Steen, C. J.; Brunekreef, B. (1997) Acute effects of
      summer air pollution on respiratory health of asthmatic children. Am. J. Respir. Crit. Care Med.
      155:2105-2108.
Godleski, J. J.; Sioutas, C.; Katler, M.; Koutrakis, P. (1996) Death from inhalation of concentrated ambient air
      particles in animal models of pulmonary disease. Am. J. Respir. Crit. Care Med. 153: A15.
Gold, D. R.; Damokosh, A. L; Pope, C. A., Ill; Dockery, D. W.; McDonnell, W. F.; Serrano, P.; Retama, A.;
      Castillejos, M. (1999) Paniculate and ozone pollutant effects on the respiratory function of children in
      southwest Mexico City. Epidemiology 10: 8-16.
Gold, D. R.; Litonjua, A.; Schwartz, J.; Lovett, E.; Larson, A.; Nearing, B.; Allen, G.;  Verrier, M.; Cherry, R.;
      Verrier, R.  (2000) Ambient pollution and heart rate variability. Circulation 101: 1267-1273.
Goldberg, M. S.; Bailar, J. C., Ill; Burnett, R. T.; Brook, J. R.; Tamblyn, R.; Bonvalot, Y.; Ernst, P.; Flegel, K. M.;
      Singh,  R. K.; Valois, M.-F. (2000) Identifying subgroups of the general population that may be susceptible to
      short-term increases in paniculate air pollution: a time-series study in Montreal, Quebec. Cambridge, MA:
      Health Effects Institute; research report 97. Available: http://www.healtheffects.org/pubs-research.htm
      [15 February, 2001].
Guenther, A.; Geron, C.; Pierce, T.; Lamb, B.; Harley, P.; Fall, R. (2000) Natural emissions of non-methane volatile
      organic compounds, carbon monoxide, and oxides of nitrogen from North America. Atmos. Environ.
      34: 2205-2230.
Gwynn, R. C.; Burnett, R. T.; Thurston, G. D. (2000) A time-series analysis of acidic paniculate matter and daily
      mortality and morbidity in the Buffalo, New York, region. Environ. Health Perspect. 108: 125-133.
Henry, R. C. (1997) History and fundamentals of multivariate air quality receptor models. Chemom. Intell. Lab.
      Syst. 37: 37-42.
Hofmann, W.; Balashazy, I.; Heistracher, T.; Koblinger, L. (1996) The significance of particle deposition patterns in
      bronchial airway bifurcations for extrapolation modeling. Aerosol Sci. Technol. 25: 305-327.
Hoyert, D. L.; Kochanek, K. D.; Murphy, S. L. (1999) Deaths: final data for 1997. Atlanta, GA: U.S. Centers for
      Disease Control and Prevention, National Center for Health Statistics. (National vital statistics reports: v. 47,
      no. 19).
Hsieh, T. H.; Yu, C. P. (1998) Two-phase pulmonary clearance of insoluble particles in mammalian species.
      Inhalation Toxicol. 10:  121-130.
Imrich, A.; Ning, Y. Y.; Koziel, H.; Coull, B.; Kobzik, L. (1999) Lipopolysaccharide priming amplifies lung
      macrophage tumor necrosis factor production in response to air particles. Toxicol. Appl. Pharmacol.
      159:  117-124.
Izzotti, A.; Camoirano, A.; DAgostini, F.; Sciacca, S.; De Naro Papa, F.; Cesarone, C. F.; De Flora, S. (1996)
      Biomarker alterations produced in rat lung by intratracheal instillations of air paniculate  extracts and
      chemoprevention with oral N-acetyIcysteine. Cancer Res. 56: 1533-1538.
Jakab, G. J.; Clarke, R. W.; Hemenway, D. R.; Longphre, M. V.; Kleeberger, S. R.; Frank, R. (1996) Inhalation of
      acid coated carbon black particles impairs alveolar macrophage phagocytosis. Toxicol. Lett. 88: 243-248.
Janssen, N. A. H.; Schwartz, J.; Zanobetti, A.; Suh, H. H. (2002) Air conditioning and source-specific particles as
      modifiers of the effect of PM10 on hospital admissions for heart and lung disease. Environ. Health Perspect.
      110:43-49.
Klemm, R. J.; Mason, R. M., Jr. (2000) Aerosol research and inhalation epidemiological study (ARIES): air quality
      and daily mortality statistical modeling-interim results. J. Air. Waste Manage. Assoc. 50: 1433-1439.
Klemm, R. J.; Mason, R. M., Jr.; Heilig, C. M.; Neas, L. M.; Dockery, D. W. (2000) Is daily mortality associated
      specifically with fine particles? Data reconstruction and replication of analyses. J. Air Waste Manage. Assoc.
      50: 1215-1222.
Kodavanti, U. P.; Schladweiler, M. C.; Ledbetter, A. D.; Watkinson, W. P.; Campen, M. J.; Winsett,  D. W.;
      Richards, J. R.; Crissman, K. M.; Hatch, G. E.; Costa, D.  L. (2000) The spontaneously hypertensive rat as a
      model of human cardiovascular disease: evidence of exacerbated cardiopulmonary injury and oxidative stress
      from inhaled emission paniculate matter. Toxicol. Appl. Pharmacol. 164: 250-263.
April 2002                                     9A-9         DRAFT-DO NOT QUOTE OR CITE

-------
Krewski, D.; Burnett, R. T.; Goldberg, M. S.; Hoover, K.; Siemiatycki, J.; Jerrett, M.; Abrahamowicz, M; White,
      W. H.. (2000) Reanalysis of the Harvard Six Cities study and the American Cancer Society study of
      paniculate air pollution and mortality. A special report of the Institute's Particle Epidemiology Reanalysis
      Project. Cambridge, MA: Health Effects Institute.
Ktinzli, N.; Tager, I. B. (1997) The semi-individual study in air pollution epidemiology: a valid design as compared
      to ecologic studies. Environ. HealthPerspect. 105: 1078-1083.
Ktinzli, N.; Medina, S.; Kaiser, R.; Quenel, P.; Horak, F., Jr.; Studnicka, M. (2001) Assessment of deaths
      attributable to air pollution: should we use risk estimates based on time series or on cohort studies? Am. J.
      Epidemiol. 153:  1050-1055.
Kuschner, W. G.; Wong, H.; D'Alessandro, A.; Quinlan, P.; Blanc, P. D. (1997) Human pulmonary responses to
      experimental inhalation of high concentration fine and ultrafine magnesium oxide particles. Environ. Health
      Perspect. 105:  1234-1237.
Laden, F.; Neas, L. M.; Dockery, D. W.; Schwartz, J. (2000) Association of fine paniculate matter from different
      sources with daily mortality in six U.S. cities. Environ. Health Perspect. 108: 941-947.
Lambert, A. L.; Dong, W.; Selgrade, M. K.; Gilmour, M. I. (2000) Enhanced allergic sensitization by residual oil fly
      ash particles is mediated by soluble metal constituents. Toxicol. Appl. Pharmacol. 165:  84-93.
Landis, M. S.; Norris, G. A.; Williams, R. W.; Weinstein, J. P. (2001) Personal exposures to PM2 5 mass and trace
      elements in Baltimore, MD, USA. Atmos. Environ. 35: 6511-6524.
Lawrence, L.; Hall, M. J. (1999) 1997 summary: national hospital discharge survey. Hyattsville, MD: U.S.
      Department of Health & Human Resources, Centers for Disease Control and Prevention, National Center for
      Health Statistics. (Advance data from vital and health statistics: no. 308).
Layton, D. W. (1993) Metabolically consistent breathing rates for use in dose assessments. Health Phys. 64: 23-36.
Li, X. Y.; Gilmour, P. S.; Donaldson, K.; MacNee, W.  (1997) In vivo and in vitro proinflammatory effects of
      paniculate air pollution (PM10). In: Driscoll, K. E.; Oberdorster, G., eds. Proceedings of the sixth
      international meeting on the toxicology of natural and man-made fibrous and non-fibrous particles;
      September 1996; Lake Placid, NY. Environ. Health Perspect. Suppl. 105(5): 1279-1283.
Li, X. Y.; Brown, D.; Smith, S.; MacNee, W.; Donaldson, K. (1999) Short-term inflammatory responses following
      intratracheal instillation of fine and ultrafine carbon black in rats. Inhalation Toxicol. 11: 709-731.
Liao, D.; Creason, J.; Shy, C.; Williams, R.; Watts, R.; Zweidinger, R. (1999) Daily variation of paniculate air
      pollution and poor cardiac autonomic control in the elderly. Environ. Health Perspect. 107: 521-525.
Linn, W. S.; Gong, H., Jr.; Shamoo, D. A.; Anderson, K. R.; Avol, E. L. (1997) Chamber exposures of children to
      mixed ozone, sulfur dioxide, and sulfuric acid. Arch. Environ. Health 52: 179-187.
Linn, W. S.; Szlachcic, Y.; Gong, H., Jr.; Kinney, P. L.; Berhane, K. T. (2000) Air pollution and daily hospital
      admissions in metropolitan Los Angeles. Environ. Health Perspect. 108: 427-434.
Lioy, P. J. (1990) Assessing total human exposure to contaminants: a multidisciplinary approach. Environ. Sci.
      Technol. 24: 938-945.
Lipfert, F. W.; Morris, S. C.; Wyzga, R. E. (2000a) Daily mortality in the Philadelphia metropolitan area and
      size-classified paniculate matter. J. Air Waste Manage. Assoc.: 1501-1513.
Lipfert, F. W.; Zhang, J.; Wyzga, R. E. (2000b) Infant mortality and air pollution: a comprehensive analysis of U.S.
      data for 1990. J.  Air Waste Manage. Assoc. 50:  1350-1366.
Lippmann, M.; Ito, K.; Nadas, A.; Burnett, R. T. (2000) Association of paniculate matter components with daily
      mortality and morbidity in urban populations. Cambridge, MA: Health Effects Institute; research report
      no. 95.
Lison, D.; Lardot, C.; Huaux, F.; Zanetti, G.; Fubini, B. (1997) Influence of particle surface area on the toxicity of
      insoluble manganese dioxide dusts. Arch. Toxicol. 71: 725-729.
Long, C. M.; Suh, H.  H.; Koutrakis, P. (2000) Characterization of indoor particle sources using continuous mass
      and size monitors. J. Air Waste Manage. Assoc. 50:  1236-1250.
Loomis, D.; Castillejos, M.; Gold, D. R.; McDonnell, W.; Borja-Aburto, V. H. (1999) Air pollution and infant
      mortality in Mexico City. Epidemiology 10:  118-123.
Lumley, T.; Heagerty, P. (1999) Weighted empirical adaptive variance estimators for correlated data regression.
      J. R. Stat. Soc. B 61(part 2): 459-477.
Lundgren, D. A.; Burton, R. M. (1995) Effect of particle size distribution on the cut point between fine and coarse
      ambient mass fractions. In: Phalen, R. F.; Bates, D. V., eds. Proceedings of the colloquium on paniculate air
      pollution and human mortality and morbidity; January 1994; Irvine, CA. Inhalation Toxicol. 7: 131-148.
April 2002                                     9A-10         DRAFT-DO NOT QUOTE OR CITE

-------
Mage, D.; Wilson, W.; Hasselblad, V.; Grant, L. (1999) Assessment of human exposure to ambient paniculate
      matter. J. Air Waste Manage. Assoc. 49: 174-185.
Mar, T. F.; Norris, G. A.; Koenig, J. Q.; Larson, T. V. (2000) Associations between air pollution and mortality in
      Phoenix, 1995-1997. Environ. Health Perspect. 108: 347-353.
McConnell, R.; Berhane, K.; Gilliland, F.; London, S. J.; Vora, H.; Avol, E.; Gauderman, W. J.; Margolis, H. G.;
      Lurmann, F.; Thomas, D. C.; Peters, J. M. (1999) Air pollution and bronchitic symptoms in southern
      California children with asthma. Environ. Health Perspect. 107: 757-760.
McDonnell, W. F.; Nishino-Ishikawa, N.; Petersen, F. F.; Chen, L. H.; Abbey, D. E. (2000) Relationships of
      mortality with the fine and coarse fractions of long-term ambient PM10 concentrations in nonsmokers.
      J. Exposure Anal. Environ. Epidemiol. 10: 427-436.
Michel, O.; Nagy, A.-M.; Schroeven, M.; Duchateau, J.; Neve, J.; Fondu, P.; Sergysels, R. (1997) Dose-response
      relationship to inhaled endotoxin in normal subjects. Am. J. Respir. Crit. Care Med. 156: 1157-1164.
Monn, C. (2001) Exposure assessment of air pollutants: a review on spatial heterogeneity and
      indoor/outdoor/personal exposure to suspended paniculate matter, nitrogen dioxide and ozone. Atmos.
      Environ. 35: 1-32.
Monn, C.; Becker, S. (1999) Cytotoxicity and induction of proinflammatory cytokines from human monocytes
      exposed to fine (PM2 5) and coarse  particles (PM10_2 5) in outdoor and indoor air.  Toxicol. Appl. Pharmacol.
      155: 245-252.
Moolgavkar, S. H. (2000a) Air pollution and daily mortality in three U.S. counties. Environ. Health Perspect.
      108: 777-784.
Moolgavkar, S. H. (2000b) Air pollution and hospital admissions for chronic obstructive pulmonary disease in three
      metropolitan areas in the United States. In: Grant, L. D., ed. PM2000: paniculate matter and health.
      Inhalation Toxicol. 12(suppl. 4): 75-90.
Moolgavkar, S. H. (2000c) Air pollution and hospital admissions for diseases of the circulatory system in three U.S.
      metropolitan areas. J. Air Waste Manage Assoc. 50: 1199-1206.
Moolgavkar, S. H.; Luebeck, E. G.; Anderson, E. L. (1997) Air pollution and hospital admissions for respiratory
      causes in Minneapolis-St. Paul and Birmingham. Epidemiology 8: 364-370.
Moolgavkar, S. H.; Hazelton, W.; Luebeck, G.; Levy, D.; Sheppard, L. (2000) Air pollution, pollens, and
      admissions for chronic respiratory disease in King County, Washington. In: Inhalation Toxicology:
      proceedings of the third colloquium on paniculate air pollution and human health; June, 1999; Durham, NC.
      Inhalation Toxicology 12(suppl. 1): 157-171.
Morris, R. D.; Naumova, E. N. (1998) Carbon monoxide and hospital admissions for congestive heart failure:
      evidence of an increased effect at low temperatures. Environ. Health Perspect. 106: 649-653.
Naeher, L. P.; Holford, T. R.; Beckett, W. S.; Belanger, K.; Triche, E. W.; Bracken, M. B.; Leaderer, B. P. (1999)
      Healthy women's PEF variations with ambient summer concentrations of PM10, PN2 5, SO42_, H+, and O3.
      Am. J. Respir. Crit. Care Med. 160: 117-125.
NARSTO. (2002) Fine particle assessment (2nd review draft). Available: http://www.cgenv.com/Narsto/ (13
      February 2002).
National Institutes of Health. (1997) Guidelines for the diagnosis and management of asthma: expert panel report 2.
      Bethesda, MD: U.S. Department of Health and Human Services, National Heart, Lung, and Blood Institute;
      publication no. 97-4051.
National Research Council. (1983) Acid deposition: atmospheric processes in eastern North America, a review of
      current scientific understanding. Washington, DC: National Academy Press.
National Research Council. (1994) Science and judgment in risk assessment. Washington, DC: National Academy
      Press.
National Research Council. (1998) Research priorities for airborne paniculate matter. I. Immediate priorities and a
      long-range research portfolio. Washington, DC:  National Academy Press.
National Research Council. (2001) Research priorities for airborne paniculate matter. III. Early research progress.
      Washington, DC: National Academy Press. Available: http://www.nap.edu/catalog/10065.html [30 April,
      2002].
Nauenberg, E.; Basu, K. (1999) Effect of insurance coverage on the relationship between asthma hospitalizations
      and exposure to air pollution. Public Health Rep. 114: 135-148.
Neas, L. M.; Dockery, D. W.; Ware, J. H.; Spengler, J. D.; Ferris, B. G., Jr.; Speizer, F. E. (1994) Concentration of
      indoor paniculate matter as a determinant of respiratory health in children. Am. J.  Epidemiol.
      139: 1088-1099.


April 2002                                     9A-11         DRAFT-DO NOT QUOTE OR CITE

-------
Neas, L. M; Dockery, D. W.; Koutrakis, P.; Tollerud, D. J.; Speizer, F. E. (1995) The association of ambient air
      pollution with twice daily peak expiratory flow rate measurements in children. Am. J. Epidemiol.
      141: 111-122.
Neas, L. M.; Dockery, D. W.; Burge, H.; Koutrakis, P.; Speizer, F. E. (1996) Fungus spores, air pollutants, and
      other determinants of peak expiratory flow rate in children. Am. J. Epidemiol. 143: 797-807.
Neas, L. M.; Schwartz, J.; Dockery, D. (1999) A case-crossover analysis of air pollution and mortality in
      Philadelphia. Environ. Health Perspect. 107: 629-631.
Nikula, K. J.; Vallyathan, V.; Green, F. H. Y.; Hahn, F. F. (2000) Influence of dose on the distribution of retained
      paniculate material in rat and human lungs. Presented at: PM2000: paniculate matter and health—the
      scientific basis for regulatory decision-making, specialty conference & exhibition; January; Charleston,  SC.
      Pittsburgh, PA: Air & Waste Management Association.
Norris, G.; Young-Pong, S. N.; Koenig, J. Q.; Larson, T. V.; Sheppard, L.;  Stout, J. W. (1999) An association
      between fine particles and asthma emergency department visits for children in Seattle. Environ. Health
      Perspect. 107: 489-493.
Norris, G.; Larson, T.; Koenig, J.; Claiborn, C.;  Sheppard, L.; Finn, D. (2000) Asthma aggravation, combustion, and
      stagnant air. Thorax 55: 466-470.
Oberdorster, G.; Ferin, J.; Lehnert, B. E. (1994) Correlation between particle size, in vivo particle persistence, and
      lung injury. Environ. Health Perspect. 102(suppl. 5): 173-179.
Osier, M.; Oberdorster, G. (1997) Intratracheal inhalation vs intratracheal instillation: differences in particle effects.
      Fundam. Appl.  Toxicol. 40: 220-227.
Ostro, B. (1995) Fine paniculate air pollution and mortality in two Southern California counties. Environ.  Res.
      70: 98-104.
Ostro, B. D.; Lipsett, M.  J.; Wiener, M. B.; Seiner, J. C. (1991) Asthmatic responses to airborne acid aerosols.
      Am. J. Public Health 81: 694-702.
Ostro, B. D.; Lipsett, M.  J.; Mann, J. K.; Krupnick, A.; Harrington, W. (1993) Air pollution and respiratory
      morbidity among adults in Southern California. Am. J. Epidemiol. 137: 691-700.
Ostro, B. D.; Broadwin, R.; Lipsett, M. J. (2000) Coarse and fine particles and daily mortality in the Coachella
      Valley, California: a follow-up study. J. Exposure Anal. Environ. Epidemiol. 10: 412-419.
Ott, W.; Wallace, L.; Mage, D. (2000) Predicting paniculate (PM10) personal exposure distributions using  a random
      component superposition statistical model. J. Air Waste Manage. Assoc. 50:  1390-1406.
Ozkaynak, H.; Xue, J.; Zhou, H.; Raizenne, M. (1996) Associations between daily mortality and motor vehicle
      pollution in Toronto, Canada. Boston, MA: Harvard University School of Public Health, Department of
      Environmental Health; March 45.
Patterson, E.; Eatough, D. J. (2000) Indoor/outdoor relationships for ambient PM2 5  and associated pollutants:
      epidemiological implications inLindon, Utah. J. Air Waste Manage. Assoc. 50: 103-110.
Pekkanen, J.; Timonen, K. L.; Ruuskanen, J.; Reponen, A.; Mirme, A. (1997) Effects of ultrafine and fine particles
      in urban air on peak expiratory flow among children with asthmatic symptoms. Environ. Res. 74:  24-33.
Peters, A.; Doring, A.; Wichmann, H.-E.; Koenig, W. (1997a) Increased plasma viscosity during an air pollution
      episode: a link to mortality? Lancet 349: 1582-1587.
Peters, A.; Wichmann, H. E.; Tuch, T.; Heinrich, J.; Heyder, J. (1997b) Respiratory effects are associated  with the
      number of ultrafine particles. Am. J. Respir. Crit. Care Med. 155: 1376-1383.
Peters, A.; Dockery, D. W.; Heinrich, J.; Wichmann, H. E. (1997c) Short-term effects of paniculate air pollution on
      respiratory morbidity in asthmatic children. Eur. Respir. J. 10: 872-879.
Peters, J. M.; Avol, E.; Gauderman, W. J.; Linn, W. S.; Navidi, W.; London, S. J.; Margolis, H.; Rappaport, E.;
      Vora, H.; Gong, H., Jr.; Thomas, D. C. (1999a) A study of twelve southern California communities with
      differing levels and types of air pollution. II. Effects on pulmonary function.  Am. J. Respir. Crit.  Care Med.
      159: 768-775.
Peters, J. M.; Avol, E.; Navidi, W.; London, S.  J.; Gauderman, W. J.; Lurmann, F.; Linn, W. S.; Margolis, H.;
      Rappaport, E.; Gong, H., Jr.; Thomas, D. C. (1999b) A study of twelve southern California communities with
      differing levels and types of air pollution. I. Prevalence of respiratory morbidity. Am. J. Respir. Crit. Care
      Med. 159: 760-767.
Peters, A.; Liu, E.; Verrier, R. L.; Schwartz, J.; Gold, D. R.; Mittleman, M.; Baliff, J.; Oh, J. A.; Allen, G.;
      Monahan, K.; Dockery, D. W. (2000) Air pollution and incidence of cardiac arrhythmia. Epidemiology 11:
      11-17.
April 2002                                     9A-12         DRAFT-DO NOT QUOTE OR CITE

-------
Pinto, J. P.; Pleasant, M; Kellogg, R.; Torok, Sz.; Stiles, D.; Willis, R. D.; Koutrakis, P.; Allen, G. A. (1995)
      Chemical characterization of the ambient aerosol in Philadelphia. Presented at: Paniculate matter: health and
      regulatory issues: an A&WMA international specialty conference; April; Pittsburgh, PA. Pittsburgh, PA: Air
      & Waste Management Association; submitted abstract.
Pinto, J. P.; Lefohn, A.  S.; Shadwick, D. S. (2002) Aspects of the spatial variability of PM25 concentrations within
      urban areas of the United States. Environ. Sci. Technol.: submitted.
Plopper, C. G.; Fanucchi, M. V. (2000) Do urban environmental pollutants exacerbate childhood lung diseases?
      Environ. Health Perspect. 108: A252-A253.
Pope, C. A., III. (1989) Respiratory disease associated with community air pollution and a steel mill, Utah Valley.
      Am. J. Public Health 79: 623-628.
Pope, C. A., Ill; Thun, M. J.; Namboodiri, M. M.; Dockery, D. W.; Evans, J. S.; Speizer, F. E.; Heath, C. W., Jr.
      (1995) Paniculate air pollution as a predictor of mortality in a prospective study of U.S. adults. Am. J.
      Respir. Crit. Care Med. 151: 669-674.
Pope, C. A.; Verrier, R. L.; Lovett, E. G.; Larson, A. C.; Raizenne, M. E.; Kanner, R. E.; Schwartz, J.; Villegas,
      G. M.; Gold, D. R.; Dockery, D. W. (1999a) Heart rate variability associated with paniculate air pollution.
      Am. Heart J. 138: 890-899.
Pope, C. A., Ill; Hill, R. W.; Villegas, G. M. (1999b) Paniculate air pollution and daily mortality on Utah's Wasatch
      Front. Environ. Health Perspect.: 107: 567-573.
Pope, C. A., Ill; Eatough, D. J.; Gold, D. R.; Pang, Y.; Nielsen, K. R.; Nath, P.; Verrier, R. L.; Kanner, R. E. (2001)
      Acute exposure to environmental tobacco smoke and heart rate variability. Environ. Health Perspect.
      109:711-716.
Pope, C. A., Ill; Burnett, R.  T.; Thun, M. J.; Calle, E. E.; Krewski, D.; Ito, K.; Thurston, G. D. (2002) Lung cancer,
      cardiopulmonary mortality, and long-term exposure to fine paniculate air pollution. JAMA J. Am. Med.
      Assoc. 287:  1132-1141.
Raizenne, M.; Neas, L. M.; Damokosh, A. L; Dockery, D.  W.; Spengler, J. D.; Koutrakis, P.; Ware, J. H.; Speizer,
      F. E. (1996) Health effects of acid aerosols on North American children:  pulmonary function. Environ.
      Health Perspect.  104: 506-514.
Ramadan, Z.; Song, X.-H.; Hopke, P. K. (2000) Identification of sources of Phoenix aerosol by positive  matrix
      factorization. J. Air Waste  Manage. Assoc. 50: 1308-1320.
Ritz, B.; Yu, F.; Fruin, S.; Chapa, G.; Shaw, G. M.; Harris, J. A. (2002) Ambient air pollution and risk of birth
      defects in southern California. Am. J. Epidemiol. 155: 17-25.
Romieu, I.; Meneses, F.; Ruiz, S.; Sienra, J. J.; Huerta, J.; White, M. C.; Etzel, R. A. (1996) Effects of air pollution
      on the respiratory health of asthmatic children living in Mexico City. Am. J. Respir. Crit. Care Med.
      154: 300-307.
Romieu, I.; Meneses, F.; Ruiz, S.; Huerta, J.; Sienra, J. J.; White, M.; Etzel, R.;  Hernandez, M. (1997) Effects of
      intermittent ozone exposure on peak expiratory flow and respiratory symptoms among asthmatic children in
      Mexico City. Arch. Environ. Health 52: 368-376.
Samet, J. M.; Dominici, F.; Zeger, S. L.; Schwartz, J.; Dockery, D. W. (2000a) National morbidity, mortality, and
      air pollution study. Part I: methods and methodologic issues. Cambridge, MA: Health Effects Institute;
      research report no. 94.
Samet, J. M.; Zeger, S. L.; Dominici, F.; Curriero, F.; Coursac, I.; Dockery, D. W.; Schwartz, J.; Zanobetti, A.
      (2000b) The national morbidity, mortality, and air pollution study. Part II: morbidity, mortality, and air
      pollution in the United States. Cambridge, MA: Health Effects Institute; research report no. 94.
Samet, J. M.; Hatch, G. E.; Horstman, D.; Steck-Scott, S.; Arab, L.; Bromberg, P. A.; Levine, M.; McDonnell,
      W.  F.; Devlin, R. B. (2001) Effect of antioxidant supplementation on ozone-induced lung injury in human
      subjects. Am. J. Respir. Crit. Care Med.  164: 819-825.
Sarnat, J. A.; Koutrakis, P.;  Suh, H. H.  (2000) Assessing the relationship between personal paniculate and gaseous
      exposures of senior citizens living in Baltimore, MD. J. Air Waste Manage. Assoc. 50: 1184-1198.
Sarnat, J. A.; Schwartz, J.; Catalano, P. J.; Suh, H. H. (2001) Gaseous pollutants in paniculate matter epidemiology:
      confounders or surrogates? Environ. Health Perspect.  109: 1053-1061.
Schlesinger, R. B.;  Ben-Jebria, A.; Dahl, A. R.; Snipes, M. B.; Ultman, J. (1997) Disposition of inhaled toxicants.
      In: Massaro, E. J., ed. Handbook of human toxicology. Boca Raton, FL: CRC Press; pp. 493-550.
Schwartz, J. (1997) Air pollution and hospital admissions for cardiovascular disease in Tucson. Epidemiology
      8: 371-377.
April 2002                                     9A-13         DRAFT-DO NOT QUOTE OR CITE

-------
Schwartz, J. (1999) Air pollution and hospital admissions for heart disease in eight U.S. counties. Epidemiology
      10: 17-22.
Schwartz, J. (2000) Harvesting and long term exposure effects in the relation between air pollution and mortality.
      Am. J. Epidemiol. 151: 440-448.
Schwartz, J.; Morris, R. (1995) Air pollution and hospital admissions for cardiovascular disease in Detroit,
      Michigan. Am. J. Epidemiol. 142: 23-35.
Schwartz, J.; Neas, L. M. (2000) Fine particles are more strongly associated than coarse particles with acute
      respiratory health effects in schoolchildren. Epidemiology.  11: 6-10.
Schwartz, J.; Zanobetti, A. (2000) Using meta-smoothing to estimate dose-response trends across multiple studies,
      with application to air pollution and daily death. Epidemiology 11: 666-672.
Schwartz, J.; Dockery, D. W.; Neas, L. M.; Wypij, D.; Ware, J. H.; Spengler, J. D.; Koutrakis, P.; Speizer, F. E.;
      Ferris, E.G., Jr. (1994) Acute effects of summer air pollution on respiratory symptom reporting in children.
      Am. J. Respir. Crit. Care Med. 150: 1234-1242.
Schwartz, J.; Dockery, D. W.; Neas, L. M. (1996) Is daily mortality associated specifically with fine particles?
      J. Air Waste Manage. Assoc. 46: 927-939.
Schwartz, J.; Norris, G.; Larson, T.; Sheppard, L.; Claiborne, C.; Koenig, J. (1999) Episodes of high coarse particle
      concentrations are not associated with increased mortality. Environ. Health Perspect. 107: 339-342.
Sexton, K.; Selevan, S. G.; Wagener, D. K.; Lybarger, J. A. (1992) Estimating human exposures to environmental
      pollutants: availability and utility of existing databases. Arch. Environ. Health 47: 398-407.
Sheppard, L.; Levy, D.; Norris, G.; Larson, T. V.; Koenig, J. Q. (1999) Effects of ambient air pollution on
      nonelderly asthma hospital admissions in Seattle, Washington, 1987-1994. Epidemiology 10: 23-30.
Smiley-Jewell, S. M.; Liu, F. J.; Weir, A. J.; Plopper,  C. G. (2000) Acute injury to differentiating Clara cells in
      neonatal rabbits results in age-related failure of bronchiolar epithelial repair. Toxicol. Pathol. 28: 267-276.
Smith, R. L.; Spitzner, D.; Kim, Y.; Fuentes, M. (2000) Threshold dependence of mortality effects for fine and
      coarse particles in Phoenix, Arizona. J. Air Waste Manage. Assoc. 50: 1367-1379.
Stieb, D. M.; Beveridge, R. C.; Brook, J. R.;  Smith-Doiron, M.; Burnett, R. T.; Dales, R. E.; Beaulieu, S.; Judek, S.;
      Mamedov, A. (2000) Air pollution, aeroallergens and cardiorespiratory emergency department visits in Saint
      John, Canada. J. Exposure Anal. Environ. Epidemiol. 10: 461-477.
Thurston,  G. D.; Ito, K.; Kinney, P. L.; Lippmann, M. (1992) A multi-year study of air pollution and respiratory
      hospital admissions in three New York State metropolitan areas: results for 1988 and 1989 summers.
      J. Exposure Anal. Environ. Epidemiol. 2: 429-450.
Thurston,  G. D.; Ito, K.; Hayes, C. G.; Bates, D. V.; Lippmann, M. (1994) Respiratory hospital admissions and
      summertime haze air pollution in Toronto, Ontario: consideration of the role of acid aerosols. Environ. Res.
      65: 271-290.
Thurston,  G. D.; Lippmann, M.;  Scott, M. B.; Fine, J. M. (1997) Summertime haze air pollution and children with
      asthma. Am. J. Respir. Crit. Care Med. 155: 654-660.
Tiittanen,  P.; Timonen, K. L.; Ruuskanen, J.; Mirme, A.; Pekkanen, J. (1999) Fine particulate air pollution,
      resuspended road dust and respiratory health among symptomatic children. Eur. Respir. J. 13: 266-273.
Tolbert, P. E.; Klein, M.; Metzger, K. B.; Peel, J.; Flanders, W. D.; Todd, K.; Mulholland,  J. A.; Ryan, P. B.;
      Frumkin, H. (2000) Interim results of the study of particulates and health in Atlanta (SOPHIA). J. Exposure
      Anal. Environ. Epidemiol. 10: 446-460.
Tsai, F. C.; Apte, M. G.; Daisey, J. M. (2000) An exploratory analysis of the relationship between mortality and the
      chemical composition of airborne particulate matter. Inhalation Toxicol. 12(suppl.): 121-135.
Turpin, B. J. (1999) Options for characterizing organic particulate matter. Environ. Sci. Technol. 33: 76A-79A.
U.S. Environmental Protection Agency. (1993) Air quality criteria for oxides of nitrogen. Research Triangle  Park,
      NC: Office of Health and Environmental Assessment, Environmental Criteria and Assessment Office; report
      nos. EPA/600/8-9!/049aF-cF. 3v. Available from: NTIS, Springfield, VA; PB95-124533, PB95-124525, and
      PB95-124517.
U.S. Environmental Protection Agency. (1996a) Air quality criteria for particulate matter. Research Triangle Park,
      NC: National Center for Environmental Assessment-RTF Office; report nos. EPA/600/P-95/00laF-cF. 3v.
U.S. Environmental Protection Agency. (1996b) Air quality criteria for ozone and related photochemical oxidants.
      Research Triangle Park, NC: Office of Research and Development; report nos. EPA/600/AP-93/004aF-cF.
      3v. Available from: NTIS, Springfield, VA; PB96-185582, PB96-185590, andPB96-185608. Available
      online at: www.epa.gov/ncea/ozone.htm.
April 2002                                     9A-14         DRAFT-DO NOT QUOTE OR CITE

-------
U.S. Environmental Protection Agency. (2000a) Air quality criteria for carbon monoxide. Research Triangle Park,
      NC: National Center for Environmental Assessment; report no. EPA/600/P-99/001F. Available:
      www.epa.gov/ncea/co/ [2000, October 6].
U.S. Environmental Protection Agency. (2000b) Health assessment document for diesel exhaust [SAB review
      draft]. Washington, DC: Office of Research and Development, National Center for Environmental
      Assessment-Washington Office; report no. EPA/600/8-90/057E. Available:
      http://cfpub.epa.gov/ncea/cfm/nceahome.cfm [29 March, 2002].
U.S. Environmental Protection Agency. (2001) National air quality and emissions trends report, 1999. Research
      Triangle Park, NC: Office of Air Quality Planning and Standards; report no. EPA/454/R-01-004. Available:
      http://www.epa.gov/oar/aqtrnd99/ [11 March, 2002].
Undem, B. I; Carr, M. J. (2002) The role of nerves in asthma. Curr. Allergy Asthma Rep. 2: 159-165.
Vedal, S.; Petkau, J.; White, R.; Blair, J. (1998) Acute effects of ambient inhalable particles in asthmatic and
      nonasthmatic children. Am. J. Respir. Crit. Care Med. 157: 1034-1043.
Wallace, L. (2000) Correlations of personal exposure to particles with outdoor air measurements: a review of recent
      studies. Aerosol Sci. Technol. 32: 15-25.
Ware, J. H.; Ferris, B. G., Jr.; Dockery, D. W.;  Spengler, J. D.; Stram, D. O.; Speizer, F. E. (1986) Effects of
      ambient sulfur oxides and suspended particles on respiratory health of preadolescent children. Am. Rev.
      Respir. Dis. 133:834-842.
Watkinson, W. P.; Campen, M. J.; Costa, D. L. (1998) Cardiac arrhythmia induction after exposure to residual oil
      fly ash particles in a rodent model of pulmonary hypertension. Toxicol. Sci. 41: 209-216.
Watkinson, W. P.; Campen, M. J.; Nolan, J. P.; Costa, D. L. (2001) Cardiovascular and systemic responses to
      inhaled pollutants in rodents: effects of ozone and paniculate matter. Environ. Health Perspect.
      109(suppl. 4): 539-546.
Whitby, K. T. (1978) The physical characteristics of sulfur aerosols. Atmos. Environ. 12: 135-159.
Whitby, K. T.; Sverdrup, G. M. (1980) California aerosols: their physical and chemical characteristics. In: Hidy,
      G. M.; Mueller, P. K.; Grosjean, D.; Appel, B. R.; Wesolowski, J. J., eds. The character and origins of smog
      aerosols: a digest of results from the California Aerosol Characterization Experiment (ACHEX). New York,
      NY: John Wiley & Sons, Inc.; pp. 477-517. (Advances in environmental science and technology: v. 9).
Wichmann, H.-E.; Spix, C.; Tuch, T.; Wolke, G.; Peters, A.; Heinrich, J.; Kreyling, W. G.; Heyder, J. (2000) Daily
      mortality and fine and ultrafine particles in Erfurt, Germany. Part I:  role of particle number and particle mass.
      Cambridge, MA: Health Effects Institute; Research Report no.  98.
Williams, R.; Suggs, J.; Zweidinger, R.; Evans, G.; Creason, J.; Kwok, R.; Rodes, C.; Lawless, P.; Sheldon, L.
      (2000a) The 1998 Baltimore Paniculate Matter Epidemiology-Exposure Study: part 1. Comparison of
      ambient, residential outdoor,  indoor and apartment paniculate matter monitoring. J. Exposure Anal. Environ.
      Epidemiol.  10: 518-532.
Williams, R.; Suggs, J.; Creason,  J.; Rodes,  C.; Lawless, P.; Kwok, R.; Zweidinger, R.;  Sheldon, L. (2000b) The
      1998 Baltimore Paniculate Matter Epidemiology-Exposure Study: part 2. Personal exposure assessment
      associated with an elderly study population. J. Exposure Anal. Environ. Epidemiol. 10: 533-543.
Williams, R. W.; Creason, J.; Zweidinger, R.; Watts, R. R.; Shy, C. (2000c) Indoor, outdoor, and personal exposure
      monitoring of paniculate air pollution: the Baltimore elderly epidemiology-exposure pilot study.
      Atmos. Environ. 34: 4193-4204.
Wilson, W. E.; Suh, H. H. (1997) Fine particles and coarse particles: concentration relationships relevant to
      epidemiologic studies. J. Air Waste Manage. Assoc. 47: 1238-1249.
Wilson, W. E.; Mage, D. T.; Grant, L. D. (2000) Estimating separately personal exposure to ambient and
      nonambient paniculate matter for epidemiology and risk assessment: why and how. J. Air Waste Manage.
      Assoc. 50: 1167-1183.
Wongphatarakul, V.; Friedlander, S. K.; Pinto,  J. P. (1998) A comparative study of PM25 ambient aerosol chemical
      databases. Environ. Sci. Technol. 32: 3926-3934.
Woodruff, T. J.; Grille, J.; Schoendorf, K. C. (1997) The relationship between selected causes of postneonatal
      infant mortality and paniculate air pollution in the United States. Environ. Health Perspect. 105: 608-612.
Zanobetti, A.; Schwartz, J.  (2000) Race, gender, and social status as modifiers of the effects of PM10 on mortality.
      J. Occup. Environ. Med. 42: 469-474.
Zanobetti, A.; Schwartz, J.  (2001) Are diabetics more susceptible to the health effects of airborne particles? Am. J.
      Respir. Crit. Care Med. 164:  831-833.
April 2002                                      9A-15         DRAFT-DO NOT QUOTE OR CITE

-------
Zanobetti, A.; Schwartz, J.; Gold, D. (2000) Are there sensitive subgroups for the effects of airborne particles?
      Environ. Health Perspect.  108: 841-845.
Zanobetti, A.; Schwartz, J.; Samoli, E.; Gryparis, A.; Touloumi, G.; Atkinson, R.; Le Tertre, A.; Bobros, J.;
      Celko, M.; Goren, A.; Forsberg, B.; Michelozzi, P.; Rabczenko, D.; Ruiz, E. A.; Katsouyanni, K. (2002)
      The temporal pattern of mortality responses to air pollution: a multicity assessment of mortality displacement.
      Epidemiology 13: 87-93.
Zeger, S. L.; Thomas, D.; Dominici, F.; Samet, J. M.; Schwartz, J.; Dockery, D.; Cohen, A. (2000) Exposure
      measurement error in time-series studies of air pollution: concepts and consequences. Environ. Health
      Perspect. 108: 419-426.
Zhang, T.; Huang, C.; Johns, E. J. (1997) Neural regulation of kidney function by the somatosensory system in
      normotensive and hypertensive rats. Am. J. Physiol. 273: R1749-R1757.
Zhang, H.; Triche, E.; Leaderer, B. (2000) Model forthe analysis of binary time series of respiratory symptoms.
      Am. J. Epidemiol.  151: 1206-1215.
Zweidinger, R.; Purdue, L.; Fitzgerald, K.; Carmichael, L.; Kellog, R. (1998)  NERL PM research monitoring
      platforms: Baltimore, Fresno and Phoenix, data report. Research Triangle Park, NC: U.S. Environmental
      Protection Agency, National Exposure Research Laboratory; report no. EPA/600/R-98/138. Available from:
      NTIS, Springfield, VA; PB99-105660.
April 2002                                     9A-16         DRAFT-DO NOT QUOTE OR CITE

-------