vvEPA
              United States
              Environmental Protection
              Agency
               Office of Research and
               Development
                     DC 20460
EPA/600/AP-95/001a
April 1995
External Review Draft
                                                      C.I
Air Quality
Criteria for
Paniculate
Matter

Volume I of III
Review
Draft
(Do Not
Cite or
Quote)
                              Notice
              This document is a preliminary draft. It has not been formally
              released by EPA and should not at this stage be construed to
              represent Agency policy. It is being circulated for comment on its
              technical accuracy and policy implications.

-------
DRAFT-DO NOT QUOTE OR CITE                                 EPA/eoo/AP-95/ooia
                                                                 April1995
                                                                 External Review Draft
                     Air Quality Criteria for
                         Particulate Matter
i


                             Volume  I  of
'' i

"*"*' i
 'l
                                     NOTICE
                  This document is a preliminary draft.  It has not been formally
                  released by EPA and should not at this stage be construed to
                  represent Agency policy. It is being circulated for comment on its
                  technical accuracy and policy implications.
                     Environmental Criteria and Assessment Office
                     Office of Health and Environmental Assessment
                         Office of Research and Development
                        U.S. Environmental Protection Agency
                         Research Triangle Park, NC  27711
                                                                 Printed on Recycled Paper

-------
                                  DISCLAIMER

     This document is an external draft for review purposes only and does not constitute
U.S. Environmental Protection Agency policy.  Mention of trade names or commercial
products does not constitute endorsement or recommendation for use.
April 1995                               I-ii       DRAFT-DO NOT QUOTE OR CITE

-------
                 Air Quality Criteria for Participate Matter


                         TABLE OF CONTENTS

                              Volume I

 1. EXECUTIVE SUMMARY  	1-1

 2. INTRODUCTION	2-1

 3. PHYSICS AND CHEMISTRY OF PARTICULATE MATTER	3-1

 4. SAMPLING AND ANALYSIS OF PARTICULATE MATTER AND
   ACID DEPOSITION	4-1

 5. SOURCES AND EMISSIONS OF SUSPENDED PARTICLES	5-1

 6. ENVIRONMENTAL CONCENTRATIONS	6-1
   Appendix 6A:  Tables of Chemical Composition of PM	6A-1

 7. EXPOSURE: AMBIENT AND INDOOR  	7-1

                              Volume II

 8. EFFECTS ON VISIBILITY AND CLIMATE	8-1

 9. EFFECTS ON MATERIALS	9-1

10. DOSIMETRY OF INHALED PARTICLES IN THE
   RESPIRATORY TRACT  	10-1

11. TOXICOLOGY OF PARTICULATE MATTER CONSTITUENTS	11-1

                              Volume III

12. EPIDEMIOLOGY STUDIES OF HEALTH EFFECTS ASSOCIATED
   WITH EXPOSURE TO AIRBORNE PARTICLES/ACID
   AEROSOLS  	12-1
   Appendix 12A:  Effects of Weather and Climate on Human Mortality and Their
                Roles as Confounding Factors for Air Pollution 	12A-1

13. INTEGRATIVE HEALTH SYNTHESIS	13-1
April 1995                        I-iii     DRAFT-DO NOT QUOTE OR CITE

-------
                        TABLE OF CONTENTS
                                                              "Jase
LIST OF TABLES  	I-xiv
LIST OF FIGURES	I-xviii
AUTHORS, CONTRIBUTORS, AND REVIEWERS 	  I-xxxi
U.S. ENVIRONMENTAL PROTECTION AGENCY PROJECT TEAM
 FOR DEVELOPMENT OF AIR QUALITY CRITERIA FOR
 PARTICULATE MATTER	I-xl
U.S. ENVIRONMENTAL PROTECTION AGENCY SCIENTIFIC
 ADVISORY BOARD, CLEAN AIR SCIENTIFIC ADVISORY
 COMMITTEE	I-xlv
1.  EXECUTIVE SUMMARY  	1-1
   1.1   PURPOSE OF DOCUMENT 	1-1
   1.2   INTRODUCTION  	1-1
   1.3   PHYSICS AND CHEMISTRY OF PARTICULATE MATTER	1-2
   1.4   SAMPLING AND ANALYSIS OF PARTICULATE MATTER AND
        ACID DEPOSITION	1-4
   1.5   SOURCES AND EMISSIONS OF SUSPENDED PARTICLES	1-5
   1.6   ENVIRONMENTAL CONCENTRATIONS  	1-7
   1.7   EXPOSURE: AMBIENT AND INDOOR 	  1-13
   1.8   EFFECTS ON VISIBILITY AND CLIMATE  	  1-16
   1.9   EFFECTS ON MATERIALS 	1-18
   1.10  DOSIMETRY MODELING OF INHALED PARTICLES IN THE
        RESPIRATORY TRACT	1-20
   1.11  TOXICOLOGY OF PARTICULATE MATTER CONSTITUENTS	  1-22
   1.12  EPIDEMIOLOGY STUDIES OF HEALTH EFFECTS ASSOCIATED
        WITH EXPOSURE TO AIRBORNE PARTICLES/ACID
        AEROSOLS	1-31
   1.13  BIOLOGICAL PLAUSIBILITY POTENTIAL MECHANISMS
        OF ACTION 	1-70
   1.14  IDENTIFICATION OF POPULATION GROUPS
        POTENTIALLY SUSCEPTIBLE TO HEALTH EFFECTS FROM
        PM EXPOSURE	1-83
   1.15  IMPLICATIONS OF RELATIVE RISK ESTIMATES 	  1-87

2.  INTRODUCTION	2-1
   2.1   LEGISLATIVE REQUIREMENTS	2-1
   2.2   REGULATORY BACKGROUND  	2-2
   2.3   SCIENTIFIC BASIS FOR THE EXISTING PARTICULATE MATTER
        STANDARDS	2-4
        2.3.1   Rationale for the Primary Standards	2-4
        2.3.2   Pollutant Indicator	2-5
        2.3.3   Averaging Time and Form of the Standards	2-6
        2.3.4   Level of the Standards	2-8

April 1995                         I_v     DRAFT-DO NOT QUOTE OR CITE

-------
                          TABLE OF CONTENTS (cont'd)
          2.3.5   Welfare Effects   	2-17
    2.4   TOPICS/ISSUES OF CONCERN FOR CURRENT CRITERIA
          DEVELOPMENT  	2-17
          2.4.1   Air Quality and Exposure	2-18
          2.4.2   Health Effects	2-22
          2.4.3   Welfare Effects   	2-28
    2.5   DOCUMENT CONTENT AND ORGANIZATION	2-29
    REFERENCES	2-32

3.   PHYSICS AND CHEMISTRY OF PARTICULATE MATTER	3-1
    3.1   INTRODUCTION  	3-1
          3.1.1   Overview	3-1
          3.1.2   Major Chemical Constituents	3-4
          3.1.3   Atmospheric Aerosol Size Distributions	3-6
          3.1.4   Chemical Composition and Its Dependence on Particle Size   	3-9
          3.1.5   Particle-Vapor Partitioning  	3-11
          3.1.6   Single Particle Characteristics	3-12
          3.1.7   Definitions 	3-13
                  3.1.7.1  Definitions of Particle Diameter	3-13
                  3.1.7.2 Definitions of Particle Size Fractions	3-15
                  3.1.7.3 Other Terminology	3-16
          3.1.8   Field Studies	3-17
          3.1.9   Dry Deposition   	3-18
          3.1.10  Atmospheric Scavenging	3-18
    3.2   PHYSICAL PROPERTIES	3-18
          3.2.1   Aerosol Size Distributions	3-18
                  3.2.1.1  Particle Size Distribution Functions	3-18
                  3.2.1.2 Log-Normal Size Distributions	3-19
                  3.2.1.3 Ambient Aerosol Size Distributions	3-20
                  3.2.1.4 Coagulation of Spherical Particles	3-20
          3.2.2   Particle Formation and Growth	3-22
                  3.2.2.1  Equilibrium Vapor Pressures	3-22
                  3.2.2.2 New Particle Formation	3-23
                  3.2.2.3 Particle Growth	3-23
                  3.2.2.4 Resuspension  	3-24
          3.2.3   Particle Removal Mechanisms and Deposition	3-30
    3.3   CHEMISTRY  AND CHEMICAL COMPOSITION	3-33
          3.3.1   Fine Particle Chemistry 	3-33
                  3.3.1.1 Acid Aerosols and Paniculate Sulfates	3-33
          3.3.2   Formation of Sulfates in Clouds  	3-35
                  3.3.2.1 Particle Formation in Clouds	3-35
          3.3.3   Aqueous-Phase Oxidation of SO2 in Clear-Air Aerosols	3-43
          3.3.3   Physical and Chemical Considerations in  Particulate Sampling and
                  Analysis	3-44

April  1995                              I-vi       DRAFT-DO NOT QUOTE OR CITE

-------
                          TABLE OF CONTENTS (cont'd)
                  3.3.3.1  Semi-Volatile Organic Compounds	3-44
          3.3.4   Paniculate Nitrates  	3-49
                  3.3.4.1  Sources	3-49
                  3.3.4.2  Gas-Phase  	3-49
          3.3.5   Water Content and Aerosol Equilibria	3-51
                  3.3.5.1  Water Content of Atmospheric Aerosols, and Its
                          Dependence on Ambient Humidity  	3-51
                  3.3.5.2  Equilibria with Water Vapor	3-52
                  3.3.5.3  Ammonium Nitrate Vaporization Equilibria	3-54
          3.3.6   Carbon-Containing Paniculate Matter	3-56
                  3.3.6.1  Introduction	3-56
                  3.3.6.2  Elemental Carbon  	3-56
                  3.3.6.3  Organic Carbon	3-59
                  3.3.6.4  Primary Organic Carbon  	3-61
          3.3.7   Metals and Other Trace Elements  	3-69
          3.3.8   Removal Processes  	3-75
    3.4   TRANSPORT AND TRANSFORMATIONS TO SECONDARY
          PARTICULATE MATTER	3-76
          3.4.1   Aqueous-Phase Chemical Equilibria and Chemical Kinetics of
                  Transformations to Secondary Paniculate Matter	3-76
                  3.4.1.1  Aqueous-Phase Equilibria	3-76
                  3.4.1.2  Aqueous-Phase Transformation of S02 to Sulfate  	3-80
                  3.4.1.3  Aqueous-Phase Transformation of NO2 to HNO3 and
                          NH4N03	3-87
          3.4.2   Transport and Transformations in Plumes  	3-87
                  3.4.2.1  Field Studies of Transport Processes	3-88
          3.4.3   Transformations in Plumes  	3-98
                  3.4.3.1  Gas-to-Particle Conversion in Plumes  	3-98
    3.5   DRY DEPOSITION	3-112
          3.5.1   Theoretical Aspects of Dry Deposition	3-112
          3.5.2   Field Studies of Dry Deposition  	3-116
                  3.5.2.1  Measured Deposition Velocities	3-119
    3.6   WET DEPOSITION	3-121
          3.6.1   Introduction  	3-121
          3.6.2   Field Studies of Wet Deposition	3-123
                  3.6.2.1  Overview of SO2 and NOX Wet Scavenging	3-128
    3.7   PHYSICAL AND CHEMICAL CONSIDERATIONS IN PARTICULATE
          MATTER SAMPLING AND ANALYSIS	3-129
          3.7.1   Size Cut-Point for Separating Fine and Coarse Particulate
                  Matter	3-129
                  3.7.1.1  Background  	3-129
                  3.7.1.2  Size Measurements .	3-129
                  3.7.1.3  Appropriate Display of Size-Distribution Data	3-131
April 1995                              I_vii       DRAFT-DO NOT QUOTE OR CITE

-------
                          TABLE OF CONTENTS (cont'd)
                                                                              Page

                   3.7.1.4 Comparison of Particle-Counting and Particle-Collection
                          Techniques	3-134
                   3.7.1.5 Review of Size Distribution Data	3-138
                   3.7.1.6 Intermodal  Region	3-144
                   3.7.1.7 Conclusions	3-169
    3.8    SUMMARY	3-169
    REFERENCES	3-172

4.   SAMPLING AND ANALYSIS OF PARTICULATE MATTER AND
    ACID DEPOSITION	,	4-1
    4.1    INTRODUCTION   	4-1
    4.2    SAMPLING FOR PARTICULATE MATTER	4-5
           4.2.1    Background	4-5
           4.2.2    Large Particle Separators  	4-6
                   4.2.2.1 Cutpoint Considerations	4-6
                   4.2.2.2 Total Suspended  Particulates (TSP)	4-11
                   4.2.2.3 Total Inhalable   	4-12
                   4.2.2.4 PM10	4-12
           4.2.3    Fine Particle Separators   	4-20
                   4.2.3.1 Cutpoint Considerations	4-20
                   4.2.3.2 Virtual Impactors	4-21
                   4.2.3.3 Cyclones   	4-23
                   4.2.3.4 Impactors  	4-25
           4.2.4    Sampling Considerations	4-26
                   4.2.4.1 Siting Criteria	4-26
                   4.2.4.2 Averaging Time/Sampling Frequency  	4-27
                   4.2.4.3 Collection Substrates	4-28
                   4.2.4.4 Chemical Speciation Sampling	4-30
                   4.2.4.5 Data Corrections/Analyses  	4-33
           4.2.5    Performance Specifications  	4-34
                   4.2.5.1 Approaches 	4-34
                   4.2.5.2 Critiques	4-36
           4.2.6    Reference and Equivalent Method Program  	4-39
           4.2.7    Determination of Size Distribution	4-43
                   4.2.7.1 Cascade Impactors	4-43
                   4.2.7.2 Single Particle Samplers   	4-45
           4.2.8    Automated Sampling	4-48
                   4.2.8.1 TEOM  	4-48
                   4.2.8.2 Beta Gauge 	4-50
                   4.2.8.3 Nephelometer	4-52
           4.2.9    Specialized Sampling	4-57
                   4.2.9.1 Personal Exposure Sampling	4-57
                   4.2.9.2 Receptor Model  Sampling  	4-60
                   4.2.9.3 Particle Acidity	4-61

April  1995                              I-viii      DRAFT-DO NOT QUOTE OR CITE

-------
                        TABLE OF CONTENTS (cont'd)
    4.3   ANALYSIS OF PARTICULATE MATTER  	4-63
          4.3.1   Mass Measurement Methods  	4-67
          4.3.2   Physical Analysis	4-69
                 4.3.2.1 X-Ray Fluorescence of Trace Elements 	4-69
                 4.3.2.2 Particle Induced X-Ray Emission of
                        Trace Elements	4-78
                 4.3.2.3 Instrumental Neutron Activation Analysis of
                        Trace Elements	4-80
                 4.3.2.4 Microscopy Analysis of Particle Size, Shape,
                        and Composition	4-84
          4.3.3   Wet Chemical Analysis  	4-86
                 4.3.3.1 Ion Chromatographic Analysis for Chloride,
                        Nitrate, and Sulfate  	4-87
                 4.3.3.2 Automated Colorimetric Analysis for Ammonium,
                        Nitrate, and Sulfate  	4-90
                 4.3.3.3 Atomic Absorption Spectrophotometric and
                        Inductive Coupled Plasma Atomic Emission
                        Spectrophotometry Analyses for Trace Elements	4-92
          4.3.4   Organic Analysis  	4-93
                 4.3.4.1 Analysis of Organic Compounds  	4-93
                 4.3.4.2 Analysis of Organic and Elemental Carbon	4-97
          4.3.5   Quality Assurance	4-99
    REFERENCES	4-101

5.  SOURCES AND EMISSIONS OF SUSPENDED PARTICLES	5-1
    5.1   INTRODUCTION  	5-1
    5.2   SUMMARY OF 1982 CRITERIA DOCUMENT  EMISSIONS REVIEW ... 5-3
    5.3   SOURCE CONTRIBUTIONS TO SUSPENDED  PARTICLES	5-5
    5.4   NATIONAL EMISSION RATES AND ANNUAL TRENDS	5-11
    5.5   EMISSIONS PROCESSES AND ESTIMATION METHODS	5-19
          5.5.1   Fugitive Dust	5-19
          5.5.2   Mobile Source Emissions  	5-28
    5.6   SIZES DISTRIBUTIONS OF PRIMARY PARTICLE
          EMISSIONS	5-31
    5.7   CHEMICAL COMPOSITIONS OF PRIMARY PARTICLE
          EMISSIONS	5-35
    5.8   EMISSIONS MODELS AND EMISSIONS INVENTORIES	5-42
    5.9   SUMMARY AND CONCLUSIONS	5-44
    REFERENCES	5-48

6.  ENVIRONMENTAL CONCENTRATIONS	6-1
    6.1    BACKGROUND, PURPOSE, AND SCOPE	6-1
          6.1.1   Dimensionality and Structuring of the Aerosol Data Space	6-2
          6.1.2   Spatial Pattern and Scales	6-4

April 1995                            I_ix      DRAFT-DO NOT QUOTE OR CITE

-------
                          TABLE OF CONTENTS (cont'd)
          6.1.3    Temporal Pattern and Scales  	6-4
          6.1.4    Space-Time Relationships  	6-5
          6.1.5    Particle Size Distribution  	6-6
          6.1.6    Aerosol Chemical Composition	6-8
          6.1.7    Chapter Organization and Approach	6-10
    6.2    CONTINENTAL AND GLOBAL AEROSOL PATTERNS	6-11
    6.3    U.S. NATIONAL AEROSOL PATTERN AND TRENDS	6-18
          6.3.1    Non-urban National Aerosol Pattern	6-18
                  6.3.1.1 Non-urban PM2 5 Mass Concentrations  	6-18
                  6.3.1.2 Non-urban PM Coarse  Concentrations	6-18
                  6.3.1.3 Non-urban PM10 Mass  Concentrations	6-21
                  6.3.1.4 PM25/PM10 Ratio at Non-urban Sites  	6-21
                  6.3.1.5 Non-urban Fine Particle Chemistry	6-21
                  6.3.1.6 Seasonally of the Non-urban Chemistry	6-27
          6.3.2    Urban National Aerosol Pattern—AIRS	6-34
                  6.3.2.1 National Pattern and Trend of AIRS PM10  	6-37
                  6.3.2.2 Eastern U.S. PM10 Pattern and Trend	6-39
                  6.3.2.3 Western U.S. PM10 Pattern and Trend	6-41
                  6.3.2.4 Short-term Variability of PM10 Concentrations	6-45
                  6.3.2.5 AIRS PM2 5  Concentrations	6-45
                  6.3.2.6 Other National Surveys	6-48
          6.3.3    Comparison of Urban and Non-urban Concentrations  	6-51
    6.4    Regional Patterns and Trends  	6-55
          6.4.1    Regional Aerosol Pattern in the Northeast	6-56
                  6.4.1.1 Non-urban Size and Chemical Composition in the
                          Northeast	6-57
                  6.4.1.2 Urban Aerosols in the Northeast  	6-57
          6.4.2    Regional Aerosol Pattern in the Southeast  	6-61
                  6.4.2.1 Non-urban Size and Chemical Composition in the
                          Southeast	6-61
                  6.4.2.2 Urban Aerosols in the Southeast	6-63
          6.4.3    Regional Aerosol Pattern in the Industrial Midwest	6-66
                  6.4.3.1 Non-urban Size and Chemical Composition in the
                          Industrial Midwest	6-66
                  6.4.3.2 Urban Aerosols in the Industrial Midwest	6-69
          6.4.4   Regional Aerosol Pattern in the Upper Midwest  	6-71
                  6.4.4.1 Non-urban Size and Chemical Composition in the Upper
                          Midwest 	6-73
                  6.4.4.2 Urban Aerosols in the  Upper Midwest	6-73
          6.4.5   Regional Aerosol Pattern in the Southwest	6-75
                  6.4.5.1 Non-urban Size and Chemical Composition in the
                          Southwest  	6-79
                  6.4.5.2 Urban Aerosols in the  Southwest  	6-79
          6.4.6   Regional Aerosol Pattern in the Northwest	6-80

April 1995                               I-x        DRAFT-DO NOT QUOTE OR CITE

-------
                          TABLE OF CONTENTS (cont'd)
                   6.4.6.1 Non-urban Size and Chemical Composition in the
                          Northwest  	6-84
                   6.4.6.2 Urban Aerosols in the Northwest	6-84
           6.4.7    Regional Aerosol Pattern in Southern California  	6-88
                   6.4.7.1 Non-urban Size and Chemical Composition in
                          Southern California	6-88
                   6.4.7.2 Urban Aerosols in Southern California	6-90
    6.5    SUBREGIONAL AEROSOL PATTERNS AND TRENDS  	6-93
           6.5.1    Subregional Aerosol Pattern in the Northeast  	6-93
                   6.5.1.1 Shenandoah National Park  	6-93
                   6.5.1.2 Washington, DC   	6-95
                   6.5.1.3 Comparison of Non-urban (Shenandoah) to Urban
                          (Washington, DC) Aerosols  	6-96
                   6.5.1.4 New York City, NY	6-100
                   6.5.1.5 Philadelphia, PA	6-105
                   6.5.1.6 Whiteface Mountain, NY	6-108
           6.5.2    Subregional Aerosol Pattern in the Southeast  	6-109
                   6.5.2.1 Winston-Salem, NC, and Florida	6-109
                   6.5.2.2 Large Southeast Metropolitan Areas  	6-111
                   6.5.2.3 Great Smoky Mountains	6-112
           6.5.3    Subregional Aerosol Pattern in the Industrial Midwest	6-112
                   6.5.3.1 Pittsburgh, PA  	6-112
                   6.5.3.2 St. Louis, MO  	6-117
                   6.5.3.3 Chicago, IL	6-119
                   6.5.3.4 Detroit, MI  	6-122
           6.5.5    Subregional Aerosol Pattern in the Southwest	6-123
                   6.5.5.1 El Paso, TX	6-123
                   6.5.5.2 Phoenix and Tucson, AZ	6-123
           6.5.6    Subregional Aerosol Pattern in the Northwest	6-128
                   6.5.6.1 South Lake Tahoe 	6-128
                   6.5.6.2 Salt Lake City, UT, Subregion  	6-130
                   6.5.6.3 Denver, CO	6-132
                   6.5.6.4 Northern Idaho-Western Montana Subregion	6-132
                   6.5.6.5 Washington-Oregon Subregion	6-135
           6.5.7    Subregional Aerosol Pattern in Southern California  	6-137
                   6.5.7.1 San Joaquin Basin 	6-137
                   6.5.7.2 Los Angeles-South Coast Air Basin	6-140
6.6 CHEMICAL COMPOSITION OF PM AEROSOLS AT URBAN AND
    NON-URBAN SITES	6-147
6.7 ACID AEROSOLS  	6-152
           6.7.1    Introduction	  6-152
           6.7.2    Geographical Distribution	  6-153
           6.7.3    Spatial Variation (City-Scale)	6-155
           6.7.4    Spatial Variation (Regional-Scale)	6-156

April 1995                              I-xi       DRAFT-DO NOT QUOTE OR CITE

-------
                         TABLE OF CONTENTS (cont'd)
          6.7.5   Seasonal Variation  	6-157
          6.7.6   Diurnal Variation	6-157
          6.7.7   Indoor and Personal	6-160
    6.8    PARTICLE NUMBER CONCENTRATION  	6-160
          6.8.1   Introduction  	6-160
          6.8.2   Ultrafine Particle Number-Size Distribution	6-161
          6.8.3   Relation of Particle Number to Particle Mass	6-164
          6.8.4   Conclusion	6-169
    6.9    AMBIENT CONCENTRATIONS OF ULTRA-FINE METALS  	6-169
          6.9.1   Introduction  	6-169
          6.9.2   Formation of Ultra-Fine Particles  	6-170
          6.9.3   Techniques for Collecting and Analyzing Ultra-Fine Metals  . .  . 6-172
          6.9.4   Observations of Ultra-Fine Metals; Stack and Source-Enriched
                 Aerosols	6-176
    6.10   SUMMARY	6-185
    REFERENCES	6-190

    APPENDIX 6A:  TABLES OF CHEMICAL COMPOSITION OF PM  	6A-1

7.   HUMAN EXPOSURE TO PARTICULATE MATTER AMBIENT AND
    INDOOR CONCENTRATIONS  	7-1
    7.1    INTRODUCTION  	7-1
          7.1.1   Ambient PM Concentration as a Surrogate for PM Dosage	7-2
          7.1.2   General Concepts for Understanding PM Exposure and
                 Microenvironments	7-4
          7.1.3   Review of State-of-Knowledge Recorded in the 1982
                 PM-SOX AQCD	7-8
    7.2    DIRECT METHODS OF MEASUREMENT OF HUMAN EXPOSURE
          TO PM BY PERSONAL MONITORING	7-11
          7.2.1   Personal Monitoring Artifacts	7-11
                 7.2.1.1 "The Hawthorne Effect"  	7-11
                 7.2.1.2 The Monitor Effect  	7-12
                 7.2.1.3 Subject Effect	7-12
                 7.2.1.4 Non-Response Error	7-12
          7.2.2   Characterization of PM Collected by Personal Monitors	7-13
          7.2.3   Microscale Variation and the Personal Cloud Effect	7-13
    7.3    NEW LITERATURE  ON PARTICLE EXPOSURES SINCE 1981  	7-14
          7.3.1   Review of the Literature	7-16
                 7.3.1.1 Results  of U.S. Studies	7-16
          7.3.2   Personal Exposures in International Studies  	7-21
          7.3.3   The Particle TEAM Study	7-22
                 7.3.3.1 Pilot Study	7-23
                 7.3.3.2 Main Study  	7-27
          7.3.4   Personal Exposures to Constituents of PM	7-36

April 1995                            I-xii      DRAFT-DO NOT QUOTE OR CITE

-------
                          TABLE OF CONTENTS (cont'd)
                                                                            "Jage
    7.4   INDIRECT MEASURES OF EXPOSURE  	7-39
          7.4.1   Personal Exposure Models Using Time-Weighted-Averages
                  of Indoor and Outdoor Concentrations   	7-41
    7.5   DISCUSSION	7-46
          7.5.1   Relation of Individual Exposures to Ambient
                  Concentration  	7-46
          7.5.2   Relation of Community Exposures to Ambient Concentrations . .  . 7-51
          7.5.3   Implications for PM and Mortality Modeling  	7-57
          7.5.4   Relative Toxicity  of Ambient PM and Indoor PM  	7-59
          7.5.5   Conclusions	7-62
    7.6   INDOOR CONCENTRATIONS AND SOURCES OF PARTICULATE
          MATTER  	7-63
          7.6.1   Introduction   	7-63
          7.6.2   Concentrations of Particles in Homes and Buildings	7-64
                  7.6.2.1  Concentrations in Homes	7-65
                  7.6.2.2  Studies in Buildings 	7-107
                  7.6.2.3  Studies in Locations Other Than Homes and Buildings .  7-114
          7.6.3   Indoor Air Quality Models and Supporting Experiments	7-114
                  7.6.3.1  Mass Balance Models	7-114
                  7.6.3.2  Models of ETS	7-122
          7.6.4   Summary and Conclusions	7-126
    REFERENCES	7-129
April 1995                             I-xiii      DRAFT-DO NOT QUOTE OR CITE

-------
                                 LIST OF TABLES
Number                                                                         Page

1-1       Characterization of Urban PM10 Data from AIRs Network by Region
          for the United States	1-9

1-2       PM10 Levels by Annual Average for Selected U.S. SMSAs for  1993	1-10

1-3       Selected U.S. PM10 Levels by Second Max PM10 for 1993	1-11

1-4       Concentration Ranges of Various Elements Associated with
          Paniculate Matter in the United States Atmosphere	1-12

1-5       Comparison of Relative Risk Estimates for Total Mortality
          from 50 jug/m3 Change in PM10	1-37

1-6       Prospective Cohort Mortality Studies	1-48

1-7       Hospital Admissions Studies for Respiratory Disease	1-51

1-8       Hospital Admissions Studies for COPD	1-52

1-9       Hospital Admissions Studies for Pneumonia  	1-53

1-10      Hospital Admissions Studies for Heat Disease	1-53

1-11      Acute Respiratory Disease Studies   	1-55

1-12      Chronic Respiratory Disease Studies	1-59

1-13      Acute and Chronic Pulmonary Function Changes  	1-61

1-14      Estimated Excess Mortality per Day In a Population of One Million
          For Which An  Increase of 50 /zg/m3 PM10 (24-H) Could Be a
          Contributing Factor  	1-95

1-15      Estimated Number of Deaths per Day In Cities of Ten Thousand
          To Ten Million for Which An Increase of 50 />ig/m3 PM10
          Could Be a Contributing Factor	1-95

1-16      Association Between Cigarette Smoking Status and Excess Mortality Risk
          from Air Pollution In the Six Cities Study  	1-96

1-17      Estimated Hospital Admissions per Day In a Population of
          One Million for Which An Increase of 50  jug/m3 (24-H) PM10
          Could Be a Contributing Factor	1-98


April 1995                             I_xiv      DRAFT-DO NOT QUOTE OR CITE

-------
                              LIST OF TABLES (cont'd)
Number                                                                        Page

1-18      Estimated Number of Hospital Admissions for Respiratory and
          Cardiovascular Causes per Day In Cities of Ten Thousand to
          Ten Million for Which An Increased of 50 ^g/m3 PM10 (24-H)
          Could Be a Contributing Factor	1-98

3-1       Lognormal Parameters for Ambient Aerosols	3-21

3-2       Values of Log P£ for Various PAHs at 20 °C	3-46

3-3       mp Values for PAHs Sorbing to UPM in Osaka, Japan	3-47

3-4       Effects of Three Types of Artifacts on Volume-Averaged Values of 
          Measured Using a Filter/Adsorbent Sampler	3-48

3-5       Some Secondary Organic Compounds Identified in Ambient Particles in
          Urban Air	3-64

3-6       Predicted Percent Contribution to Secondary Organic Aerosol
          Concentrations at Los Angeles	3-65

3-7       Amount  of Secondary Aerosol  Produced in a  Typical Los Angeles Smog
          Episode  According to Functional Groups 	3-65

3-8       Reactivity Scale for the Electrophilic Reactions of PAH	3-68

3-9       Concentration Ranges of Various Elements Associated  with Paniculate
          Matter in the Atmosphere   	3-70

3-10      Compounds Observed in Aerosols by a Roadway at Argonne National
          Laboratory	3-72

3-11      Compounds Observed in Aerosols in a Forested Area,  State
          College, PA	3-72

3-12      Henry's  Law Coefficients of Some Atmospheric Gases Dissolving in
          Liquid Water	3-77

3-13      Recent Field Studies of a-Mesoscale Transport and Trajectory Model  ....  3-93

3-14      Scavenging Ratios  	3-124
3-15      Relative  Humidity of Deliquescence and Crystallization for Several
          Atmospheric Salts	  3-153

3-16      Summary of Hygroscopic Growth  Factors	3-157

April 1995                              I-xv       DRAFT-DO NOT QUOTE OR CITE

-------
                             LIST OF TABLES (cont'd)
Number                                                                       Page

3-17     Comparison of Sulfate Concentration and Mass Mean Diameters of
         Aerosols for Days with Higher and Lower Relative Humidity  	3-163

4-1      EPA-designated Reference and Equivalent Methods for PM10	4-41

4-2      Instrumental Detection Limits for Particles on Filters	4-65

4-3      Minimum Detectable Limits for XRF Analysis of Air Filters	4-73

4-4      INAA Counting Scheme and Elements Measured  	4-81

5-1      Receptor Model Source Contributions to PM10	5-7

5-2      Typical Chemical Abundances in Source Emissions	5-36

6-1      Aerosol Information Needs for Assessing Effects  	6-2

6-2      Spatial Regions and Scales	6-4

6-3      Maximum SO^" and H+ Concentrations Measured in North American
         Cities	6-154

6-4      Regulated Metals and the Volatility Temperature  	6-171

6-5      Composition of the Aerosols Present at Grand Canyon National Park
         in the Summer of 1984, for the Two Sulfate Episodes of
         August 15 and 16  	6-179

6-6      Measurements of fine and ultra-fine metals	6-180

6-7      Measurements of fine and ultra-fine metals—lead and nickel  	6-183

6-8      Comparison of Selected Species, Shenandoah National Park, and
         Washington, DC, San Gorgonio Wilderness, CA, and Grand Canyon
         National Park, Summer,  1993	6-187

7-1      Quantile Description of Personal, Indoor, and Outdoor PM3 5
         Concentrations by Location in Two Tennessee  Communities  	7-17

7-2      Summary of WHO/UNEP GEMS/HEAL PM,  Personal Exposure Pilot
         Study Results	7-22

7-3      Summary of Daily Indoor,  Outdoor, and Personal Exposures to PM10
         During Cooking as a Function of Fuel Type in Three Cities in Asia	7-23

April 1995                              I-xvi      DRAFT-DO NOT QUOTE OR CITE

-------
                              LIST OF TABLES (cont'd)
Number

7-4       PTEAM Prepilot Study:  Mean Outdoor Particle Concentrations (/*g/m3) .  . .  7-25

7-5       PTEAM Prepilot Study:  24-Hour Particle Concentrations	7-25

7-6a      PTEAM Prepilot Study:  24-Hour PM10 Concentrations	   7-26

7-6b      PTEAM Prepilot Study:  24-Hour PM2 5 Concentrations	7-27

7-7       Population-Weighted Concentrations and Standard Errors  	7-30

7-8       Comparison of PEM Exposure of Individuals to the Simultaneous Ambient
          PM Concentration in 10 U.S. Cities and Four Foreign Cities	7-35

7-9       Concentrations of Particles (PM2 5) in Homes of Children Participating
          in the Harvard Six-City Study	7-67

7-10a     Reconstructed Source Contributions to PM2 5 Mass for Steubenville	7-74

7-1 Ob     Reconstructed Source Contributions to PM2 5 Mass for Portage   	7-74

7-11      Weighted Summary Statistics by County for Respirable Suspended
          Paniculate (PM2 5) Concentrations	7-75

7-12      Weighted Analysis of Variance of Respirable Suspended Particulate
          (PM2 5) Concentrations in the Main Living Area of Homes
          Versus Source Classification	7-76

7-13      Respirable Suspended Particulate Concentration	7-76

7-14      Weighted Distributions of Personal, Indoor, and Outdoor Particle
          Concentrations  	7-79

7-15      Weighted Distributions of PM2 5/PM10 Concentration Ratio	7-80

7-16      Stepwise Regression Results for Indoor Air Concentrations of PM10,
          PM25, and Nicotine:   Coefficients  (Standard Errors of Estimates)	7-84
7-17      Penetration Factors, Decay Rates, and Source Strengths:  Nonlinear
          Estimates	7-87

7-18      Indoor-Outdoor Mean Concentrations of Fine Particles in
          Three Large-Scale Studies	7-93

7-19      Influence of Recent Cigarette Smoking on Indoor Concentrations of
          PM (Size Unspecified)  	7-98

April 1995                              I_xvii      DRAFT-DO NOT QUOTE OR CITE

-------
                           LIST OF TABLES (cont'd)
Number                                                                   Page

7-20     Indoor and Outdoor PM in Buildings in Helsinki, Finland, as a Function
         of Season and Location	7-99

7-21     Indoor Average PM2 5 and PM10 by Reported Smoking in the
         Home and Evaporative Cooler Use During Sampling Week	7-100

7-22     Indoor PM10 and PM2 5 by Season and Environmental Tobacco
         Smoke  	'	7-101

7-23     Smoking, Nonsmoking, and Outdoor RSP Concentrations and Ratios ....  7-108
 April 1995                            I-xviii     DRAFT-DO NOT QUOTE OR CITE

-------
                                   LIST OF FIGURES
Number                                                                           Page

3-1       Particle size related to RH	3-5

3-2       Number of particles as a function of particle diameter	3-7

3-3       Particle volume distribution as a function of particle diameter   	3-7

3-4       Particle volume distribution as a function of particle diameter in a
          freeway-influenced area .	3-8

3-5       In concentration as a function of particle size   	3-10

3-6       Efficiency values for size-selective sampling criteria 	3-16

3-7       Model dust emissions for  the United States	3-30

3-8       Diffusion constants and settling velocities for particles	3-31

3-9       Particle deposition from wind tunnel studies  	3-32

3-10      Sedimentation and inertia  effects on large particle deposition	3-32

3-11      Comparison of observed H2O2 depletions and observed
          sulfate yields   	3-41

3-12      Extrapolations from correlations of wind tunnel measured deposition
          velocities for z =  1 m, densities of 1, 4, and 11.5 g cm"3	3-117

3-13      An example of histogram  display and fitting to log-normal functions for
          particle-counting size distribution data	3-132

3-14      An example of an effective display of impactor data 	3-133

3-15      Size distributions of sulfate, Long Beach, June, 1987,  showing use of
          fitted log-normal distributions to describe diurnal variations in size
          and concentration	3-134

3-16      Effect of changing endpoints	3-135

3-17      These size distributions, obtained during an EPA study of the Denver
          brown cloud represent one of the few efforts to compare particle-counting
          and particle-collection size-distribution measurements	3-136
April 1995                               I_xix      DRAFT-DO NOT QUOTE OR CITE

-------
                              LIST OF FIGURES (cont'd)
Number

3-18      Grand average volume-size distributions from the Aerosol Characterization
          Experiment in 1972  	3-139

3-19      Volume-size distribution taken in the midwestern U.S. near the
          Cumberland Power Plant in Tennessee	3-140

3-20      Examples of size distribution histograms for total mass, sulfate, and
          iron obtained at two visibility levels using an Anderson impactor   	3-141

3-21      Impactor size distribution measurement generated by Lundgren et al.
          with the Wide Range Aerosol Classifier: Philadelphia and Phoenix	3-143

3-22      Example of aged and fresh coarse mode particle size distributions	3-144

3-23      Size distributions reported by Noll from the Chicago area using an
          Anderson impactor for the smaller particles and a Noll Rotary Impactor
          for the larger particles	3-145

3-24      Size distribution of dust generated by driving a truck over an unpaved
          test track	3-147

3-25      Size distribution of emissions from a pulverized-coal power plant and
          the particle size distributions remaining after several types of control
          devices	3-149

3-26      Size distributions from a fluidized-bed, pulverized coal combustor,
          after initial cleanup by a cyclone collector and after final cleanup
          by a baghouse	3-150

3-27      Particle growth curves showing  fully reversible hygroscopic growth
          of sulfuric acid particles  	3-152

3-28      Theoretical predictions and experimental measurements of growth
          of NH4HSO4 and (NH4)2SO4 particles at RH between 95 and 100%   .... 3-154

3-29      Tandem Differential Mobility Analyzer measurements of the sensitivity
          of particle size to relative humidity at  Claremont, CA	3-156

3-30      Example of growth in particle size due primarily to increases in
          relative humidity from Uniontown, PA  	3-159

3-31      Mass size distribution of non-volatile aerosol material  	3-160
April 1995                                I-xx       DRAFT-DO NOT QUOTE OR CITE

-------
                              LIST OF FIGURES (cont'd)
Number                                                                          Page

3-32      Example of particle-counting volume distribution obtained in
          Claremont, CA	3-162

3-33      Relative humidity versus sulfur, during the 1986 Carbonaceous
          Species Methods Comparison Study, for particles with Dp>0.56 /mi  ....  3-165

3-34      Data from the South Coast Air Quality Study  	3-166

3-35      Log-log plot of sulfate mode concentration versus mode diameter
          from Claremont during summer SCAQS (John et al., 1990)	3-167

3-36      Typical results of size-distribution measurements taken with a Berner
          impactor in a Vienna street with heavy automotive traffic (Berner et al.,
          1993)	3-168

4-1       Characteristics of aerosol measurement instruments	4-3

4-2       American Conference of Governmental Industrial Hygienists,
          British Medical Research Council, and International Organization
          for Standardization size-selective sampling criteria  	4-7

4-3       Sampling efficiency of IOM ambient inhalable aerosol sampler for three
          different types of test aerosol   	4-9

4-4       Liquid particle sampling effectiveness curves with solid particle
          points superimposed for the Wedding IP10 and the Andersen Samplers
          Model 321A inlets at 8 km/h  	4-14

4-5       Andersen sampler 	4-15

4-6       Sampling characteristics of two-stage size-selective inlet for liquid
          aerosols  	4-17

4-7       Penetration of particles for 16.67 1 pm PM10 inlets	4-18

4-8       Collection performance variability as a function of wind speed	4-19

4-9       Calibration of a 2.5 /xm impactor	4-22

4-10      Percent collection as  a function of aerodynamic diameter  	4-24

4-11      Performance of glass fiber filters compared to greased substrate	4-27

4-12      Schematic diagram of an annular denuder system	4-32

April 1995                               I-xxi       DRAFT-DO NOT QUOTE OR CITE

-------
                              LIST OF FIGURES (cont'd)
Number

4-13      Aerosol calibration of a cascade impactor	4-44

4-14      TEOM	4-49

4-15      Beta gauge	4-51

4-16      Integrating nephelometer	4-53

4-17      Particle-scattering coefficient as a function of particle size	4-54

4-18      Correlation of bsn and fine fraction mass  	4-56
                        SH

4-19      Collection efficiency of the MSP personal aerosol sampler  	4-59

4-20      Modified dichotomous sampler  	4-62

4-21      Schematic of a typical X-ray fluorescence system	4-72

4-22      Example of an X-ray fluorescence spectrum  	4-75

4-23      Schematic of a PIXE/PESA analysis system  	4-79

4-24      Typical gamma-ray spectra observed for long counts	4-82

4-25      Typical gamma-ray spectra observed for short counts	4-83

4-26      Schematic representation of an ion chromatography system	4-88

4-27      Example of an ion chromatogram showing the separation of fluoride,
          chloride, nitrite,  nitrate, phosphate, and sulfate ions  	4-89

4-28      Schematic of a typical automated colorimetric system	4-91

5-1       Primary PM10 emissions estimated for 1983 to 1992	5-12

5-2       Sub-categories of non-fugitive dust emissions, 1983 to 1992   	5-13

5-3       National emissions of sulfur dioxide,  1983 to 1992	5-14

5-4       National emissions for oxides of nitrogen, 1983 to 1992	5-15

5-5       National emissions for volatile organic compounds, 1983 to 1992	5-16

5-6       Size distribution of California source emissions,  1986  	5-32

April 1995                               I-xxii      DRAFT-DO NOT QUOTE OR CITE

-------
                              LIST OF FIGURES (cont'd)
Number                                                                          Page

5-7       Size distribution of California particle emissions, 1986	5-33

5-8       Particle size distribution in laboratory resuspension chamber	5-34

5-9       Chemical abundances for PM2 5 profiles of road dust	5-37

5-10      Chemical abundances for PM2 5 profiles of vehicle exhaust	5-37

5-11      Chemical abundances for PM25 profiles of wood burning	5-38

5-12      Chemical abundances for PM25 profiles of coal-fired power plant	5-38

6-1       Time scales for particle emissions  	6-5

6-2       Relationship of spatial and temporal scales for coarse and fine
          Particles	6-7

6-3       Residence time in the lower troposphere for atmospheric particles
          between 0.1 and 1.0 pirn  	6-8

6-4       Space-time relationship in urban and mountainous areas  	6-9

6-5a      Continental scale pattern of aerosols derived from visibility
          observations over land and satellite monitoring over the
          oceans: eastern North America	6-12

6-5b      Continental scale pattern of aerosols derived from visibility
          observations over land and satellite monitoring over the
          oceans: western North America   	6-13

6-5c      Continental scale pattern of aerosols derived from visibility
          observations over land and satellite monitoring over the
          oceans: southern North America	6-14

6-6       Global pattern of oceanic aerosols derived from satellite observations	6-15

6-7       Seasonal pattern of oceanic aerosols derived from satellite
          observations	6-17

6-8       Fine mass concentration derived from non-urban IMPROVE/NESCAUM
          networks	 6-19

6-9       Coarse mass concentration derived from non-urban IMPROVE/NESCAUM
          networks	6-20

April 1995                              I-xxiii     DRAFT-DO NOT QUOTE OR CITE

-------
                             LIST OF FIGURES (cont'd)
Number                                                                         Page

6-10      PM10 mass concentration derived from non-urban IMPROVE/NESCAUM
          networks	6-22

6-11      Fine fraction of pm10 derived from non-urban IMPROVE/NESCAUM
          networks	6-23

6-12      Chemical mass balance of fine particles derived from non-urban
          IMPROVE/NESCAUM networks	6-24

6-13a,b   Yearly average absolute and  relative concentrations for sulfate,
          nitrate,  organics, and soot  	6-25

6-13c,d   Yearly average absolute and  relative concentrations for sulfate, nitrate,
          organics,  and soot	6-26

6-14      Seasonal pattern of non-urban aerosol concentrations for the entire
          United States  	6-28

6-15      Seasonal pattern of non-urban aerosol concentrations for the eastern
          United States  	6-31

6-16      Seasonal pattern of non-urban aerosol concentrations for the western
          United States  	6-33

6-17      Trend of  valid  pm10 monitoring stations in the AIRS database  	6-35

6-18      AIRS PM10 quarterly concentration maps using all available data   	6-36

6-19      AIRS PM10 and PM2 5 concentration pattern for the conterminous
          United States  . . .	6-38

6-20      AIRS PM10 and PM2 5 concentration pattern for east of the Rockies	6-40

6-21      AIRS PM10 and PM2 5 concentration pattern for west of the Rockies	6-42

6-22      Short-term PM10 concentration time  series for Missoula, MT, and
          Knoxville, TN 	6-46

6-23      Logarithmic  standard deviation airs PM10 concentrations	6-47

6-24      Annual PM2.5 concentration  pattern obtained from IMPROVE/NESCAUM
          and AIRS networks	6-49
April 1995                              I-xxiv     DRAFT-DO NOT QUOTE OR CITE

-------
                             LIST OF FIGURES (cont'd)
Number                                                                       Page

6-25      Monthly mean concentrations in /ig/m3 of IP, fine fraction, and S
          as (NH4)2SO4 in Portage, WI; Topeka, KS; Harriman, TN; Watertown, MA;
          St. Louis, MO; and Steubenville, OH	6-50

6-26      Spatial maps  of PM10 concentration difference between AIRS and
          IMPROVE/NESCAUM networks	6-52

6-27      Urban excess concentrations (AIRS minus IMPROVE) for the United
          States, eastern United States, and western United States	6-54

6-28      Aerosol regions of the conterminous United States  	6-56

6-29      IMPROVE/NESCAUM concentration data for the Northeast	6-58

6-30      AIRS concentration data for the Northeast  	6-60

6-31      Short-term variation of PM10 average for the Northeast  	6-62

6-32      Urban excess concentration (AIRS minus IMPROVE) for the Northeast  .  . . 6-62

6-33      IMPROVE/NESCAUM concentration data for the Southeast  	6-64

6-34      AIRS concentration data for the Southeast  	6-65

6-35      Short-term variation of PMj0 average for the Southeast  	6-67

6-36      Urban excess concentration (AIRS minus IMPROVE) for the Southeast  .  . . 6-67

6-37      IMPROVE/NESCAUM concentration data for the industrial Midwest	6-68

6-38      AIRS concentration data for the industrial Midwest	6-70

6-39      Short-term variation of PM10 average for the industrial Midwest	6-72

6-40      Urban excess concentration (AIRS minus IMPROVE) for the industrial
          Midwest	6-72

6-41      IMPROVE/NESCAUM concentration data for the Upper Midwest  	6-74

6-42      AIRS concentration data for the Upper Midwest	6-76

6-43      Short-term variation of PM10 average for the Upper Midwest	6-77



April 1995                             I-xxv      DRAFT-DO NOT QUOTE OR CITE

-------
                            LIST OF FIGURES (cont'd)
Number                                                                      Page

6-44     Urban excess concentration (AIRS minus IMPROVE) for the Upper
         Midwest	6-77

6-45     IMPROVE/NESCAUM concentration data for the Southwest	6-78

6-46     AIRS concentration data for the Southwest	6-81

6-47     Short-term variation of PMj0 average for the Southwest	6-82

6-48     Urban excess concentration (AIRS minus IMPROVE) for the
         Southwest	6-82

6-49     IMPROVE/NESCAUM concentration data for the Northwest	6-83

6-50     AIRS concentration data for the Northwest	6-86

6-51     Short-term variation of PM10 average for the Northwest	6-87

6-52     Urban excess concentration (AIRS minus IMPROVE) for the
         Northwest	6-87

6-53     IMPROVE/NESCAUM concentration for  Southern California  	6-89

6-54     AIRS concentration for Southern California	6-91

6-55     Short-term variation of PM10 average for Southern California  	6-92

6-56     Urban excess concentration (AIRS minus IMPROVE) for Southern
         California  	6-92

6-57     IMPROVE/NESCAUM concentration for  Shenandoah National Park	6-94

6-58     IMPROVE/NESCAUM concentration for  Washington, DC	6-97

6-59     Excess aerosol concentration at Washington, DC compared to
         Shenandoah National Park	6-99

6-60     Daily concentration of fine mass and fine  sulfur at Washington, DC
         and Shenandoah National Park  	6-101

6-61     Aerosol concentration map, trend and seasonality in the New York City
         region	6-102

6-62     Fine, coarse, and PM10 particle concentration near New York City	6-103

April 1995                             I-xxvi     DRAFT-DO NOT QUOTE OR CITE

-------
                              LIST OF FIGURES (cont'd)
Number                                                                          Page

6-63      Aerosol concentration map, trend and seasonality in the Philadelphia
          region  	6-106

6-64      Fine, coarse, and PM10 particle concentration near Philadelphia	6-107

6-65      PM10 concentration seasonality at Whiteface Mountain and neighboring
          low elevation sites	6-109

6-66      Aerosol concentration pattern at North Carolina and Florida sites  ......  6-110

6-67      Aerosol concentration pattern in Texas and Gulf states	6-113

6-68      Aerosol concentration pattern and trends in the Pittsburgh subregion  ....  6-114

6-69      Fine, coarse, and PMIO concentration near Pittsburgh  	6-116

6-70      Aerosol concentration pattern and trends in the St. Louis subregion	6-118

6-71      Fine, coarse, and PM10 concentration pattern near St. Louis  	6-120

6-72      Aerosol concentration pattern and trends in the Chicago subregion	6-121

6-73      Aerosol concentration pattern and trends in the El Paso subregion	6-124

6-74      Fine, coarse, and PM10 concentration pattern near El Paso  	6-125

6-75      Aerosol concentration pattern and trends in the Phoenix-Tucson
          subregion  	6-126

6-76      Excess aerosol concentration at South Lake Tahoe compared to Bliss
          State Park	6-129

6-77      Aerosol concentration pattern near Salt Lake City	6-131

6-78      Aerosol concentration pattern and trends in the Northern
          Idaho-Northwestern Montana subregion	6-133

6-79      PM10 concentration pattern at sites in Northern Idaho-Northwestern
          Montana subregion	6-134

6-80      Aerosol concentration pattern in Washington State and Oregon	6-136

6-81      Aerosol concentration pattern and trends in the San Joaquin Valley  	6-138


April  1995                              I-XXVii      DRAFT-DO NOT QUOTE OR  CITE

-------
                             LIST OF FIGURES (cont'd)
Number                                                                         Page

6-82      Fine, coarse, and PM10 concentration pattern in the
          San Joaquin Valley	6-139

6-83      Aerosol concentration pattern and trends at Los Angeles	6-141

6-84      Fine, coarse, and PM10 concentration pattern near Los Angeles  	6-142

6-85a     Mass apportionment:  eastern United States	6-149

6-85b     Mass apportionment:  central United States	6-150

6-85c     Mass apportionment:  western United States  	6-151

6-86      Mean air pollutant concentrations for days when winds were from
          the southerly direction, plotted versus population density  	6-156

6-87      Average monthly aerosol strong acidity for Year 1 sites of the Harvard
          24-city study  	6-158

6-88      Diurnal Pattern of sulfate and hydrogen ion at Harriman,  TN	6-159

6-89      Aerosol number and volume size distributions from an urban site
          at Long Beach, CA  	6-162

6-90      Aerosol number and volume size distributions from a background site
          in the Rocky Mountains,  CO  	6-163

6-91      Number concentrations as a  function of time of day at Long Beach,
          CA  	6-164

6-92      Number and volume size distributions  at the Rocky Mountain site
          showing an intrusion of urban air	6-165

6-93      Number and volume size distributions  from Los  Angeles, CA,
          showing comparison of three measurement techniques  	6-166

6-94      Relationship between particle number and particle volume	6-168

6-95      Impact of treatment temperature on the enrichment of metals
          in the fly ash after the thermal treatment of soils from a
          Superfund site	 6-173
April 1995                             I-xxviii      DRAFT-DO NOT QUOTE OR CITE

-------
                              LIST OF FIGURES (cont'd)
Number                                                                         Page

6-96      Average normalized concentrations as a function of stage number,
          for Se, S, Ca, Al, Si, K, Mo, W, Ni, and Cr for five BLPI samples
          from a coal-fired power plant	6-177

6-97      Fine and ultra-fine sulfur at Grand Canyon National Park,
          Summer, 1984  	6-178

6-98      Fine and ultra-fine metals, nickel, selenium, and lead, in
          Long Beach, CA, December  10 through 13, 1987, in
          four-hour increments  	6-182

6-99      Patterns of zinc,  arsenic, sulfur, and selenium in the United States  	6-186

6-100     Apparent deposition of automotive lead aerosol in the respiratory
          tract of one of the authors as determined by cascade impactor and
          PIXIE, as a function of aerodynamic diameter for  > 4, 4 to 2, 2 to 1,
          1 to 0.5, 0.5 to 0.25, and <0.25 jum particles of size classes 1, 2, 3,
          4, 5, and 6, respectively  	6-188

7-1       Sizes of indoor particles   	7-5

7-2       An example of personal exposure to respirable particles	7-10

7-3       Central-site mean of two dichotomous samplers versus residential
          outdoor monitors	7-31

7-4       Personal exposures  versus residential (back yard) outdoor PM10
          concentrations	7-32

7-5       Increased concentrations of elements  in the personal versus the
          indoor  samples  	7-33

7-6       Source  apportionment of PTEAM PM10 Personal Monitoring data	7-34

7-7       Personal versus outdoor SO4=	7-37

7-8       Estimated ("best fit" model) versus measured personal SO4=	7-38

7-9       Personal activity cloud and exposure	7-42

7-10      Components of personal exposure ,„.,..„......	7-47

7-11      Personal exposure to PM in Phillipsburg,  NJ (Winter, 1988)	7-52


April 1995                              I_xxix     DRAFT-DO NOT QUOTE OR  CITE

-------
                             LIST OF FIGURES (cont'd)
Number                                                                         Page

7-12      Personal exposure to PM in Bejing, China (Winter, 1985)	7-53

7-13      Personal exposure to PM in Azusa, CA (Spring, 1989)  	7-54

7-14      Personal exposure to PM in Riverside, CA	7-55

7-15      Personal exposure to PM in Phillipsburg, NH, with concentration
          outliers removed	7-60

7-16      Venn diagram (Mage, 1985) showing focusing of information to more
          completely specify toxicity of a given PM mixture  	7-61

7-17      Percentage of time spent in different microenvironments by U.S.
          residents	7-64

7-18      The annual mean concentration of respirable particles for the
          highest and lowest site from the network of indoor and outdoor monitors
          in each city  	7-67

7-19      Distribution of numbers of children living in households with varying
          respirable paniculate matter (PM2 5) as a function of parental smoking
          status	7-69

7-20      Distribution percentiles for annual average concentrations of indoor
          respirable particulate matter (PM2 5) by household smoking status and
          estimated number  of cigarette packs smoked in the home  	7-71

7-21      PM2 5 in smoking and non-smoking homes in three of the
          Harvard Six-Citys Study sites  	7-73

7-22      Cumulative frequency distribution of 24-h personal, indoor, and outdoor
          PM10 concentrations in Riverside, CA	7-82

7-23      Cumulative frequency distribtuion of 24-h indoor and outdoor PM2 5
          concentrations in Riverside, CA	7-82

7-24      Forty-eight-day sequence of PM10 and PM2  5 in Riverside, CA, PTEAM
          study	7-83

7-25      Average indoor and outdoor 12-h concentrations of PM10 during  the
          PTEAM study in Riverside, CA   	7-83

7-26      Sources of fine particles (PM2 5)  and respirable particles (PM10)
          in all homes (Riverside, CA)   	7-89

April 1995                              I-xxx      DRAFT-DO NOT QUOTE OR CITE

-------
                                   \

                              LIST OF FIGURES (cont'd)
 Number                                                                         Page

 7-27      Sources of fine particles (PM2 5) and respirable particles (PM10)
          in homes with smokers (Riverside, CA)	7-90

 7-28      Sources of fine particles (PM2 5) and respirable particles (PM10)
          for homes with cooking during data collection (Riverside, CA)	7-91

 7-29      Comparison of respirable particles in smoking and nonsmoking areas
          of 38 buildings in the Pacific Northwest	7-109

 7-30      Respirable particles in smoking and nonsmoking areas of homes for
          the elderly (arithmetic mean for 72 hours)  	7-113

 7-31      Predicted fate of particles penetrating into buildings of three
          California museums as  a function of particle size  	7-120
April 1995                              I_xxxi      DRAFT-DO NOT QUOTE OR CITE

-------
                  AUTHORS, CONTRIBUTORS, AND REVIEWERS


                       CHAPTER 1. EXECUTIVE SUMMARY

Principal Authors

Ms. Beverly Comfort-Environmental Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Mr. William Ewald-Environmental Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Lester D.  Grant-Environmental Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Dennis Kotchmar-Environmental Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Ms. Annie M. Jarabek-Environmental  Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. James McGrath-Environmental Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. William Wilson-Atmospheric Research and Exposure Assessment Laboratory (MD-75),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711


Contributors and Reviewers

Dr. Judith Graham-Environmental Criteria and Assessment Office (MD-52), Environmental
Protection Agency, Research Triangle Park, NC 27711

Dr. Jeanette Wiltse, Office of Health and Environmental Assessment, Office of Research and
Development (8601), Waterside Mall,  401 M. St. S.W., Washington, DC  20460


                           CHAPTER 2. INTRODUCTION

Principal Author

Dr. Lester D. Grant-Environmental Criteria and Assessment Office (MD-52), U.S.
Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Dennis Kotchmar-Environmental Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

April 1995                            I-xxxii     DRAFT-DO NOT QUOTE OR CITE

-------
              AUTHORS, CONTRIBUTORS, AND REVIEWERS (cont'd)


      CHAPTER 3.  PHYSICS AND CHEMISTRY OF PARTICULATE MATTER
Principal Authors

Dr. Paul Altshuller-Environmental Criteria and Assessment Office, U.S. Environmental
Protection Agency, Research Triangle Park, NC  27711

Mr. William Ewald-Environmental Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Dale Gillette-Atmospheric Research and Exposure Assessment Laboratory (MD-81),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Suzanne Hering-Aerosol Dynamics, Inc., 2319 Fourth Street, Berkeley, CA 94710

Dr. Paul Lioy-Environmental Occupational Health and Science Institute, Exposure
Measurement and Assessment Division, 681 Frelinghuysen Road, Piscataway,
NJ  08855-1179

Dr. Kenneth Noll-Illinois Institute of Technology, Chicago,  IL 60616

Dr. Spyros Pandis-Carnegie-Mellon University, Pittsburgh, PA  15146

Dr. James Pankow-Oregon Graduate Center, Beaverton, OR 97229-3678

Dr. Steven Schwartz-Brookhaven National Laboratory,  Upton, NY  11934

Dr. William Wilson-Atmospheric Research and Exposure Assessment Laboratory (MD-75),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711


Contributors and Reviewers

Dr. Michael Barnes-Atmospheric Research and Exposure Assessment laboratory (MD-46),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Thomas  Cahill-University of California at Davis, Crocker Nuclear Lab., Davis, CA
95616

Ms. Beverly Tilton-Environmental Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711
April 1995                             I-xxxiii     DRAFT-DO NOT QUOTE OR CITE

-------
              AUTHORS, CONTRIBUTORS, AND REVIEWERS (cont'd)


Contributors and Reviewers (cont'd)

Dr. Sheldon Friedlander-University of California, 5531 Boelter Hall, Los Angeles, CA
90095

Dr. Peter McMurray-University of Minnesota, Department of Mechanical Engineering, 111
Church Street, SE, Minneapolis, MN 55455-0111

Dr. Sidney Soderholm-NIOSH, 1095 Willowdale Rd., Morgantown, WVA  26505

Dr. Barbara Turpin-Rutgers University, Box 231, New Brunswick, NJ  08903
   CHAPTER 4.  SAMPLING AND ANALYSIS OF PARTICULATE MATTER AND
                               ACID DEPOSITION
Principal Authors

Dr. Judith Chow-Desert Research Institute, P.O. Box 60220, Reno, NV 89506-0220

Dr. Steven McDow-University of North Carolina at Chapel Hill, Chapel Hill, NC 27599

Dr. Charles Rodes-Research Triangle Institute, Center for Aerosol Technology,
P.O. Box 12194, Research Triangle Park, NC 27709-2194


Contributors and Reviewers

Dr. Robert Burton-Atmospheric Research and Exposure Assessment Laboratory (MD-56),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Thomas Cahill-University of California - Davis,  Crocker Nuclear Laboratory, Davis,
CA 95616

Dr. Delbert Eatough-Brigham Young University, Department of Chemistry, Provo, UT
84602

Mr. William Ewald-Environmental Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Sheldon Friedlander-University of California, Department of Chemical Engineering,
5531 Boelter Hall, Los Angeles, CA 90095


April 1995                            I-xxxiv      DRAFT-DO NOT QUOTE OR CITE

-------
              AUTHORS, CONTRIBUTORS, AND REVIEWERS (cont'd)
Contributors and Reviewers (cont'd)

Dr. Peter McMurry-University of Minnesota, Department of Mechanical Engineering,
111 Church Street, S.E., Minneapolis, MN 55455-0111

Dr. Sidney Soderholm-National Institute of Occupational Safety and Health,  1095
Williowdale Road, Room 111,  Morgantown, WV 26505

Dr. Barbara Turpin-Rutgers University, Environmental Sciences Building, Box 231,
College Farm Road, New Brunswick, NJ 08903

Dr. Russell Wiener-Atmospheric Research and Exposure Assessment Laboratory (MD-77),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Nancy Wilson-Atmospheric Research and Exposure Assessment Laboratory (MD-56),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711
                     CHAPTER 5. SOURCES AND EMISSIONS


Principal Authors

Dr. Judith Chow-Desert Research Institute, P.O. Box 60220, Reno, NV  89506-0220

Dr. John Watson-Desert Research Institute, P.O. Box 60220, Reno, NV  89506-0220


Contributors and Reviewers

Mr. William Ewald-Environmental Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Timothy Larson-University of Washington, Department of Civil Engineering, Seattle,
WA  98195

Dr. Barbara Turpin-Rutgers University, Environemntal Sciences Building, Box 231, College
Farm Road, New Brunswick, NJ  08903
April 1995                             I-XXXv      DRAFT-DO NOT QUOTE OR CITE

-------
              AUTHORS, CONTRIBUTORS, AND REVIEWERS (cont'd)


               CHAPTER 6. ENVIRONMENTAL CONCENTRATIONS


Principal Authors

Dr. Thomas Cahill-University of California, Davis

Dr. Suzanne Her ing-Aerosol Dynamics, Inc., Berkeley, CA

Dr. Rudolf Husar-Washington University, Center for Air Pollution and Impact and Trend
Analyses, St. Louis, MO  63130

Mr. Joseph Pinto-Atmospheric Research and Exposure Assessment Laboratory (MD-84),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Mr. Robert Stevens-Atmospheric Research and Exposure Assessment Laboratory (MD-47),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Mr. Robert Willis-ManTech Environmental Technology, Inc., P.O. Box 12313, Research
Triangle Park, NC 27709.

Dr. William Wilson-Atmospheric Research and  Exposure Assessment Laboratory (MD-75),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Michael Zelenka-Atmospheric Research and Exposure Assessment Laboratory (MD-75),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711


Contributors and Reviewers

Dr. John Core-WESTAR Council, 1001 SW 5th Avenue, Suite 1100, Portland, OR  97204

Mr. William Ewald-Environmental Criteria and  Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

Dr. Timothy Larson-University of Washington,  Department of Civil Engineering, Seattle,
WA 98195

Dr. Brian Leaderer-John B. Pierce Laboratory,  290 Congress Avenue, New Haven,
CT 06519

Dr. Helen  Sub-Harvard University, School of Public Health, 665 Huntington Avenue,
Boston, MA  02115


April 1995                            I-xxxvi      DRAFT-DO NOT QUOTE OR CITE

-------
              AUTHORS, CONTRIBUTORS, AND REVIEWERS (cont'd)
                            CHAPTER 7.  EXPOSURES
Principal Authors

Dr. Dave Mage-Atmospheric Research and Exposure Assessment Laboratory (MD-56),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Lance Wallace-Office of Research and Development, U.S. Environmental Protection
Agency, Wallington, VA 22186

Dr. Robert Ziegenfus-Kutztown University,  Department of Geography, Kutztown, PA 19550
Contributors and Reviewers

Dr. Gerald Akland-Atmospheric Research and Exposure Assessment Laboratory (MD-56),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Timothy Buckley-Atmospheric Research and Exposure Assessment Laboratory (MD-56),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Robert Burton-Atmospheric Research and Exposure Assessment Laboratory (MD-56),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Mr. William Ewald-Environmental Criteria and Assessment Office (MD-52),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711

Dr. Paul Lioy-Environmental Occupational Health Science Institute, Exposure Measurement
and Assessment Division, 681 Frelinghuysen Road, Piscataway, NJ  08855-1179

Dr. Peter McMurry-University of Minnesota, Department of Mechanical Engineering,
111 Church Street, SW, Minneapolis, MN 55455

Dr. William Wilson-Atmospheric Research and Exposure Assessment Laboratory (MD-56),
U.S. Environmental Protection Agency, Research Triangle Park, NC  27711
April 1995                            I-xxxvii     DRAFT-DO NOT QUOTE OR CITE

-------
                                     FORWARD

     As explained in Chapter 2 (Introduction), the present draft document, Air Quality
Criteria for Particulate Matter, is being prepared to meet Clean Air Act provisions which
require that the Administrator of the United States Environmental Protection Agency (U.S.
EPA) periodically review and revise, as appropriate, the criteria and National Ambient Air
Quality Standards (NAAQS)  for paniculate matter and other listed "criteria air pollutants".
This criteria assessment, therefore, contains evaluations of new scientific information that has
become available since the last prior criteria reviews for paniculate matter were carried out
by U.S. EPA in the 1980's, as also described in Chapter 2.
     Responsibility for preparation of this External Review Draft of the subject Particulate
Matter Air Quality Criteria Document (PM AQCD) falls within the mission of U.S. EPA's
Environmental Criteria and Assessment Office in Research Triangle Park, NC.  That office
(ECAO/RTP) is a component of the Office of Health and  Environmental Assessment
(OHEA) within the Office of Research and Development (ORD),  the scientific arm of
U.S. EPA. Members of the  U.S. EPA Project Development Team for Development of the
present draft document are listed in ensuing pages of the front matter for this volume (I of
III) of the document, and include both ECAO/RTP staff and a few other scientists on
temporary assignment to ECAO/RTP.
     The U.S. EPA Project Development team has carried out preparation of an overall
Project Development Plan for preparation of the  subject PM AQCD including identification
of key issues to be addressed, planned content of the document, description of the process
and schedule for preparation  and review of draft materials, and identification of U.S. EPA
staff scientists and non-EPA consultants expected to serve as authors of sections of the
document.  That Project Development Plan was reviewed  by members  and consultants of the
Clean Air Act Scientific Advisory Committee (CASAC) listed on  Pages xlv-xlvii, and it was
appropriately modified in response to their comments and recommendations.  The EPA
Project Team has  also coordinated and implemented the planning and execution of specific
logistical arrangements, by which the writing, word processing, editing, and assembly  of
draft chapter materials and their preliminary peer review and subsequent revision to the
present External Review Draft version have been accomplished.

April 1995                             I-xxxviii     DRAFT-DO NOT QUOTE OR CITE

-------
                                  FORWARD (cont'd)

      Initial drafts of various chapters (other than Chapters 1 and 13) were prepared mainly
during summer/fall of 1994 and subsequently underwent preliminary peer review in early
1995, either (a) as part of public expert workshops held in January 1995 where most key
chapters (4, 6, 7, 10, 11, 12) were discussed in open forum, or (b) via receipt of oral or
written comments on  the other chapters from internal EPA scientific staff and/or selected
non-EPA experts.  Revisions made in response to such preliminary reviews were then
incorporated into revised chapters  to produce this External Review Draft of the document.
      The principal authors of materials contained in each chapter of the document are
identified in the front matter for each of the three volumes of this  External Review Draft,  as
shown on pages preceding this forward for Volume I chapters.  Other contributors and
reviewers,  who provided comments on initial drafts of particular chapters or other
information importantly considered as inputs to revisions incorporated into the chapters, are
also identified under "Contributors and Reviewers" in the front matter materials  for each
volume. However, the evaluations and conclusions contained in this External  Review Draft
do not necessarily reflect the individual views of all identified authors, contributors, and
reviewers.
      The present External Review Draft of the subject  PM AQCD has been prepared for
release  for public comment and review by CAS AC, as  mandated by the Clean Air Act.  The
public comment period, extending  from May 1, 1995 to August 1, 1995, will  be followed
soon after by a public meeting of CASAC in early August, 1995 (specific site and dates to be
announced  in the  Federal Register).  Further revisions will then be incorporated  into this
draft document in response to public comments and CASAC peer review to produce a final
version of this document by no  later than January, 1996.
April 1995                              I-xxxix     DRAFT-DO NOT QUOTE OR CITE

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

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

Dr. Dennis Kotchmar—Project Manager, Medical Officer, Environmental Criteria and
     Assessment Office (MD-52), U.S. Environmental Protection Agency, Research Triangle
     Park,  NC  27711

Ms. Beverly Comfort—Deputy Project Manager/Technical Project Officer, Health Scientist,
     Environmental Criteria and Assessment Office (MD-52), U.S. Environmental Protection
     Agency, Research Triangle Park, NC 27711

Mr. Norman Childs—Branch Chief, Environmental Media Assessment  Branch,
     Environmental Criteria and Assessment Office (MD-52), U.S. Environmental Protection
     Agency, Research Triangle Park,  NC  27711

Dr. A. Paul Altshuller—Technical Consultant, Environmental Criteria and Assessment Office
     (MD-52), U.S. Environmental Protection Agency, Research Triangle Park,  NC  27711

Mr. William Ewald—Technical Project Officer, Health Scientist, Environmental Criteria and
     Assessment Office (MD-52), U.S. Environmental Protection Agency, Research Triangle
     Park,  NC  27711

Dr. Jasper Garner—Technical Project Officer, Ecologist,  Environmental Criteria and
     Assessment Office (MD-52), U.S. Environmental Protection Agency, Research Triangle
     Park,  NC  27711

Dr. Judith A. Graham—Associate Director, Environmental Criteria and Assessment Office
     (MD-52), U.S.  Environmental Protection Agency, Research Triangle Park,  NC  27711

Dr. Allan Marcus—Technical Project Officer, Statistician, Environmental Criteria and
     Assessment Office (MD-52), U.S. Environmental Protection Agency, Research Triangle
     Park,  NC  27711

Ms. Beverly Tilton—Technical Project Officer, Physical Scientist, Environmental Criteria
     and Assessment Office (MD-52), U.S. Environmental Protection Agency, Research
     Triangle Park, NC 27711
April 1995                               I_xl       DRAFT-DO NOT QUOTE OR CITE

-------
                  U.S. ENVIRONMENTAL PROTECTION AGENCY
        PROJECT TEAM FOR DEVELOPMENT OF AIR QUALITY CRITERIA
                           FOR PARTICULATE MATTER
                                     (cont'd)
Ms. Annie M. Jarabek—Technical Project Officer, Toxicologist, Environmental Criteria and
     Assessment Office (MD-52), U.S. Environmental Protection Agency, Research Triangle
     Park, NC  27711

Dr. James McGrath—Technical Project Officer, Visiting Senior Health Scientist,
     Environmental Criteria and Assessment Office (MD-52), U.S. Environmental Protection
     Agency, Research Triangle Park, NC  27711

Dr. William Wilson—Technical Consultant, Physical Scientist, Atmospheric Research and
     Exposure Assessment Laboratory (MD-75), U.S. Environmental Protection Agency,
     Research Triangle Park,  NC 27711
Technical Support Staff

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

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

Ms. Diane H. Ray—Program Analyst, Environmental Criteria and Assessment Office
     (MD-52), U.S. Environmental Protection Agency, Research Triangle Park, NC 27711

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

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

Mr. Richard Wilson—Clerk, Environmental Criteria and Assessment Office (MD-52),
     U.S.  Environmental Protection Agency, Research Triangle Park, NC 27711
April 1995                              I_xli       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

Ms. Marianne Barrier—Graphic Artist, ManTech Environmental Technology, Inc.,
P.O. Box 12313, Research Triangle Park, NC  27709

Mr. John R. Barton—Document Production Coordinator, ManTech Environmental
Technology, Inc., P.O. Box 12313, Research Triangle Park, NC  27709

Ms. Suzanne Borneman—Word Processor, ManTech Environmental Technology, Inc.,
P.O. Box 12313, Research Triangle Park, NC  27709

Ms. Lynette D. Cradle—Word Processor, ManTech Environmental Technology, Inc.,
P.O. Box 12313, Research Triangle Park, NC  27709

Ms. Jorja R. Followill—Word Processor, ManTech Environmental Technology, Inc.,
P.O. Box 12313, Research Triangle Park, NC  27709

Ms. Sheila R. Lassiter—Word Processor, ManTech Environmental Technology, Inc.,
P.O. Box 12313, Research Triangle Park, NC  27709

Ms. Wendy B. Lloyd—Lead Word Processor, ManTech Environmental Technology, Inc.,
P.O. Box 12313, Research Triangle Park, NC  27709

Ms. Cheryl B. Thomas—Word Processor, ManTech Environmental Technology, Inc.,
P.O. Box 12313, Research Triangle Park, NC  27709

Mr. Peter J. Winz—Technical Editor, Mantech Environmental Technology, Inc.,
P.O. Box 12313, Research Triangle Park, NC  27709
Technical Reference Staff

Mr. John A. Bennett—Bibliographic Editor, ManTech Environmental Technology, Inc.,
P.O. Box 12313, Research Triangle Park, NC 27709

Ms. S. Blythe Hatcher—Bibliographic Editor, Information Organizers, Inc.,
P.O. Box 14391, Research Triangle Park, NC 27709

Ms. Susan L. McDonald—Bibliographic Editor, Information Organizers, Inc.,
P.O. Box 14391, Research Triangle Park, NC 27709

April 1995                             I-xlii      DRAFT-DO NOT QUOTE OR CITE

-------
                 U.S. ENVIRONMENTAL PROTECTION AGENCY
        PROJECT TEAM FOR DEVELOPMENT OF AIR QUALITY CRITERIA
                         FOR PARTICULATE MATTER
                                    (cont'd)
Ms. Carol J. Rankin—Bibliographic Editor, Information Organizers, Inc., P.O. Box 14391,
Research Triangle Park, NC  27709

Ms. Deborah L. Staves—Bibliographic Editor, Information Organizers, Inc.,
P.O. Box 14391, Research Triangle Park, NC 27709

Ms. Patricia R. Tierney—Bibliographic Editor, ManTech Environmental Technology, Inc.,
P.O. Box 12313, Research Triangle Park, NC 27709
April 1995                           I.xiiii      DRAFT-DO NOT QUOTE OR CITE

-------
                 U.S. ENVIRONMENTAL PROTECTION AGENCY
                          SCIENCE ADVISORY BOARD
                CLEAN AIR SCIENTIFIC ADVISORY COMMITTEE

                PARTICULATE MATTER CRITERIA DOCUMENT
                    PROJECT DEVELOPMENT PLAN REVIEW
Chairman
Dr. George T. Wolff—General Motors Corporation, Environmental and Energy Staff,
General Motors Bldg., 12th Floor, 3044 West Grand Blvd., Detroit, MI  48202
Members

Dr. Stephen Ay res—Office of International Health Programs, Virginia Commonwealth
University, Medical College of Virginia, Box 980565, Richmond, VA  23298

Dr. Jay Jacobson-Boyce Thompson Institute, Tower Road, Cornell University, Ithaca,
NY 14853

Dr. Benjamin Liu-Department of Mechanical Engineering, Institute of Technology,
111 Church Street, S.E., Minneapolis, MN 55455

Dr. Joseph Mauderly—Inhalation Toxicology Research Institute, Lovelace Biomedical and
Environmental Research Institute, P.O. Box 5890, Albuquerque,  NM 87185

Dr. Paulette Middleton—University Cooperation for Atmospheric Research, P.O. Box 3000,
Boulder, CO 80307

Dr. James H. Price, Jr.—Research and Technology Section, Texas Natural Resources
Conservation Commission, P.O. Box 13087, Austin, TX 78711
Invited Scientific Advisory Board Members

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

Dr. Roger O. McClellan—Chemical Industry Institute of Toxicology, P.O. Box 12137,
Research Triangle Park, NC 27711
Consultants

Dr. Stanley Auerbach—Environmental Sciences Division, Oak Ridge National Laboratory,
Bldg. 1505, MS 6036, Oak Ridge, TN 37831

April 1995                             I-xlv     DRAFT-DO NOT QUOTE OR CITE

-------
                 U.S. ENVIRONMENTAL PROTECTION AGENCY
                          SCIENCE ADVISORY BOARD
                CLEAN AIR SCIENTIFIC ADVISORY COMMITTEE
                                     (cont'd)
Consultants (cont'd)

Dr. Petros Koutrakis-Harvard School of Public Health, 665 Huntington Avenue, Boston,
MA 02115

Dr. Kinley Larntz-Department of Applied Statistics, University of Minnesota, 352 COB,
1994 Beuford Avenue, St. Paul, MN 55108-6042

Dr. Allan Legge-Biosphere Solutions, 1601 llth Avenue, N.W., Calgary, Alberta  T2N
1H1, Canada

Dr. Daniel Menzel-Department of Community and Environmental Medicine, University of
California, 19172 Jamboree Boulevard, Irvine,  CA 92717-1825

Dr. Frederick J. Miller-Chemical Industry Institute of Toxicology,  P.O. Box 12137,
Research Triangle Park, NC 27709

Dr. William R. Pierson-Energy and Environmental Engineering Center, Desert Research
Institute, P.O. Box 60220,  Reno, NV 89506-0220

Dr. Christian Seigneur-ENSR Consulting and Engineering, 1320 Harbor Bay Parkway,
Suite 210,  Alameda, CA 94502

Dr. Carl M. Shy-Department of Epidemiology, School of Public Health, University of North
Carolina, CB #7400 McGravran-Greenberg Hall, Chapel Hill, NC  27599-7400

Dr. Jan Stolwijk-Epidemiology and Public Health, Yale University, 60 College Street,
New Haven, CT  06510

Dr. Warren White-Washington University, Campus Box 1134, 684 Waterman Avenue,
St. Louis, MO  63130-4899

Dr. Ron Wyzga-Electric Power Research Institute, 3412 Hill view Avenue, P.O.  Box  10412,
Palo Alto,  CA  94303

Dr. Mark J. Utell—Pulmonary Disease Unit, Box 692, University of Rochester Medical
Center, 601 Elmwood Avenue, Rochester, NY  14642
April 1995                             I-xlvi     DRAFT-DO NOT QUOTE OR CITE

-------
                 U.S. ENVIRONMENTAL PROTECTION AGENCY
                         SCIENCE ADVISORY BOARD
                CLEAN AIR SCIENTIFIC ADVISORY COMMITTEE
                                    (cont'd)
Designated Federal Official

Mr. Randall C. Bond—Science Advisory Board (1400), U.S. Environmental Protection
Agency, 401 M Street, SW, Washington, DC  20460
Staff Assistant

Ms. Janice M. Cuevas—Science Advisory Board (1400), U.S. Environmental Protection
Agency, 401 M Street, SW, Washington, DC  20460
Secretary

Ms. Lori Anne Gross—Science Advisory Board (1400), U.S. Environmental Protection
Agency, 401 M Street, SW, Washington, DC  20460
April 1995                           I-xlvii     DRAFT-DO NOT QUOTE OR CITE

-------
 i                          1.   EXECUTIVE SUMMARY
 2
 3
 4      1.1   PURPOSE OF THE DOCUMENT
 5          The purpose of this document is to present air quality criteria for paniculate matter
 6      (PM) in accordance with the Clean Air Act (CAA). Two sections of the CAA (Sections 108
 7      and 109) govern the establishment, review, and revision of National Ambient Air Quality
 8      Standards (NAAQS). Section 108 directs the Administrator of the U.S.  Environmental
 9      Protection Agency (EPA) to list pollutants that may reasonably be anticipated to endanger
10      public health or welfare and to issue air quality criteria for them. The air quality criteria are
11      to reflect the latest scientific information useful in indicating the kind and extent of all
12      exposure-related effects on public health and welfare that may be expected from the presence
13      of the pollutant in ambient  air.
14
15
16      1.2   INTRODUCTION
17          Air Quality Criteria for Particulate Matter evaluates the latest scientific information
18      useful in deriving  criteria that form the scientific basis for U.S. Environmental Protection
19      Agency (EPA) decisions regarding the National Ambient Air Quality Standards (NAAQS) for
20      paniculate matter  (PM).  This Executive Summary concisely  summarizes key conclusions
21      from the document which comprises thirteen chapters.  The Executive Summary is followed
22      by a general introduction in Chapter 2.  Chapters 3 through 7 provide background
23      information on physical and chemical properties of PM and related compounds; sources  and
24      emissions; atmospheric transport,  transformation,  and fate of PM; methods for the collection
25      and measurement of PM; and ambient air concentrations and factors affecting exposure of the
26      general population.  Chapter 8 describes effects on visibility, and Chapter 9 describes
27      damage to materials attributable to PM.  Chapters 10 through 13  evaluate information
28      concerning the health effects of PM.  More specifically, Chapter  10 discusses dosimetry of
29      inhaled particles in the respiratory tract; Chapter 11 summarizes information on the
30      toxicology of specific types of PM constituents, and includes  experimental toxicological
31      studies of animals  and human clinical studies.  Chapter 12 discusses epidemiological studies.
        April 1995                               1-1       DRAFT-DO NOT QUOTE OR  CITE

-------
 1     Chapter 13 integrates information on critical health issues derived from studies reviewed in
 2     the prior chapters.  The following sections conform to the chapter organization of the criteria
 3     document.
 4
 5
 6     1.3  PHYSICS AND CHEMISTRY OF PARTICULATE MATTER
 7          Chapter three first describes the physical properties, chemistry, and chemical
 8     composition of PM. Next, the transport and transformation to secondary paniculate matter
 9     are presented. Then, both dry and wet deposition are discussed. Finally, the physical and
10     chemical considerations in paniculate matter sampling and analysis are examined. The
11     following brief comments present some basic characteristics of PM.
12          Atmospheric particles originate from a variety of sources and possess a range of
13     morphological, chemical, physical, and thermodynamic properties.  Examples include
14     combustion-generated particles such as diesel soot or fly ash, photochemically produced
15     particles  such as those found in urban haze, salt particles formed from sea spray, and soil-
16     like particles from resuspended dust. Some particles are liquid, some are solid; others
17     contain a solid core surrounded by liquid.  Atmospheric particles contain inorganic ions and
18     elements, elemental carbon,  organic compounds, and crustal compounds.  Some atmospheric
19     particles  are hygroscopic and contain particle-bound water. The organic fraction is especially
20     complex.  Hundreds of organic compounds have been identified in atmospheric aerosols,
21     including alkanes, alkanoic and carboxylic acids, polycyclic aromatic hydrocarbons,  and
22     nitrated organic compounds.
23          Particle diameters span more than four orders of magnitude, from a few nanometers to
24     one hundred micrometers. Combustion-generated particles, such as those from power
25     generation, from automobiles, and in tobacco smoke, can be as small as 0.01 /mi and as
26     large as 1 /urn. Particles produced in the atmosphere by photochemical processes range in
27     diameter  from 0.05 to 2 ^m.  Fly ash produced by coal combustion ranges from 0.1 to
28     50 jum or more.  Wind-blown dust,  pollens, plant fragments, and cement dusts are generally
29     above 2 /xm in diameter. Particles as small as a few nanometers and as large as 100
30     have been measured in the atmosphere.
       April 1995                               1-2      DRAFT-DO NOT QUOTE OR CITE

-------
  1           The composition and behavior of airborne particles are fundamentally linked with those
  2      of the surrounding gas.  Although the term aerosol is often used to refer to suspended
  3      particles, aerosol is defined as a dilute suspension of solid or liquid particles in gas.
  4      Paniculate material can be primary or secondary.
  5           Primary particles are those emitted in particulate form and include wind-blown dust, sea
  6      salt, road dust, mechanically generated particles and combustion-generated particles such as
  7      fly ash and soot.  The concentration of primary particles depends on their emission rate,
  8      transport and dispersion,  and removal rate from the atmosphere.
  9           Secondary particulate material may form from condensation of high temperature  vapor
10      or from vapors generated at as a result of chemical reactions  involving gas-phase precursors.
11      Secondary formation processes can result in either the formation of new particles or the
12      addition of particulate material to preexisting particles.  Most atmospheric sulfate is formed
13      from atmospheric oxidation of sulfur dioxide.  Atmospheric nitrate is also essentially
14      secondary, formed from reactions involving oxides of nitrogen to form nitric acid.  A portion
15      of the organic aerosol is also attributed to secondary processes.  Secondary aerosol formation
16      can depend on concentrations of other gaseous reactive species such as ozone or hydrogen
17      peroxide, atmospheric conditions including  solar radiation  and relative humidity, and the
18      interactions of precursors and preexisting particles with cloud or fog droplets.  As a result, it
19      is considerably more difficult to relate ambient concentrations of secondary species to  sources
20      of precursor emissions than it is to identify the sources of  primary  particles.
21           Airborne particulate matter can be anthropogenic or biogenic  in origin.  Both
22      anthropogenic and  biogenic particulate material can occur from  either primary or secondary
23      processes.  Anthropogenic refers to particulate matter which is directly emitted or formed
24      from precursors which are emitted as a result of human activity. Primary anthropogenic
25      sources include fossil fuel combustion, fireplace emissions, and  road dust.   Secondary
26      anthropogenic particulate material can be generated photochemically from anthropogenic
27      SO2, NOX, or organic gases.  Primary biogenic sources include  leaf waxes and other plant
28      fragments from plants. In addition, plants emit gaseous species such as terpenes. Terpenes
29      are photochemically reactive, and in the presence of nitrogen oxides can form secondary
30      organic particles.  Other types  of primary particulate material such as sea salt and wind-
31      generated dust from soil undisturbed by man also are of non-anthropogenic  origin.

        April 1995                                 1-3       DRAFT-DO NOT QUOTE OR CITE

-------
  1      1.4   SAMPLING AND ANALYSIS OF PARTICULATE MATTER AND
  2            ACID DEPOSITION
  3           This chapter is intended to provide supplemental information to Chapter 3 and other
  4      discussions of aerosol measurement methodologies used in support of the existing PM10
  5      standards and/or potentially useful in considerations  related to the possible need for
  6      development of  a "fine particle" standard with an upper cut-point in the 1 to 3 /^m range.
  7      The discussion of ambient PM monitoring methods is also included to enhance understanding
  8      of exposure data (and their interpretation) used in epidemiology analyses assessed in Chapter
  9      12 of this document.  An important contribution of the sampling  and analytical sections is the
10      extensive compilation of salient peer-reviewed technical references that can be consulted by
11      the reader for more detailed information.
12           Chapter four briefly describes the technical capabilities  and  limitations of aerosol
13      sampling and analytical procedures focusing on  those that were used: (1) to collect data
14      supporting other sections in this  document;  (2) to support the existing PM10, TSP, and Pb
15      regulations; (3)  to support health and welfare  effects studies;  and/or (4) have application to
16      development of  a possible fine particle standard; or (5) illustrate the attributes of several new
17      technologies.  The discussion of aerosol separation technologies is divided between (1)
18      devices used to mimic the larger particle (> 10 /im)  penetration rationales for the upper
19      respiratory tract airways, and (2) those devices generally used to  mimic smaller particle
20      penetration (< 10 pm) to the thoracic regions.  These device descriptions are followed by
21      sampling considerations for their applications.
22           The applications of performance specifications  to define these measurement systems for
23      regulatory purposes are discussed,  along with a  number of critical observations suggesting
24      that  the current  specification process does not always ensure the accuracy or
25      representativeness necessary in the field.
26           The EPA program designating PM10 reference  and equivalent sampling systems is
27      briefly described, along with a current list of designated devices.  Selected measurement
28      systems used to  provide more detailed characterization of aerosol  properties  for research
29      studies are discussed, with  a focus on the determination of particle size distributions.
30      Aerosol sampling systems for specialty applications,  including automated samplers, personal
       April 1995                                1-4       DRAFT-DO NOT QUOTE OR CITE

-------
  1      exposure samplers, and the sampling systems used in aerosol apportionment studies are
  2      briefly described.
  3
  4
  5      1.5   SOURCES AND EMISSIONS OF SUSPENDED PARTICLES
  6           Chapter five is organized to present first a concise summary of key information on PM
  7      emissions derived from the previous criteria review in the 1980's and then to provide a more
  8      extensive discussion of newer information appearing in recent years.
  9           The main objectives of  Chapter 5 discussions are:
 10           •     To identify the sources  that are major contributors to suspended particle
 11                 concentrations in the United States.
 12
 13           •     To describe the particle sizes and chemical properties of source emissions.
 14
 15           •     To evaluate the limitations and uncertainties of emissions rate estimates and
 16                 source contributions for suspended particles and their gaseous precursors.
 17           The ambient atmosphere contains both primary and secondary particles; the former are
 18      emitted directly by sources, and the latter are formed from gases (SO2, NOX, NH3, VOCs).
 19      Fugitive dust is a primary pollutant.  Major sources of particle emissions are classified as
 20      major point sources, mobile sources, and area sources; these are anthropogenic.  Natural
 21      sources also contribute to ambient concentrations.
 22           The 1982 Criteria Document emphasized emissions from industrial sources, especially
 23      primary particles.  SO2 was the only precursor of secondary particles considered.  Since
 24      1982, many of these sources have been controlled, yet particle standards are exceeded in
 25      many areas.
 26           Source and receptor models are used to quantify major contributions to excess PM10
 27      concentrations. Source models use emissions inventories and meteorological data to predict
 28      particle formation  dispersion and particle concentrations measured at receptors.  Receptor
29      models use the chemical composition of emissions (finger points) and concentrations at
30      receptor sites to estimate the contribution of sources.  The latter  are used to identify sources
31      in non-attainment areas.
       April 1995                                !_5       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Fugitive dust is a major contribution to PM10 at nearly all sampling sites, although the
 2      average fugitive dust source contribution is highly variable among sampling sites within the
 3      same areas, and is highly variable between seasons.
 4           Primary motor vehicle exhaust makes up as much as 40% of average PM10 at many
 5      sampling  sites.  Vegetative burning outdoor and residential wood burning are significant
 6      sources in residential areas.  Fugitive dust from paved and unpaved roads,  agricultural
 7      operations, construction, and soil erosion constitute —90% of nationwide primary  emissions.
 8      Fugitive dust consists of geological material that is suspended  into the atmosphere  by natural
 9      wind and by anthropogenic activities from sources such as  paved and unpaved roads,
10      construction and demolition of buildings and roads, storage piles, wind erosion, and
11      agricultural tilling. There are obvious discrepancies between the proportion of fugitive dust
12      in primary emissions and geological contributions  to PM10 calculated by receptor models,
13      due to contributions from secondary aerosols, which are not included in the primary PM10
14      emission estimates.  Even when secondary  aerosol is subtracted, however,  other sources such
15      as vegetative burning and wood combustion make larger relative contributions to ambient
16      concentrations than is indicated by the emissions inventories.  Fugitive dust estimates are
17      especially affected by the general limitations of emissions inventories. All of the emissions
18      have remained relatively constant over the 8-year period except for those from soil erosion.
19           The major non-fugitive dust emitters are other industrial  processes and exhaust from
20      highway vehicles.  Fuel combustion from utilities, industrial, and other sources together
21      contribute between 1  to 2% to  total primary particle emissions.  Industrial  fuel combustion
22      emissions were reduced by one-third and other fuel combustion emissions were reduced by
23      one half between  1983 and 1992.  On-highway vehicle emissions increased by 50%,
24      primarily due to large increases in the number of vehicle miles traveled.  Electric  utilities
25      account for the largest fraction of sulfur dioxide, nearly 70% of total emissions. These
26      emissions have not changed substantially over the  10 years reported. Annual averages do not
27      reflect the seasonality of certain emissions, residential wood burning in fireplaces and stoves,
28      for example.  Cold weather also affects motor vehicle exhaust emissions, both in terms of
29      chemical composition and emission rates. Planting, fertilizing, and harvesting  are also
30      seasonal.
        April 1995                                1-6       DRAFT-DO NOT QUOTE OR CITE

-------
  1           Mobile sources are major emitters of primary particles, oxides of nitrogen, and volatile
  2      organic compounds.  They are also minor emitters of sulfur dioxide and ammonia.  On-road
  3      motor vehicles using gasoline-and diesel-fueled engines are by far the largest component of
  4      mobile source emissions,  and the emissions estimation methods are most highly developed
  5      for these vehicles.  Studies show that while vehicle emissions models may function well
  6      under idealized conditions, they underestimate the effects of high emitting vehicles that may
  7      be major sources of VOCs.  Motor vehicle exhaust contains high concentrations of organic
  8      and elemental carbon, but their ratios are much different from those found in wood
  9      combustion with the abundance of elemental carbon being nearly equal to  the organic carbon
10      abundance.
11           There are major discrepancies between the relative amounts of emissions and
12      contributions to suspended particles found in many  areas.  Some major re-design is needed to
13      create more accurate emissions models that can improve the quantification of source-receptor
14      relationships.  Emissions models are intended to estimate the emissions rates as a function of
15      space and time of selected pollutants from point,  area, and mobile sources. In contrast to an
16      emissions inventory, which is a static catalogue of emissions estimates for a given
17      geographical area and averaging time, an emissions model is capable of accessing activity
18      data bases from a multitude of information-gathering agencies and determining actual
19      emissions for relatively small regions and averaging times.
20
21
22      1.6   ENVIRONMENTAL CONCENTRATIONS
23           Chapter six summarizes PM concentrations  over the United States, including the
24      spatial, temporal,  size, and chemical aspects.   This chapter mainly aims to provide
25      background information on U.S. PM concentrations to help set a context for discussions in
26      later  chapters on the characterization and quantification of PM health effects.  The general
27      approach in the chapter is to organize, evaluate, and summarize  the existing large scale
28      aerosol data sets over the  United States.  Emphasis  is placed on complete national coverage
29      as well as the fusion and reconciliation of multiple data sets.
30           The main organizing dimension used to structure Chapter 6 is space. Accordingly, PM
31      concentrations are presented on global, continental, national, regional, and sub-regional or

        April 1995                               1-7       DRAFT-DO NOT QUOTE OR CITE

-------
 1     urban scales.  Within each spatial domain, the spatial-temporal structure, aerosol size and
 2     chemical composition are also presented. The presentation of aerosol pattern begins with a
 3     global and continental perspective.  Next, national U.S. aerosol patterns are examined, as
 4     derived from non-urban and urban PM10 and PM2 5 monitoring networks, and the aerosol
 5     characteristics over seven subregions of the contiguous United States are examined in more
 6     detail. Ten year trends, seasonal patterns, as  well as the PM2 5/PM10 relationship and fine
 7     particle chemical composition are examined for each region.  An ensuing section then
 8     focuses further on the sub-regional and urban-scale aerosol patterns over representative areas
 9     of the United States.
10          The aerosol concentration pattern over the United States has been reported by many
11     aerosol researchers over the past decade. In particular the research groups associated with
12     the IMPROVE aerosol monitoring networks have been prolific producers of high quality
13     data, reports, and analysis of non-urban data. This section draws heavily on their
14     contribution but the maps, charts, and computations have been re-done for sake of
15     consistency with other (urban) data from the  AIRS network. Each of the sections are
16     augmented by suitable but not exhaustive references to the pertinent literature.
17          Table 1-1 summarizes,  for illustrative purposes, annual  average urban PM10 data  from
18     the AIRS Network by U.S. region. The annual means for all regions show  declines from
19     1985 to 1993.  Tables 1-2 and 1-3 provide more specific illustrative information on 24-h
20     mean, second highest maximum, and annual  average values for PM10 concentrations in 1993
21     for selected U.S. cities from various U.S. regions.
22           Next, a section on chemical composition of PM aerosols at urban and non-urban sites
23     summarizes available data on the composition of atmospheric particles.  Emphasis has been
24     placed on the Harvard six-city study and the  inhalable paniculate network (1980-1981).
25     However, data for fine particle mass and elemental composition only were available from
26     these studies.  Data for sulfate, nitrate, and elemental and organic carbon content are
27     included from other studies to provide an overview of the chemical composition of the
28     atmospheric aerosol in the United States.  Extensive tables in this section provide detailed
29     representation of atmospheric properties of aerosols to which U.S. populations are exposed.
30     Unfortunately, data this complete are generally collected over limited time periods and are
31     not of sufficient duration to be useful for most epidemiological

       April  1995                                1-8       DRAFT-DO NOT QUOTE OR CITE

-------
        TABLE 1-1.  CHARACTERIZATION OF URBAN PM10 DATA FROM AIRS NETWORK BY REGION FOR THE

                                                   UNITED STATES
Region
Northeast


Southeast

Industrial Midwest


Upper Midwest

Southwest


Northwest

Southern California


PM10
1993
22


24

25


25

26


25

32


(/*g/m3)
1985
36


32

38


31

52


50

45


SDa
30%


17%

28%


19%

45%


45%

40%


Seasonality
Summer Peak July


Summer High
July, August
High June, August
Low November, February

Slightly Lower Levels
December/January
April-June Peak October-
November Peak
August-September Dip
Peak December
Low March-May
Peak November
Low March

Seasonal
Variation %
20 %b


37%

37%


19%




36%

27%


PM2.5/PM10
62%


58%

59%


38%

37%


59%

50%


Influences
Canadian and Gulf
airmasses, local sources,
long range transport
Flat, poor regional
ventilation
Winter cold Canadian
airmasses. Summer moist
Gulf Coast masses
Agricultural Heartland
windblown dust influence
low precipitation, coarse
particle dominant, dust
contribution to PM10
Meteorology highly
variable
Air flow from Pacific, dry
summer, low in remote
Basin wide elevation
O
O
aStandard deviation among monitoring stations within regions.

bSeasonal range expressed as percent.
O
H

O
c;
o
H
w

o
n
HH
H
W

-------
    TABLE 1-2. PM10 LEVELS BY ANNUAL
           AVERAGE FOR SELECTED U.S.
SMSAs FOR 1993
Metropolitan Statistical Area
Santa Fe, NM
Amarillo, TX
Santa Rosa, CA
Springfield, MD
Casper, WY
Danbury, CT
Glens Falls, NY
Titusville Area, FL
New London Area, CT/RI
Bridgeport, CT
Fort Lauderdale, FL
Asheville, NC
Montgomery, AL
Honolulu, HI
Oakland, CA
Charleston, SC
San Francisco, CA
Dallas, TX
Louisville, KY
Baltimore, MD
Birmingham, AL
Mobile, AL
Orange County, CA
Phoenix, AZ
New York, NY
1990
Population
117,043
187,547
388,222
239,971
61,226
187,867
118,539
398,978
266,819
443,722
1,255,480
174,821
292,517
836,231
2,082,914
506,875
1,603,678
2,553,362
952,662
2,382,172
907,810
476,923
2,410,556
2,122,101
8,546,846
PM10
WTD AM1
0*g/m3)
15
16
18
18
18
19
19
19
19
21
21
22
23
24
26
26
29
30
33
35
36
38
38
44
47
PM10
2nd Max2
Og/m3)
35
29
52
39
41
46
44
57
41
50
71
58
48
58
71
58
72
74
73
70
85
71
80
92
86
O3 (ppm)3
—
-
—
—
—
0.14
—
—
0.13
0.17
—
—
—
—
0.13
—
—
0.14
0.14
0.15
0.13
—
0.17
0.13
—
'Weighted Annual Mean
2Highest Second Maximum 24-hour Concentration
3Highest O3 Second Daily Maximum 1-hour Concentration
April 1995
      1-10
DRAFT-DO NOT QUOTE OR CITE

-------
           TABLE 1-3.  SELECTED U.S. PM10 LEVELS BY 2ND MAX PM10 FOR 1993
Metropolitan Statistical Area
St. Louis, MO
Los Angeles, CA
San Diego, CA
El Paso, TX
Medford, OR
Seattle, WA
Gary, IN
Flint, MI
Bakerville, CA
Fresno, CA
Denver, CO
Chicago, IL
Eugene, OR
Salt Lake City, UT
Spokane, WA
Pittsburgh, PA
Riverside, CA
Steubenville, OH
New Haven, CT
Provo, UT
Philadelphia, PA
1990
Population
2,444,099
8,863,164
2,498,016
591,610
146,389
1,972,961
604,526
430,459
543,477
667,490
1,622,980
6,069,974
282,912
1,072,227
361,364
2,056,705
2,588,793
142,523
638,220
263,590
4,856,881
PM10
2nd Max2
(Mg/m3)
101
102
105
106
106
119
122
127
128
131
142
147
151
156
166
167
172
177
178
209
531
PM10
WTD AM1
(Mg/m3)
44
47
34
37
41
35
34
24
54
53
41
47
28
42
46
38
73
40
52
40
34
Note
O33 0.13
O3 - 0.25, CO4-
14
03 0.16
03 0.14, CO-11
—
—
—
—
O3 0.16
O3 0.14
—
—
—
—
CO 12
SO25 0.155
O3 0.23
S02 0.244
O3 0.15
CO 10
O3 0.14
       1 Weighted Annual Mean
       2Highest Second Maximum 24-hour Concentration
       3Highest O3 Second Daily Maximum 1-hour Concentration
       4Highest CO Second Maximum Non-overcapping 8-hour Concentration
       5Highest SO2 Second Maximum 24-hour Concentration
1     investigations.  The tables do, however, provide insights as to the types of information that
2     could be collected as part of future monitoring efforts in support of human exposure
3     investigations.  Table 1-4 provides illustrative data from an earlier 1987 study with regard to
      April 1995
1-11
DRAFT-DO NOT QUOTE OR CITE

-------
              TABLE 1-4.  CONCENTRATION RANGES OF VARIOUS ELEMENTS
                     ASSOCIATED WITH PARTICULATE MATTER IN THE
                            UNITED STATES ATMOSPHERE (ng/m3)
Elements
As
Cd
Ni
Pb
V
Zn
Co
Cr
Cu
Fe
Hg
Mn
Se
Sb
Remote
0.007 to 1.9
0.003 to 1.1
0.01 to 60
0.007 to 64
0.001 to 14
0.03 to 460
0.001 to 0.9
0.005 to 11.2
0.029 to 12
0.62 to 4160
0.005 to 1.3
0.01 to 16.7
0.0056 to 0.19
0.0008 to 1.19
Rural
1.0 to 28
0.4 to 1000
0.6 to 78
2 to 1700
2.7 to 97
11 to 403
0.08 to 10.1
1.1 to 44
3 to 280
55 to 14530
0.05 to 160
3.7 to 99
0.01 to 3.0
0.6 to 7
Urban (USA)
2 to 2320
0.2 to 7000
1 to 328
30 to 96270
0.4 to 1460
15 to 8328
0.2 to 83
2.2 to 124
3 to 5140
130 to 13800
0.58 to 458
4 to 488
0.2 to 30
0.5 to 171
       Source: Schroeder et al., 1987.
 1     concentration ranges of various metals found in remote, rural, and U.S. urban areas as
 2     common specific airborne PM constituents.
 3          Chapter 6 also includes a section on acid aerosols.  Acid aerosols are secondary
 4     pollutants formed primarily through oxidation of sulfur dioxide (SO2), a gas emitted by the
 5     combustion of fossil fuels. Oxidation of SO2 forms sulfuric acid (H2SO4), the major
 6     component of acid aerosols.  Sulfuric acid is formed to a lesser extent through the oxidation
 7     of sulfur species (H2S and CH3SCH3) from natural sources.  H+ is found in the fine particle
 8     size fraction i.e., particles with aerodynamic diameter (Dp) < 1.0 pirn. Although recent
 9     research has shown a typically high correlation between 804 and acidity, data from
10     summertime sampling have shown that SO4= is not always a reliable predictor of H+ for
11     individual events at a given site.
12          A major determinant of the lifetime of H+  in the atmosphere is the rate of
13     neutralization by ammonia (NH3).  Ammonia reacts with H2SO4 to form ammonium sulfate
14     [(NH4)2SO4] and ammonium bisulfate (NH4HSO4).  The major sources of ammonia in the
       April 1995
1-12
DRAFT-DO NOT QUOTE OR CITE

-------
 1      environment are animals and humans.  In North America, ambient concentrations of H+ tend
 2      to be regional in nature with the highest concentrations found in the northeastern United
 3      States and southwestern Canada.  Recent research has shown that regional transport is
 4      important to acid sulfate concentrations, as elevated levels of ambient H+ were measured
 5      simultaneously during a regional episode at multiple sites located from Tennessee to
 6      Connecticut.
 7           Recent work has suggested that ultrafine particles may be responsible for some of the
 8      health effects associated with exposure to paniculate matter, leading to an interest in the
 9      number concentration of ambient particles.  A Chapter 6 section examines data on particle
10      number concentration and the relationship between particle number and particle mass or
11      volume.   In some  situations the ultrafine mode can be the dominant size range for selected
12      components of atmospheric aerosol particles.  One example is the case of metallic aerosols
13      for which fine particles (Dp<2.5/>im) concentrations can be dominated by the ultrafine mode
14      despite the strength of the processes that tend to remove particles from this mode.  While
15      there is consensus that ultrafine metals are abundantly produced and emitted into the
16      atmosphere,  there  are not much data on ambient concentrations of ultrafine metals.  The few
17      direct measurements available can be extended with some confidence using indirect methods;
18      i.e., from particle counting techniques that have size information but no chemical
19      information, or from filter collection methods that have limited size information but detailed
20      compositional information.
21
22
23      1.7   EXPOSURE:   AMBIENT AND INDOOR
24           Chapter seven focuses on studies which include information on measurement of
25      simultaneous personal PM exposures,  indoor-residential PM concentrations, and ambient PM
26      concentrations.  The literature on concentrations of PM in indoor settings are also presented.
27           For any air pollutant, the total exposure of an individual consists of a variety of
28      sequential exposures to a variety of microenvironments.  They are typically, outdoor, indoors
29      at-home, at-work,  in-traffic, and many other indoor microenvironments.  The principle of
30      superposition is a useful mechanism to visualize the summation process.  For any identified
31      air pollutant, the ambient environment is one source of indoor pollution due to air exchange

        April 1995                               M3       DRAFT-DO NOT QUOTE OR CITE

-------
 1     and infiltration.  Whether the ambient is a significant or dominant source of indoor pollution
 2     depends on the relative strength of indoor sources and sinks.
 3          Paniculate matter studies  have detected a "personal cloud"  related to the activities of an
 4     individual which may generate  significant levels  of airborne PM in his/her vicinity which
 5     may not be picked up by an indoor PM monitor at a distance. Other studies have identified
 6     significant sources in the home, e.g.  cooking and smoking.
 7          In PM of size fractions that include coarse particles, some  studies have identified
 8     statistically significant  relationships between personal exposures and particle concentrations
 9     from fixed-site ambient or indoor monitors, and other studies have not, probably due to
10     overwhelming effects of indoor sources, "personal clouds" and other individual activities.
11     For PM of a fine size  fraction  - such as sulfates, there seems to be more of a relationship
12     between ambient concentration and personal exposure, than for coarser PM, perhaps because
13     of the  ability of fine PM to penetrate into indoor settings.
14          For  a study population in which there is a detectable correlation between personal
15     exposures and ambient concentrations,  the ambient concentration can predict the mean
16     personal exposure with much less uncertainty than it can predict the personal exposure of any
17     given individual in the population.
18          The three largest studies of indoor air particles in the U.S. have all found that the
19     single  largest indoor source of fine (PM3 5 or PM2.5) particles is cigarette smoke.  The
20     estimate of the impact of smoking on a home PM levels ranges from about 30 to 45  /*g/m3,
21     and of a single cigarette from 1 to 2 jug/m3 for a 24-h period. Homes  without smoking have
22     indoor particle concentrations (both PM10 and PM2 5) that are sometimes below and
23     sometimes above the outdoor levels.  At low outdoor levels (as in most of the cities in the 6-
24     City and New York State studies) indoor concentrations are generally higher—at high
25     outdoor levels, they are slightly lower.  Indoor concentrations are considerably higher during
26     the day, when people are  active, than at night.
27          The second largest source of indoor particles, as determined by the PTEAM Study, is
28     cooking.  Estimates  of the effect of cooking ranged from about 10 to 20  /xg/m3.  A few small
29     studies confirm the effect of cooking on indoor particle levels, both PM10 and PM2 5.  The
30     two other large-scale studies did not directly test for the effect of cooking,  although the
        April 1995                               1-14       DRAFT-DO NOT QUOTE OR CITE

-------
  1     residual indoor concentrations in multivariate calculations led the authors to suggest that
  2     cooking could have contributed to the residual.
  3          Kerosene heater use was determined to contribute about 15 /^g/m3 to indoor
  4     concentrations in one county but not the other in the New York State study.   Also, a later
  5     effort using mass-balance calculations was unsuccessful in determining a contribution to
  6     particle mass from kerosene heater use  in either county, although a somewhat smaller set of
  7     homes may have been responsible for this result. Gas stoves,  wood stoves, and fireplaces
  8     were found to have no noticeable impact on total concentrations of particles, although many
  9     studies show an increase in PAH concentrations and some show an increase in mutagenicity
 10     of indoor air due to these combustion sources.
 11          Vacuuming, dusting, and sweeping were found to contribute slightly but with doubtful
 12     significance to indoor levels in the PTEAM Study.  House volume had a  significant but small
 13     effect on particle concentrations,  with values of —1 to —2 /zg/m3 per 1,000 cubic feet. Air
 14     exchange rates were also significant at times, but with different impacts depending on the
 15     relative indoor and outdoor concentrations—at high outdoor concentrations, increased air
 16     exchange resulted in increases in the indoor air particle levels.
 17          Unknown indoor sources were found to account for a substantial fraction (25%) of
 18     indoor concentrations (both PM2 5 and PM10) in the PTEAM study.  This suggests a need for
 19     further research to determine the source or sources of these particles.
 20          Decay rates for fine (PM2 5) particles were determined to be about 0.4 h"1 compared to
 21     1 h'1 for coarse particles,  with an intermediate estimate of 0.65 h"1 for PM10.   For a home
 22     with no indoor sources and  a typical air exchange rate  of about 0.75, this would imply that
 23     fine particles indoors would be about  0.757(0.4+0.75) = 65% of the outdoor value at
 24     equilibrium, indoor PM10 would be about 54%  of outdoor levels,  and indoor coarse  particles
 25     would be about 43% of outdoor levels.  Since few homes were  observed to have
 26     concentrations  this  low, it can be inferred that few homes are free of important indoor
27     sources of particles.
28         Studies in buildings also indicated  that smoking was the major indoor source of
29     particles, with a geometric mean of 44 versus 15 /ig/m3 (arithmetic mean  of 70 versus
30     18 /*g/m3) observed for smoking versus  nonsmoking areas in 38 Pacific Northwest buildings.
       April 1995                                1_15       DRAFT-DO NOT QUOTE OR CITE

-------
 1     This difference of 29 to 52 /xg/m3 is similar to the difference of 30 to 45 /^g/m3 estimated
 2     from the three major studies of U.S. homes.
 3          Indoor air quality models have been employed with increasing success to estimate
 4     source emission rates and particle decay rates. Cigarettes smoked in homes with normal
 5     activities emit about 14 mg/cigarette, a result that agrees well with various chamber studies
 6     with smokers or smoking machines.  Cooking was estimated to emit 4 mg/min,  a result that
 7     needs confirmation.  Elemental emission profiles have been determined for both smoking and
 8     cooking, with potassium  and chloride being major markers for smoking, and iron and
 9     calcium for cooking.  Particle decay rates have been estimated for homes to range between
10     0.4 and 1.0 h"1.  Studies in telephone equipment buildings and museums have established
11     particle deposition velocities for sulfates and other ions,  although differences in the estimates
12     suggest that further research is needed.
13
14
15     1.8   EFFECTS ON VISIBILITY AND CLIMATE
16          Chapter eight discusses  factors affecting visibility,  ways to measure it, historical trends,
17     and methods to determine its  value.  Paniculate matter effects on climate are also discussed.
18     Much of the information contained in the section on visibility is a summary of information
19     from the previous 1982 Criteria Document for Paniculate Matter and Sulfur Oxides.
20          Traditionally, visibility has been defined in terms of the distance from an object that is
21     necessary to produce a minimum detectable contrast between that object and its background.
22     Although visibility is often defined by this "visual range," it includes not only being able to
23     see or not see a target, but also seeing  targets at shorter distances and appreciating the details
24     of the target, including its colors.  Visibility impairment can manifest itself in two ways:  (1)
25     as a layer of haze (or a plume), which is visible because it has a visual discontinuity between
26     itself and its background, or (2) as a uniform haze which reduces atmospheric clarity.  The
27     type and degree of impairment  are determined by the distribution, concentrations,  and
28     characteristics of atmospheric particles and  gases, which scatter and absorb light traveling
29     through the atmosphere.  Scattering and absorption  determine light extinction.
30          On a regional scale, the extinction of light is generally dominated by  particle scattering.
31     In urban areas, absorption by particles  becomes important and occasionally dominant.

       April 1995                                1-16       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Extinction by particles is usually dominated by particles of diameter 0.1 to 2 pm (fine par-
  2      tides).  In general,  scattering by particles accounts for 50 to 95% of extinction, depending
  3      on location,  with urban sites in the 50 to 80% range and nonurban sites in the 80 to 95%
  4      range.
  5           Available visibility monitoring methods measure different aspects of visibility
  6      impairment.   Generally, contrast-type measurements  (such as photography, telephotometry,
  7      and human eye observations) relate well to the perception of visual  air quality,  while
  8      extinction or scattering measurements (such as transmissometry and nephelometry) relate to
  9      the cause of visibility degradation. The above measurement methods can be used to approx-
10      imate visual  range.
11           Current knowledge  indicates that fine paniculate matter is composed of varying
12      amounts of sulfate,  ammonium, and nitrate ions, elemental carbon,  organic carbon
13      compounds,  water,  and smaller amounts of soil dust, lead compounds, and trace species.
14      Sulfate often dominates the fine mass and light scattering, while elemental carbon is the
15      primary light-absorbing species.  Ammonium ion typically accounts for 5 to 15%  of the fine
16      mass and often correlates well with sulfate levels.  Mean nitrate concentrations  can  represent
17      up to 37% of the total fine particle mass  in urban cities.
18           Visibility has value  to individual economic agents primarily through its impact upon
19      activities of  consumers and producers.  Most economic studies of the effects of air pollution
20      on visibility  have focused on the aesthetic effects which are believed to be the most
21      significant economic impacts of visibility degradation caused by  air  pollution in the  U.S. It
22      is well established that people notice changes in visibility and that visibility conditions affect
23      the well-being of individuals.
24           Paniculate matter of submicron size in the  earth's atmosphere  perturbs the radiation
25      field. There is no doubt  that anthropogenic aerosols have the potential to affect climate; the
26      question is by how much. There are two chief avenues through which aerosols impact the
27      radiation budget of the earth. The direct effect is that of enhanced reflection of solar
28      radiation by  particles in a cloud-free atmosphere. Since aerosols, even those containing some
29      absorptive component, are primarily reflective, their  impact is felt as a negative radiative
30      forcing (i.e., a cooling) on the climate system.  Although there is some uncertainty  in the
31      global distribution of such aerosols and in the chemical and radiative properties of the

        April 1995                                 1_17      DRAFT-DO NOT QUOTE OR CITE

-------
 1     aerosols, the radiative effects can still be modeled within certain bounds.  Estimates of this
 2     forcing range from — 0.3 W m"2 to about twice that value for current conditions over pre-
 3     industrial times.
 4           The indirect forcing results from the way in which aerosols,  in their role as cloud
 5     condensation nuclei (CCN), affect cloud microphysical properties.  The most important is the
 6     effective radius of cloud droplets,  which decrease as CCN concentrations increase.   This
 7     effect is most pronounced when the concentration, N, is very low, and clouds are moderately
 8     reflective.  Other effects are the enhancement of  cloud lifetimes and also changes in the
 9     nucleating ability of CCN through chemical changes.  Although estimates of the indirect
10     effect are uncertain by at least a factor of 2,  it appears to be potentially as  important as the
11     direct effect.  On a global mean basis, anthropogenic emissions of anthropogenic aerosols
12     could have offset substantially the positive radiative forcing  due to greenhouse gas emissions.
13           The one crucial difference between aerosol  forcing and greenhouse (gas) forcing is the
14     atmospheric lifetime of aerosols and gases and hence, forcing.  The aerosol forcing is fairly
15     regional, whereas the greenhouse forcing is global. One should, therefore, expect
16     inter-hemispheric differences in the forcing and perhaps climate response.  However, climate
17     models are not currently at  the level of sophistication needed to determine  climate response
18     unambiguously.  Global observations of surface temperature cannot separate natural and
19     anthropogenic causal mechanisms, with few exceptions.
20
21
22     1.9   EFFECTS  ON MATERIALS
23           Chapter nine briefly discusses the effects of paniculate matter exposure on the aesthetic
24     appeal and physical damage to  different types of  building materials and economic
25     consequences, including background information  on the physics and chemistry of atmospheric
26     corrosion. Where possible, the chapter discusses only those effects associated with particle
27     exposure;  however, most of the data are on the effects of particles in combination with SO2.
28           A significant detrimental effect of particulate matter pollution is the soiling of painted
29     surfaces and other building  materials.  Soiling is  defined as  a degradation mechanism that can
30     be remedied by cleaning or washing, and depending on the  soiled  surface,  repainting.
31     Available data on pollution  exposure indicates that particulate matter can result in increased

       April 1995                                1-18      DRAFT-DO NOT QUOTE OR CITE

-------
  1      cleaning frequency of the exposed surface, and may reduce the life usefulness of the material
  2      soiled.  Data on the effects of particulate matter on other surfaces are even less well
  3      understood.  Some evidence also shows damage to fabrics, electronics, and works of art
  4      composed of one or more materials, but this evidence is largely qualitative and sketchy.
  5           The damaging and soiling of materials by airborne pollutants have an economic impact,
  6      but this impact is difficult to measure.  The accuracy of economic damage functions is
  7      limited by several factors.  One of the problems has been to separate costs related to
  8      particulate matter-related materials  from other pollutants, as well as from those related to
  9      normal maintenance.  Cost studies typically involve broad assumptions about the kinds of
 10      materials that are exposed in a given area and then require complex statistical analysis  to
 11      account for a selected number of variables.  Attitudes regarding maintenance may vary
 12      culturally, further confounding the problem of quantifying economic impact.
 13           The nature and extent of damage to materials by particulate matter have been
 14      investigated by field and laboratory studies. Both physical and economic  damage functions
 15      have been developed for specific damage/effect parameters associated with exposure to
 16      particulate matter. To date, only a few of these functions are relatively reliable in
 17      determining damage, while none has been generally accepted for estimating costs.
 18           In recent years, fairly reliable damage functions for soiling of exterior wall paints have
 19      been developed.  The available damage functions are few in number but represent a major
 20      fraction of the total surface that is exposed and sensitive to pollution damage.
 21           Although there still remains a lack of sensitive materials distribution data, the
 22      geographic  resolution of available data is about as good as that of environmental monitoring
 23      data.  These limitations may hinder accurate estimates of total material damage and soiling,
 24      but they do not prevent estimates within ranges of error. Studies have used various
 25      approaches  to determine pollutant-related costs for extra cleaning, early replacement, more
 26      frequent painting, and protective coating of susceptible  materials, as  well as other indicators
27      of the adverse economic effects of pollutants.  No study has produced completely satisfactory
28      results, and estimates of cost vary widely.
29
30
        April 1995                                1-19      DRAFT-DO NOT QUOTE OR CITE

-------
  1      1.10 DOSIMETRY MODELING OF  INHALED PARTICLES IN THE
  2            RESPIRATORY TRACT
  3           Particles are deposited in the respiratory tract by mechanisms of impaction,
  4      sedimentation, interception, diffusion, and electrostatic precipitation.  Ventilation rates differ
  5      for various activity patterns in humans, for different ages, and among species.  These
  6      ventilation differences coupled with differences in upper respiratory tract structure and in
  7      size, branching pattern, and structure of the lower respiratory tract among species and
  8      healthy versus diseased states result in significantly different patterns  of airflow that in turn
  9      affect particle deposition in the respiratory tract regions.  For a given aerosol, the two most
10      important parameters determining deposition are the mean aerodynamic diameter and the
11      distribution of particles about that mean.  Subsequent clearance of a deposited dose is
12      dependent on the initial site of deposition, physicochemical properties of the  particles (e.g.,
13      dissolution half-time), and on time since deposition.
14           An accurate description  of the exposure-dose-response relationship for the observed
15      health effects of PM should account,  to the extent possible, for these  mechanistic
16      determinants of particle disposition.  Deposited dose may be an appropriate metric for
17      "acute" effects, (e.g.,  mortality), especially if the particles exert their primary action on the
18      surface contacted.  An alternative to consider is dose rate (jwg/min) per unit surface area
19      because insoluble particles deposit and clear along the surface of the respiratory tract.
20      "Chronic" effects (e.g., certain types of morbidity) may  be better described by retained dose
21      estimates because clearance is affected by the time since deposition and the aerosol solubility,
22      characterized by dissolution-absorption half-times.
23           The human model chosen to make deposited and retained dose predictions is a semi-
24      empirical compartmental model that is able to describe particle deposition and clearance by
25      three routes (absorption into blood, transport to gastrointestinal tract,  and transport to
26      lymphatics). Two different models were used to model particle deposition and clearance in
27      laboratory animals.
28           The predictions of deposited and retained doses  show anticipated differences due to the
29      influence of aerosol particle diameter and distribution, minute ventilation, and species-
30      specific morphometry.  For example, mouth breathing alters the deposition fraction of
31      ambient aerosols in the tracheobronchial and alveolar regions when compared to nasal

        April 1995                                1_20       DRAFT-DO  NOT QUOTE OR CITE

-------
 1      breathing.  The differences shown in the these predictions for deposition point to the
 2      importance of characterizing the differences between genders and the impact of age on
 3      deposition.  The chosen model has predicted differences between children of 1 year and
 4      adults across particle diameters ranging from the diffusion to aerodynamic range of
 5      approximately 2.5-fold in the tracheobronchial region and 2-fold in the alveolar region.  The
 6      direction and magnitude are a function of aerosol particle diameter and distribution.
 7      Differences in ventilation and  morphometry for diseased  states can also be anticipated and
 8      recent studies and other model predictions show an increased deposition in subjects with
 9      chronic  obstructive pulmonary disease.
10           The various species  used in inhalation toxicology studies that serve as  the basis for
11      exposure-dose-response assessment do not receive  identical doses in a comparable respiratory
12      tract region when exposed to the same aerosol.  Such interspecies differences are important
13      because the toxic effect is likely more related to the quantitative pattern of deposition within
14      the respiratory tract than to the exposure concentration; this pattern determines not only the
15      initial respiratory tract tissue dose but also the specific pathways by which the inhaled
16      material is  cleared and redistributed.  Thus, accounting for differences in dosimetry can
17      change the apparent effect levels among species.  To illustrate, for the same aerosol of 0.5
18      jim MM AD and a of 1.3, using deposition normalized to surface area for an effect in  the
                         6
19      tracheobronchial region, an exposure concentration of  100 /xg/m3 to  rats and guinea pigs
20      would predict a human equivalent exposure  concentration of 939 and 79 /ig/m3, respectively,
21      assuming species sensitivity to the deposited tracheobronchial dose were equal. However,
22      for chronic exposures to the same aerosol, retained alveolar dose (/-ig/g lung tissue) may be
23      more appropriate as a dose metric.  Assuming it is a relatively insoluble  aerosol (i.e.,
24      assuming a dissolution-absorption half-time of 1,000 days), a human equivalent exposure
25      concentration would be predicted to be 22 and 784 /xg/m3 based on the rat versus guinea pig,
26      respectively.
27           These examples  show that relevance of a particular animal model should be considered
28      together with dosimetry and the  appropriateness of the metric for a given health endpoint.  In
29      general, the objective  should be  to provide a metric that is mechanistically motivated by the
30      observed health effect of interest for extrapolation.
        April 1995                                1-21       DRAFT-DO NOT QUOTE OR CITE

-------
 1          Dosimetry modeling at the moment can address important mechanistic factors of
 2     particle deposition and clearance including the aerosol particle diameter and distribution,
 3     intra and interspecies differences in deposition as a function of ventilation and morphometry,
 4     and intra and interspecies differences in clearance rates.  Use of dosimetry modeling and
 5     judicious choice of appropriate dose metrics should be used to interpret the observed health
 6     effects data related to PM10 exposures.
 7          Further,  these predictions were  based on the use of mass as the exposure metric.
 8     Recent data suggest that particle number,  or possibly particle surface area, may be a more
 9     appropriate exposure metric because the fine mode aerosols are small in mass but have
10     extremely high concentrations of particle numbers. Also, normalizing factors such as
11     number of alveoli  or number of macrophages may be more appropriate for certain
12     pathogenesis mechanisms.  Creating these dose metrics for various species will depend on the
13     availability of morphometric information.
14
15
16     1.11  TOXICOLOGY OF PARTICIPATE MATTER CONSTITUENTS
17          Chapter eleven reviews results on exposure to specific PM constituents, based on
18     controlled human clinical studies, selected occupational studies, and animal toxicology
19     studies.  It focuses on those studies published since the 1982 PM Criteria Document and
20     includes coverage  of specific PM  species  selected for discussion based on their being
21     commonly present in ambient aerosols at concentrations > 1 ng/m3.
22          Paniculate matter is a broad term that encompasses thousands  of chemical species,
23     many of which have not been investigated in controlled animal or human studies.  However,
24     even a full discussion of all the types of particles that have been studied is well beyond the
25     scope of the chapter.  Most of the animal toxicological and occupational epidemiological
26     studies summarized used very high paniculate concentrations,  relative to ambient, even when
27     animal-to-human dosimetric differences are considered.  In spite of  these difficulties, the
28     array of animal studies does illustrate certain toxicological principles for particles.  To
29     identify but a few  here, the data base clearly shows that the site of respiratory tract
30     deposition (and hence particle size) clearly influences the health  outcome and that toxicity is
31     dependent on the chemical species.

       April 1995                               1-22      DRAFT-DO NOT QUOTE OR CITE

-------
 1     Effects of Controlled Exposure to Acid Aerosols
 2          The effects of acid aerosols are primarily related to strong acidity (i.e., H+ content).
 3     That is, H2SO4 is more potent than NH4HSO4 which is more potent than NH4(H)2SO4.  The
 4     size of acid aerosols also affects their potency, but the exact relationships are dependent upon
 5     the endpoint being examined.
 6         Sulfuric acid affects pulmonary function. Healthy subjects are only affected by very high
 7     levels of acute exposure (around 2,000 ng/m3), even if they exercise and gargle with acidic
 8     material to reduce  neutralization by oral ammonia. Asthmatics, especially adolescent
 9     asthmatics, appear to be more sensitive. For example, a few, but not all studies, found that
10     an acute exposure  to around 70 ^g/m3 caused small  decrements in the pulmonary function of
11     adolescent  asthmatics.
12         Sulfuric acid affects mucociliary  clearance in humans and animals. The direction (i.e.,
13     increase or decrease) and the magnitude of the effect is dependent on the concentration and
14     duration of exposure, as well as the specific region of the lung being measured. Humans
15     exposed to levels of H2SO4 as low as 100 jiig/m3 (Ih) experienced a decrease in mucociliary
16     clearance.  Animal studies have shown that H2SO4 can also affect alveolar clearance.
17        Chronic exposure to H2SO4 causes a variety of structural changes in the lung. For
18     example, mucus-secreting cells are affected and can be found in deeper regions of the lung
19     than usual.  Pulmonary function is also altered.
20         Several studies have sought to define interactions of acids with  other pollutants,
21     especially ozone.   Work with animals has demonstrated additivity, synergism, and
22     antagonism, depending upon the species, exposure, and endpoint.  More  recent human
23     clinical studies found that acute exposure to  100  ptg/m3 H2SO4 may potentiate the response
24     to ozone on pulmonary function.
25
26     Complex Mixtures
27         The 1982 Air  Quality Criteria Document for Particulate Matter and Sulfur Dioxide
28     concluded  from its review of studies on the genotoxicity and carcinogenicity of atmospheric
29     particles that "all the major types  of  airborne paniculate matter may contain adsorbed
30     compounds that are mutagenic and/or carcinogenic to animals and may contribute in some
31     degree to the human cancer associated with exposure to urban  air pollution."  Recent

       April 1995                                1-23       DRAFT-DO NOT QUOTE OR CITE

-------
 1      research activity has added data that support this conclusion, but do not warrant that it be
 2      changed significantly.
 3          The direct relevance of the evidence for the mutagenicity and tumorigenicity of extracts
 4      of particulate matter in experimental systems to exposure scenarios experienced by humans is
 5      uncertain at this time.  Recent analytical epidemiological studies, that adjusted for tobacco
 6      smoking and other major potential risk factors, have found a weak to non-existent association
 7      between human lung cancer and indices of exposure to air pollution including particulate
 8      matter.  Most investigators believe that the epidemiological evidence obtained thus far does
 9      not substantiate causality, although the hypothesis remains credible.
10
11      Diesel Emissions
12           Acute toxic effects caused by exposure to diesel exhaust are mainly attributable to the
13      gaseous components (i.e., mortality from carbon monoxide intoxication and lung injury from
14      respiratory irritants).  When the exhaust  is diluted to limit the concentrations of these gases,
15      acute effects are not seen.
16           Ten different long-term (> 1 year) animal inhalation studies of diesel engine emissions
17      have been conducted. The focus of these studies has been on respiratory tract effects in the
18      alveolar region.  Effects in the  upper respiratory tract and in other organs were not found
19      consistently in chronic animal exposures.  The pathogenic sequence following the inhalation
20      of diesel exhaust as determined histopathologically and biochemically begins with the
21      phagocytosis of diesel particles  by alveolar macrophages (AMs).  These activated
22      macrophages release chemotactic factors  that attract neutrophils and additional AMs. As the
23      lung burden of diesel particles increases, there is an aggregation of particle-laden AMs in
24      alveoli adjacent to terminal bronchioles, increases in the number of Type 2 cells  lining
25      particle-laden alveoli, and the presence of particles within alveolar and  peribronchial
26      interstitial tissues and associated lymph nodes.  The PMNs and macrophages release
27      mediators of inflammation and oxygen radicals and particle-laden AMs  are functionally
28      altered resulting in decreased viability and impaired phagocytosis and clearance of particles.
29      There is a substantial body of evidence for an impairment of particulate clearance from  the
30      bronchio-alveolar region of rats  following exposure to diesel exhaust.  The latter series of
        April 1995                                1-24       DRAFT-DO NOT QUOTE OR CITE

-------
  1      events may result in the presence of pulmonary inflammatory, fibrotic, or emphysematous
  2      lesions.
  3           The noncancer toxicity of diesel emissions is considered to be due to the particle rather
  4      than the gas phase,  since the long-term effects seen with whole diesel are not found or are
  5      found to a much lesser extent in animals exposed  to similar dilutions of diesel exhaust
  6      filtered to remove most of the particles. Chronic  studies in rodents have  demonstrated
  7      pulmonary effects at 200 to 700 jug/m3. No-effect levels have been reported ranging from 60
  8      to 260 Mg/m3.
  9           Several epidemiologic studies have evaluated the effects of chronic exposure to diesel
 10      exhaust on occupationally exposed workers.  None of these studies are useful for a
 11      quantitative evaluation of noncancer toxicity because  of inadequate exposure characterization,
 12      either because exposures  were not well defined or because the possible confounding effects  of
 13      concurrent exposures could not be evaluated.
 14           The U.S. Environmental Protection Agency  has developed a draft qualitative and
 15      quantitative cancer assessment for diesel emissions.  The summary to follow was drawn from
 16      that document.  This draft is currently undergoing external review by  the public and the
 17      Clean Air Scientific Advisory Committee.  As a result of limited evidence from
 18      epidemiological data, supported by adequate evidence for carcinogenicity  of diesel engine
 19      emissions in animal studies, as well as positive evidence for mutagenicity, it was concluded
 20      that diesel engine emissions best fit into cancer weight-of-evidence Category Bl.  Diesel
 21      engine emissions are thus considered to be probable human carcinogens.  This is in
 22      agreement with a 2A classification by the International Agency  for Research on Cancer.
 23           Using a dosimetry model that accounted for  animal-to-human differences in lung
 24      deposition efficiency, lung particle clearance rates, lung surface area,  ventilation, metabolic
 25      rate, as well as elution rates of organic chemicals  from the particle surface, equivalent human
 26      doses were calculated  on the basis of particle concentration per unit lung  surface  area.
27      Following dosimetric adjustment, risk estimates were derived using a linearized multistage
28      model. A unit risk  estimate of 3.4 x 10"5 (the upper 95% bound of the cancer risk from
29      lifetime exposure to 1  /ig/m3 diesel paniculate  matter) is recommended.  This estimate is
 30      based on the geometric mean of estimates derived from three separate animal bioassays using
31      Fischer 344 rats.

        April 1995                                 1-25       DRAFT-DO NOT QUOTE  OR CITE

-------
 1     Metals
 2          A summary of the health effects of airborne metals follows. The descriptions are
 3     qualitative in nature.  Most of the literature on these compounds deals with high
 4     concentration animal toxicological studies or occupational epidemiological studies.
 5          The primary target for inhaled aluminum compounds is the respiratory tract.
 6     Commonly reported symptoms include asthma, cough, and decreased pulmonary function;
 7     fibrosis has also been reported. Laboratory animal studies support findings from human
 8     studies that aluminum acts via an irritant, rather  than by an allergic, mechanism.
 9          The respiratory tract is the primary target organ for antimony (trioxide) following
10     inhalation exposure. Respiratory effects have been reported in workers chronically exposed
11     to mg levels of antimony dust.  Other reported effects include altered ECGs, gastrointestinal
12     symptoms, ocular and dermal effects, and reproductive effects.
13          The toxicity data on inhalation exposures to arsenic are limited.  Long-term
14     occupational exposure leads to lung cancer and causes skin changes and peripheral nerve
15     damage in workers. Respiratory tract tumors occurred in hamsters exposed  to intratracheal
16     doses of arsenic combined with a carrier dust.
17          Data on barium are extremely limited, with no epidemiological data available and no
18     standard inhalation toxicity studies in animals. Occupational case studies, supported by
19     histopathological studies in rats, indicate that the respiratory tract is a target for barium
20     compounds.
21          The kidney and the respiratory  tract are the primary target organs for cadmium by
22     inhalation exposure in the human; toxicity is dependent  on cumulative exposure, with renal
23     tubular dysfunction and associated increased excretion in urine of proteins,  amino acids, and
24     essential metals  being key outcomes of long-term exposures.  Acute high-level exposure in
25     humans causes intense respiratory tract irritation, and milder effects on pulmonary function
26     follow chronic low-level exposure.  Rat studies show that cadmium can cause lung cancer;
27     there is evidence that lung cancer has been observed in humans following high occupational
28     exposure.
29          The respiratory tract is the primary target for inhaled chromium compounds. Human
30     and animal data agree on the nature of nasal effects.  Laboratory animal studies have
31     reported lung  lesions and evidence of inflammation.  Human and animal data agree that

       April 1995                               1-26       DRAFT-DO NOT  QUOTE OR CITE

-------
  1      Cr(VI) compounds cause lung cancer.  Human studies have also reported early signs of renal
  2      damage with exposure to Cr(VI) compounds.
  3           The respiratory tract is the major target for inhaled cobalt compounds. In humans,  two
  4      major types of effects are observed, interstitial lung disease and asthma. Inflammation and
  5      decreased lung function have been observed in laboratory animals.
  6           Limited data support the respiratory system as a major target for inhaled copper and
  7      copper compounds.  In humans the data  are limited to subjective reporting of respiratory
  8      symptoms and radiographic evidence of pulmonary effects after acute and chronic inhalation.
  9           The respiratory tract is the primary target for iron oxides by inhalation exposure.  In
10      humans, respiratory effects  have been reported in workers chronically exposed to iron dust.
11      In laboratory  animals, hyperplasia and alveolar fibrosis have been reported  after inhalation or
12      intratracheal administration  of iron oxide.
13           The nervous system is the most sensitive target for elemental mercury following acute
14      or chronic inhalation exposures.  Effects range from reversible neurological symptoms to
15      psychomotor and neurobehavioral changes and peripheral nerve dysfunction.  Respiratory,
16      gastrointestinal, and cardiovascular symptoms have also been reported in case reports and
17      occupational studies with exposure  to high concentrations of mercury.  The kidney is a
18      sensitive target toxicity following elemental mercury exposure in humans.
19           The nervous system and the respiratory tract are primary targets for inhaled manganese.
20      Acute occupational exposures are associated with pneumonitis, while chronic exposures
21      mainly impact the central nervous system. Limited information suggests that prenatal and/or
22      postnatal exposure of laboratory rodents  to inhaled manganese oxide may depress
23      neurobehavioral activity.
24           Limited data on the inhalation of magnesium and its compounds support the respiratory
25      tract as a target.  Acute high-level exposure of humans or laboratory animals to magnesium
26      oxide fume results in a reaction similar to zinc oxide metal fume fever.  Suggestive evidence
27      indicates chronic exposure to magnesium dusts may produce pneumoconiosis. In laboratory
28      animals, fibrosis is observed with chronic exposure to high levels of magnesium dusts.
29           The respiratory tract appears to be the main target in humans and animals after
30      inhalation exposure to molybdenum compounds; however, inhalation exposure to
31      molybdenum has also been associated with nonspecific effects in humans including general

        April 1995                                1-27       DRAFT-DO NOT QUOTE OR CITE

-------
  1      weakness and dizziness.  Animal inhalation studies indicate that toxicity varies with the
  2      molybdenum compound.
  3           The respiratory tract is the primary target for nickel compounds following inhalation
  4      exposure.  In humans, respiratory effects include asthma and altered pulmonary function.  In
  5      laboratory animals,  inflammatory responses suggest an immunological response in the lungs.
  6      Immunological changes have been reported in refinery workers exposed to nickel.  The
  7      potential for lung and nasal cancer was evident in occupational and laboratory animal studies.
  8           When ingested in relatively low concentrations, potassium is an essential metal, but
  9      available data on inhaled  potassium compounds are insufficient to assess toxicity.  Similarly,
10      with inhalation exposure, the respiratory tract is the target for selenium, another essential
11      metal.  In humans, respiratory effects have been reported in workers chronically exposed to
12      selenium; and similar effects have been reported in laboratory animal studies.
13           Inorganic tin is relatively inert lexicologically, and effects are limited to mild
14      respiratory effects, along  with the formation of radio-opaque nodules in the lungs.  No other
15      target systems for inhalation exposure to inorganic tin have been reported. Limited data
16      indicate the nervous, hepatic,  renal,  and respiratory systems are targets for inhalation
17      exposure to organotin compounds.  The respiratory tract is the primary target for titanium
18      following inhalation exposure.  No histopathology of other organs was found in rats
19      chronically exposed to titanium tetrachloride at up to  6,000 /igTi/m3.   Titanium is not
20      translocated in the body,  even with chronic exposure  and high concentrations. Titanium
21      dioxide inhalation results  in pneumoconiosis in humans and signs of inflammation in
22      laboratory animals.
23           The respiratory tract is the primary target for inhaled vanadium compounds.  Vanadium
24      damages alveolar macrophages, and  toxicity is  related to compound solubility and  valence.
25      Human occupational case studies and epidemiological studies indicate symptoms of
26      respiratory distress.  Symptoms of systemic effects have been observed following chronic
27      occupational exposure and in laboratory animal studies.
28           Following inhalation exposure, the respiratory tract is the primary target for  zinc,
29      another essential metal.  In humans, metal fume fever, characterized by respiratory
30      symptoms and pulmonary dysfunction, was observed  in workers and experimental subjects
        April 1995                                1_28       DRAFT-DO NOT QUOTE OR CITE

-------
 1      during acute exposures to high levels of zinc oxide. Zinc can produce inflammatory
 2      responses in both human and animal species.
 3
 4      Silica
 5           Silica can occur in two chemical forms, amorphous and crystalline.  Crystalline forms
 6      include quartz,  which is the most prevalent; cristobalite; tridymite; and a few other rare
 7      forms. Freshly fractured crystalline silica is more lexicologically reactive than aged forms of
 8      crystalline silica or forms that may be coated with other chemical compounds.  Amorphous
 9      silica  is less well studied and may have similar toxic endpoints but is less potent than
10      crystalline silica.  With sufficient exposure,  crystalline silica is toxic to the respiratory
11      system.  Acute  high exposure in both humans  and animals causes lung inflammation and,  if
12      the exposure is  high enough, rapid onset of a fibrotic lung disease (acute silicosis) which can
13      be fatal.  Occupational studies show that chronic exposure to crystalline silica causes
14      inflammation of the lung which is followed by fibrosis and a human fibrotic disease called
15      silicosis which can lead to early mortality.  Silocotic individuals are also at higher risk for
16      other  diseases, e.g., tuberculosis. Some occupational studies also show a concurrent
17      incidence of lung cancer.
18           The role,  if any, of silica-induced lung inflammation, fibrosis, and silicosis in the
19      development of lung  cancer is hypothesized but not adequately demonstrated.  Crystalline
20      silica  interaction with DNA has been shown. Chronic exposure  animal studies  in rats also
21      show  a similar pattern of lung inflammation, fibrosis, and lung cancer.  In 1987, the
22      International Agency for Research on Cancer classified crystalline silica as a "possible"
23      human carcinogen owing to a sufficient level of evidence in animal studies but  inadequate
24      evidence in human studies.  The evidence for amorphous silica carcinogenicity  was said to be
25      inadequate for both humans and animals, placing  it in Group 3 (agent not classifiable).
26      While active surveillance of the U.S. population for fibrosis and silicosis is not standard
27      practice, U.S. health statistics do  not reveal a general population increase of crystalline silica
28      diseases.  However, there is an increase in  these diseases among the occupational work
29      force.
30           An assessment of the  occupational  risk of silicosis was made using recent studies from
31      South Africa and Canada, both of which examined medical histories of over 2000 miners.

        April  1995                                 1-29       DRAFT-DO NOT QUOTE OR CITE

-------
 1     Both predicted zero risk for cumulative silica exposures of 0.6 mg/m3 • years (equivalent to
 2     a 20-year workplace exposure to an average concentration of 30 /zg/m3).  At higher
 3     exposures, excess risk was observed in these workers (e.g., 2% risk at 1.6 mg/m3 • years).
 4     These  effective occupational exposures  are greater and the particle sizes smaller than those
 5     likely to be experienced by the public; however, the public would be expected to  include
 6     susceptible subpopulations.  Information gaps still exist for both the exposure-response
 7     relationship (especially in potentially susceptible subgroups) for levels of exposure within the
 8     general population.
 9
10     Asbestos
11          The mechanisms underlying the development of asbestos-induced pulmonary fibrosis in
12     rats is complex.  While the acute response to asbestos results  in pulmonary inflammation and
13     cell proliferation, the pattern of fibrosis following chronic exposures becomes more complex.
14     It is likely that the retention of inhaled  fibers and  consequent  accumulation of interstitial
15     fibers  concomitant with prolonged inflammation will contribute to the development of a
16     diffuse and progressive pattern of pulmonary fibrosis.  The pathogenesis of asbestos-related
17     lung tumors clearly is a complex process and requires further investigation.
18
19     Ultrafine Particles
20          Certain freshly-generated ultrafine particles when inhaled as singlets at very low mass
21     concentrations (10 to 50 /xg/m3)  can be highly toxic to the lung.  Mechanisms  responsible for
22     this high toxicity could include (1) high pulmonary deposition efficiencies of these particles
23     (2) the large numbers per unit mass of these particles, (3) their increased surface  area
24     available for reaction, and (4) the presence of radicals on the particle surface,  depending on
25     the process of generation of the particles.  Results of studies with ultrafine model particles
26     indicate that particle  number may be of more import as a dose parameter, than just particle
27     mass.
        April 1995                                1-30       DRAFT-DO NOT QUOTE OR CITE

-------
  1      1.12  EPIDEMIOLOGY STUDIES OF HEALTH EFFECTS ASSOCIATED
  2            WITH EXPOSURE TO AIRBORNE PARTICLES/ACID AEROSOLS
  3          Chapter twelve assesses epidemiological evidence relating human health effects to
  4      exposure to airborne particles, which constitutes some of the most crucial information
  5      assessed in this latest PM criteria review.  Much new information has appeared since EPA's
  6      publication of the 1982 document on Air Quality Criteria for Paniculate Matter, and sulfur
  7      oxides (PM/SOX), its second Addendum (1986), and a later Acid Aerosol Issue Paper (1989).
  8      A rapidly  growing body of epidemiologic literature examines relationships between PM
  9      concentrations and human health effects, ranging from respiratory function changes and
10      symptoms to exacerbation of respiratory disease and excess mortality associated with
11      premature death.
12          The  time-series mortality studies reviewed  in this and past criteria documents provide
13      evidence that PM-containing ambient air pollution can cause increases in daily human
14      morbidity  and mortality. The newly available epidemiology studies provide indications that
15      very small increases in relative risk for such effects are associated with ambient air pollutant
16      mixtures containing  low or moderate concentrations of PM, as indexed by a variety of
17      monitoring methods (e.g., black smoke, TSP, COH, PM10,PM25). This includes emerging
18      new evidence indicating likely associations of health effects with PM across a wide range of
19      routine ambient concentration levels seen in the United States and other countries, including
20      levels that extend below present U.S. PM air quality standards.
21          The  1982 EPA criteria document earlier concluded that the most clearly defined effects
22      on mortality arising  from exposure to PM were sudden increases in the number of deaths
23      occurring, on a day-to-day basis, during episodes of high pollution, as  occurred in the Meuse
24      Valley in 1930, in Donora in 1948, and in London in 1952.  During the December, 1952
25      London episode 3,000 to 4,000 excess deaths were attributable  to air pollution, with the
26      greatest increase  in the death rate most dramatic for those > 45 years old and occurring
27      most notably in those with chronic lung disease and heart disease due to pollution-induced
28      cardio-respiratory problems.  Other episodes with associated notable increases in mortality
29      occurred in London  during various winters from 1948 to  1962.  Collectively, studies of these
30      and other early episodes left little doubt that airborne particles contribute to mortality
       April 1995                              1-31      DRAFT-DO NOT QUOTE OR CITE

-------
 1     associated with very high concentrations of urban aerosol mixes dominated by combustion
 2     products (e.g., from burning coal) and/or their transformation products (e.g., H2SO4).
 3           Besides evaluating mortality associated with major episodes, the 1982  criteria
 4     document also focused on studies of more moderate day-to-day variations in  mortality within
 5     large cities in relation to PM pollution.  Various methodological problems were identified for
 6     most of the studies, precluding quantitative conclusions regarding  exposure-response
 7     relationships of importance for deriving air quality standards. Among the main problems
 8     were inadequate measurement or control for potentially confounding variables and inadequate
 9     quantification of exposure to airborne particles and other associated pollutants (e.g., sulfates
10     or acid aerosols).  Despite such problems, the 1982 document concluded that the then
11     available  studies collectively indicated that mortality was clearly and substantially increased
12     when airborne particle 24-h concentrations exceeded 1,000 /-ig/m3 (as measured by the black
13     smoke, or BS, method) in conjunction with sulfur dioxide (SO2) elevations in excess  of 1,000
14     Mg/m3 (with the elderly or others with severe preexisting cardiovascular or respiratory
15     disease mainly being affected).
16           The 1986 addendum to the 1982 criteria document later considered several additional,
17     then-new analyses of acute PM exposure mortality in London during the 1958-1959 through
18     1971-1972 winter periods.  After reviewing the new data analyses, and taking into account
19     the previously reviewed London results and the above noted methodological  considerations,
20     the following conclusions were drawn:
21
22           (1)    Markedly increased mortality occurred, mainly among the elderly and chronically
23                 ill, in association with BS and SO2 concentrations above 1,000 /xg/m3, especially
24                 during episodes when such pollutant elevations occurred for several consecutive
25                 days;
26
27           (2)    During such episodes, coincident high humidity  or fog was also  likely important,
28                 possibly by providing conditions leading to formation of sulfuric acid (H2SO4) or
29                 other  acidic aerosols;
30
31           (3)    Increased risk of mortality is associated with exposure to BS and SC^ levels in
32                 the range of 500 to 1,000 /xg/m3, for SO2 most clearly at concentrations in
33                 excess of =700 /ig/m3; and
34
35           (4)    Convincing evidence indicates that relatively small, but statistically significant,
36                 increases in the risk of mortality exist at BS (but not  SO2) levels below 500
37                 Mg/m3. with no indications of any specific threshold level having been

       April 1995                                 1-32      DRAFT-DO NOT QUOTE OR CITE

-------
  1                 demonstrated at lower concentrations of BS (e.g., at  < 150 /*g/m3). However,
  2                 precise quantitative specification of the lower PM levels associated with mortality
  3                 is not possible, nor can one rule out potential contributions of other possible
  4                 confounding variables at these low PM levels.
  5
  6      In setting the current U.S. PM standards, the BS levels noted above were taken as indexing
  7      particles roughly in the same size range as inhalable particles reaching tracheobronchial or
  8      alveolar regions of the respiratory tract; and, taking into account other evidence of morbidity
  9      effects (e.g., worsening of chronic  bronchitis symptoms),  the U.S. 24 h primary NAAQS
10      was set as 150 jig/m3 PM10.
11           The decade or so since the 1986 EPA Addendum has seen the publication of numerous
12      new time series analyses of associations between human mortality or morbidity and acute
13      exposures to PM concentrations at or below the lower end of the range indexed by the above
14      studies of London mortality or the level of the current U.S. 24-h standard.  Some utilized
15      TSP or other measures (e.g., COH, BS, etc.) as an indices of PM exposure, but during the
16      last few  years, the analyses have mainly focused on PM10 as a measure of PM.
17
18      Short-Term PM Exposure Mortality Studies
19           Based on the new time-series  analyses, numerous investigators have reported very
20      small, but statistically significant associations between increased relative risk for mortality
21      and various indices of PM (e.g., BS,  COH, TSP, PM10, PM2.5, etc.) for many different
22      cities in  the United States and in other countries, as well.  The elderly (>65 yr old),
23      particularly those with preexisting cardiopulmonary disease, are found to have distinctly
24      higher risks than younger age groups.  The small relative risk estimates for  PM are generally
25      reduced  when other likely important (potentially confounding) factors are also controlled for
26      in the models,  but the PM association still usually remains statistically significant, although
27      typically accounting for much less of the variance in mortality than temperature or
28      combinations of variables used to index contributions of weather-related mortality.  Thus,
29      qualitatively, the newly emerging database appears  to provide indications that polluted
30      atmospheres containing relatively low concentrations of particles may contribute (along with
31      other factors) to a very small increase  in relative risk for human mortality, especially in the
32      elderly with preexisting cardiopulmonary diseases.


        April 1995                                1_33       DRAFT-DO NOT QUOTE OR CITE

-------
 1           It is important to note that differences in opinion exist within the scientific community
 2      with regard to:  (a) how adequately other likely important confounding factors (including
 3      weather and copollutants) and/or other seasonal factors were controlled for across the various
 4      new analyses; and (b) interpretation of reported associations between increased relative risk
 5      estimates and indices of ambient PM  concentrations.  For example, introduction of one or
 6      more other commonly-present ambient air pollutants (e.g., SO2, O3, CO, NOX) into models
 7      of PM effects generally reduces the estimated PM effect, often by as much as 50% and, at
 8      times, to statistically non-significant levels.  In a few studies, however, the size of the PM
 9      effect remained essentially the same or increased slightly with other copollutants in the
10      model.   Similarly, analyses of PM-mortality effects by  season (winter, spring, summer,  fall),
11      as in a  few studies,  so far have  yielded varying patterns of PM-mortality effects being
12      significant in one or another season(s) but not all, with specific effective seasons differing
13      from one locale to another.  The copollutant and seasonality  analyses results, in particular,
14      have  led to considerable debate  in the scientific community, typified on the one hand by (a)
15      skepticism about the size  and the "realness" of reported low-level PM effects and, on the
16      other hand, (b) countervailing views asserting that the effect of PM (or any other  weakly
17      contributing factor) on mortality can be made to "disappear" by overspecification of
18      applicable models (i.e., by introduction of sufficient other, possibly extraneous, variables
19      into the models or by more detailed breakdowns of data (e.g., by season) that may reduce
20      the power to detect a PM effect).
21           No clear resolution of this debate or "consensus"  opinion in the scientific community
22      has yet crystallized, but some agreement appears to be  emerging that the results for models
23      containing only PM and no other copollutants may provide upper bound estimates for effects
24      of ambient particle-containing mixes of pollutants, whereas results  derived from analyses
25      including other copollutants and extensive controls for weather,  seasonality,  and/or other
26      likely important contributing factors should be viewed as lower-bound estimates of PM
27      effects  (which may be 50% or more lower than the upper bound or even include zero).  Key
28      points regarding derivation of quantitative estimates of PM-related mortality and morbidity
29      effects, taking into account the types  of uncertainties  and scientific debate just noted  are
30      summarized below.  Major emphasis is placed first on  results derived from studies of PM10
        April  1995                                1-34       DRAFT-DO NOT QUOTE OR CITE

-------
  1      as the PM indicator of most interest in relation to the current U.S. PM standards and, then,
  2      additional key quantitative points for PM2 5 and acid aerosols are summarized.
  3
  4      PM10 Relative Risk Analyses
  5           This section discusses key findings from studies conducted since the 1986 PM criteria
  6      document addendum that have employed PM10 in their analyses of the human mortality
  7      effects of acute exposures to PM, as discussed in more detail in Chapter 12.  Some studies
  8      considered daily mortality in the entire population (i.e., all ages) and some by cause; some
  9      also considered subpopulations (e.g, the elderly).
10           Two earlier published summaries of the PM literature converted all results to a
11      PM10-equivalence basis and provided quantitative intercomparisons and after such summaries
12      used TSP as the reference PM metric.  The results from such summaries suggest about a 1
13      percent change in acute total mortality for a  10 /xg/m3 change in PM10, but the estimates
14      range from 0.3 to 1.6% (i.e., a factor of 5). While most of the 95% confidence intervals
15      (CIs) of these estimates overlap, CIs of the highest and lowest estimates do not overlap,
16      indicating significant differences between these estimates. Note that the effects indicated for
17      a 10 /ig/m3 PM10 change cannot be reliably converted to other PM increments (e.g., 50 or
18      100 /xg/m3 PM10), as differences in model specification (e.g., linear versus  log models) will
19      cause them to differ in their conversions to other particle concentration ranges.  The reasons
20      for the approximately  five-fold effect estimate difference noted among studies are not
21      obvious,  but one factor appears to be the  PM exposure averaging time, as estimates using
22      multiple day PM10 averages are all 1 % or higher.  This is not unexpected, given that any
23      lagged effects from prior days of PM10 exposure will be added to the effects estimate when a
24      multi-day average is employed, increasing the estimated effect on a per /xg/m3 basis.
25           It is also important to note that other air pollutants were generally not addressed in
26      deriving the coefficients reported by the above summaries.   Differences among coefficients
27      are to be expected, given that the composition (and, potentially toxicity) of the PM, as well
28      as the demographic characteristics in each city, can be expected to  differ.  Moreover, the
29      conversions from other PM metrics to PM10  necessarily introduce much additional
30      uncertainty.  However, though not all of these results may therefore be the most appropriate
31      available  for quantifying a PM10 effect, they  do consistently indicate that there is an

        April 1995                                 1_35      DRAFT-DO NOT QUOTE OR CITE

-------
  1      association between acute exposure to ambient air mixes containing PM and increased daily
  2      mortality.  Moreover, the by-cause results also reported in the summaries indicate that PM
  3      effect estimates are greater  for respiratory causes, which lends support to the biological
  4      plausibility of the noted PM associations.
  5           In an effort to more clearly quantify daily PM10-total acute mortality associations,
  6      Table 1-5 summarizes total mortality relative risk (RR) estimates reported for a 50 /xg/m3
  7      increase in PM10 estimated  from nine studies reviewed in Chapter 12, which employed PM10
  8      data in their analysis of total mortality data (or which had on-site PM10 reference data to
  9      convert other PM metrics with more certainty).  The studies listed were selected for this
10      analysis mainly because they can most readily  be intercompared and provide direct data
11      related to PM10 levels.  The RR's calculated were based upon a 50 /ig/m3  increase above the
12      mean PM10 24-h concentration, which is approximately the order of magnitude of the typical
13      difference between the mean and maximum in  these cities evaluated and is fairly
14      representative of the mean/max differences seen for most of the U.S. cties in Tables 1-2 and
15      1-3.  This is noted because  in non-linear models as often employed in the studies discussed in
16      Chapter 12, the RR estimate associated with a  given /ng/m3 PM10 increase  will vary
17      depending upon the baseline concentration to which it is added.
18           From results presented in Chapter 12, it is  apparent that these studies  generally have
19      yielded at least marginally significant PM10 coefficients, but the resultant excess risk
20      estimates differ by a factor  of five across these studies (from 1.5% to 8.5% per 50 /ig/m3).
21      The mean and maximum PM10 concentration data are noted for each  study. If the PM10
22      coefficient decreased  as the mean level of PM10 decreased, then confounding as a function of
23      varying PM level would be suggested.  However, the data presented  indicate that the
24      variability in coefficients is  not a function of PM10 level, as sites with high or low PM10
25      concentrations can report either high  or low RR's.  In Chapter  12, the statistical
26      methodology characteristics of each  study were concisely summarized, in order to determine
27      if any factors are important to help explain the variability observed from study to study in the
28      PMio RR estimate.  As noted earlier, the RR estimate for acute mortality  associated with
29      short-term exposure to PM10 is usually larger when other pollutants are not simultaneously
30      considered in the model. Those studies  which considered PM10 both alone and with other
31      pollutants in the model often yielded smaller, and usually more marginally significant, PM10

        April 1995                               1_36      DRAFT-DO  NOT QUOTE OR  CITE

-------
TABLE 1-5. COMPARISON OF RELATIVE RISK (RR) ESTIMATES FOR TOTAL MORTALITY FROM 50 jtg/m3
 CHANGE IN PM10, USING STUDIES WHERE PM10 WAS MEASURED OR WAS CALIBRATED FOR THE SITE
Co

H
O'
O

1
H
O
e:
H
tn
O
Q
H
W

Study

Utah Valley, UT




St. Louis, MO

Kingston, TN

Birmingham, AL
Athens, Greece

Toronto, ON Canada
Los Angeles, CA

Chicago, IL
Santiago, Chile





Chicago, IL

'Calculated on a basis of 50





Reference PM10 (/ig/m3)
Mean Maximum
Pope et al.(1992) 47 297




Dockery et al. (1992) 28 97

Dockery et al. (1992) 30 67

Schwartz (1993) 48 163
Touloumi et al. (1994) 78 306

Ozkaynak et al. (1994) 40 96
Kinney et al. (1995) 58 177

Ito et al. (1995) 38 128
Ostro et al. (1995a) 115 367





Styer et al. (1995) 37 365

/ig/m3 increase from 50 to 100 jug/m3.




Other Pollutants
In Model

None
None, Winter
None, Summer
Max O3, Summer
Avg O3, Summer
None
03
None
03
None
None
SO2, CO
None
None
03, CO
O3, CO
None
None
None, Poisson
SO2, Poisson
NO2, Poisson
O3, Poisson
None







Lag Times, d

< 4 d
< 4d
< 4 d
< 4 d
< 4 d
< 3 d
< 3 d
< 3d
< 3d
< 3 d
1 d
1 d
Od
1 d
1 d
< 3 d
1 d
< 4 d
1 d
1 d
1 d
1 d
3 d






RRper
50 /ig/m3

1.08
1.085
1.11
1.19
1.14
1.08
1.06
1.085
1.09
1.05
1.034
1.015
1.025
1.025
1.017
1.025
1.04
1.07
1.0221
1.0261
1.0431
1.0261
1.04






95 Percent
Confidence Interval

(1.05, 1.11)
(1.03, 1.14)
(0.92, 1.35)
(0.96, 1.47)
(0.92, 1.41)
(1.005, 1.15)
(0.98, 1.15)
(0.94, 1.25)
(0.94, 1.26)
(1.01, 1.10)
(1.025, 1.044)
(1.00, 1.03)
(1.015, 1.034)
(1.00, 1.055)
(0.99, 1.036)
(1.005, 1.05)
(1.005, 1.06)
(1.04, 1.10)
(1.003, 1.042)
(1.005, 1.047)
(1.020, 1.066)
(1.005, 1.047)
(1.00, 1.08)







-------
 1      relative risks when other pollutants were also considered.  This ranges from roughly a 20 to
 2      50 percent reduction in the estimate of excess risk associated with PM10 (e.g., in Athens,
 3      Greece, the PM10 RR declines from 1.07 to 1.03 per 100 jug/m3 when other pollutants are
 4      considered).  Such a reduction is to be expected when co-linear variables are added.
 5           Older studies using BS or TSP often found high correlations between SO2 and the PM
 6      indicator which reduced the apparent PM effect and attenuated its statistical significance.
 7      However, studies using a variety of PM  indicators at cities were SO2 levels were so low as
 8      to have little likelihood of SO2 being a significant confounder of a PM effect found
 9      quantitatively similar significant PM effects.  While there is some possibility that
10      summertime  PM effects may be partially confounded with those of other pollutants (e.g., O3)
11      derived from motor vehicle fuel combustion or transformation products, winter effects of PM
12      are clearly detectable when O3 levels are much lower. If PM effects on mortality were so
13      completely confounded with those co-pollutants so as to be undetectable, then one would
14      need to invoke many different confounders in different studies of communities.   While this
15      explanation is not impossible, it appears  highly unlikely, but cannot be precluded altogether
16      since PM may  derive from different sources in these studies, have varying size  and chemical
17      composition  from one  locale to another,  and therefore may have different  characteristics that
18      affect health  outcomes such as mortality.
19           Another factor clearly affecting the PM10 RR estimates is the PM10 averaging period.
20      Most of the studies which utilized multi-day averages of PMjo in their regressions (i.e., for
21      Utah Valley; St. Louis; eastern TN; Santiago; Chicago; and Birmingham) yielded higher RR
22      estimate studies. However, the  increase indicated for these studies is not proportional to the
23      averaging time. Indeed, in sub-analyses for Utah Valley data, the PM10 mortality risk is
24      indicated to be roughly doubled by using a five day average versus a single day
25      concentration,  and sub-analyses  for Santiago also indicate approximately a doubling in the
26      PMio RR when a  3 day average is considered (i.e., from RR = 1.04 for a single day PM10
27      value to RR  =  1.07 for a 3d average PM10 value).  This may be due to the fact that, since
28      autocorrelation exists in the PM10 concentrations  from day to day, the single day
29      concentration is "picking  up"  some of the effect of multi-day pollution episodes, even though
30      they are not  explicitly  modeled.  These results suggest that a multi-day rather than a
        April 1995                               1-38       DRAFT-DO NOT QUOTE OR CITE

-------
 1      single-day average PM10 concentration may provide a more relevant index to gauge the
 2      effects of short-term PM exposures over several consecutive days.
 3           Table 1-5 shows that total acute mortality relative risk estimate associated with a 50
 4      Mg/m3 increase in the one-day 24-h average PM10 ranged from 1.015 to 1.085, depending
 5      upon the site (i.e., the PM10 composition and population demographics) and whether PM10 is
 6      modeled as the sole index of air pollution.  Relative Risk estimates with PM10 as the only
 7      pollutant index in the model range from RR = 1.025 to 1.085, while the PM10 RR with
 8      multiple pollutants in  the model range from 1.015 to 1.025.  As noted earlier, the former
 9      range might be viewed as approximating an upper bound of the best estimate, as  any
10      mortality effects  of co-varying pollutants are likely to be "picked up" by the PM10 index.
11      On the other hand, the latter multiple pollutant model range might be viewed as
12      approximating a  lower bound of the best estimate, as the inclusion of highly correlated
13      covariates may weaken the PM10 estimate.  Overall, consistently  positive PM-mortality
14      associations are seen throughout these analyses, even with the use of various  modeling
15      approaches and after controlling for major confounders such as season, weather, and
16      co-pollutants.  The 24-h 50 jig/m3 PM10 total mortality effect estimate most typically falls in
17      approximately the RR = 1.025  to 1.05  range (representing an expected 2.5 to 5.0% increase
18      in risk of death over daily background mortality rates for which a 50 /xg/m3 increment in
19      ambient PM10 concentration could be a  contributing factor).
20           It is logical to assume that the  bulk of the total mortality effects suggested by these
21      studies are among the elderly.  During the historic London, 1952 pollution episode the
22      greatest increase  in the mortality rate was among older citizens and those having  respiratory
23      diseases. An analysis of mortality in Philadelphia, PA  during 1973 through 1980 comparing
24      mortality during  the 5% highest versus the 5% lowest TSP days also found the greatest
25      increase in risk of death to be among those aged 65  to 74 and those > 74  year of age
26      (mortality risk ratios = 1.09 and 1.12, respectively, between high and low TSP days).  Also,
27      in a time series analyses of Philadelphia daily mortality during this period, the TSP-mortality
28      coefficient was significantly higher (6 = 0.000910 ± 0.000161) for persons  > 65 yrs old
29      than for the younger population (6 = 0.000271 + 0.000206). These coefficients indicate an
30      effect size for the elderly roughly three  times that for the younger population (10% versus
31      3%, respectively, for a 100 pig/m3 increase in TSP).  Also,  two other recent PM10 analyses

        April 1995                                 1_39       DRAFT-DO NOT QUOTE OR CITE

-------
  1      (one for Sao Paulo, Brazil and the other for Santiago, Chile) directly considered the question
  2      of PM10-mortality associations among the elderly population (> 65 years of age).  The first
  3      of these two analyses evaluated PM10-mortality associations for the elderly in Sao Paulo
  4      during 1990 through 1991 in Sao Paulo. The reported PM10 relative risk (RR =  1.13 for a
  5      100 jitg/m3 increase) is higher than noted above for total mortality analyses addressing
  6      multiple pollutants (100 /xg/m3 RR = 1.03 to 1.05), supporting past observations  that the
  7      elderly represent a population especially sensitive to the health effects of air pollution.  The
  8      second study in Santiago, Chile found a 24-h PM10 100 peg/in3 RR estimate of 1.08, for the
  9      overall population but the RR estimates for a 100 /ig/m3 increase in PM10 rose to  an RR =
10      1.11 for the elderly (aged 65+) using the same model specification.  Thus, these directly
11      comparable estimates (i.e., using the same model specification and population) suggest that
12      the elderly experience roughly a 40 percent higher excess  risk from exposure to PM air
13      pollution than the overall population.
14           Overall, considering the historical pollution episode evidence and the results  of recent
15      PM10-mortality analyses evaluating elderly populations, it  seems evident that elderly adults
16      represent a population especially at risk for mortality implications  of acute exposure to air
17      pollution,  including PM.
18           Relatively few studies have directly examined the PM-mortality  association in children.
19      It is difficult, given the limited and somewhat conflicting results  available at this time, to
20      ascribe any such association to low-level PM pollution in particular.  This is an area where
21      further research is clearly needed to broaden the base upon which  to assess the potential for
22      PM to increase mortality among children.
23           In studies alluded to above and in  others disused in more detail in Chapter 12, a
24      consistent trend was for acute PM exposure effect estimates to be higher for the respiratory
25      mortality category than for total mortality from all (non-accident) causes.  This lends support
26      to the biological plausibility  of a PM air pollution effect, as the breathing of toxic particles
27      would be expected to most directly affect the respiratory tract.  Of particular interest is
28      comparison of relative risk values from  those studies that made most direct and appropriate
29      comparisons. In a Santa Clara study, the PM-respiratory mortality RR was 4.3 times as
30      large as for deaths as a whole (i.e., 3.5/0.8); for Philadelphia, the PM (TSP)-respiratory
31      mortality RR was 2.7 times as large as for total mortality  (i.e., 3.3/1.2); for Utah Valley,

        April  1995                                1-40      DRAFT-DO NOT QUOTE OR CITE

-------
 1     the PM10-respiratory mortality RR was 2.5 times as large as for deaths as a whole (i.e.,
 2     3.7/1.5); for Birmingham, AL, the respiratory mortality RR of PM10 was 1.5 times as large
 3     as for deaths as a whole (i.e.,  1.5/1.0); and for Santiago, Chile the reported excess
 4     respiratory mortality RR of PM10 was 1.8 times as large as for deaths as a whole (comparing
 5     1.15/1.08 RR per 100 /*g/m3). Thus, in these studies, the PM RR for respiratory diseases is
 6     indicated to range from 50 to over 400%  higher for respiratory disease categories than for all
 7     causes of death, indicating that increases in respiratory deaths are a major contributor to the
 8     overall PM-mortality associations noted previously.  Moreover,  since evidence suggests that
 9     an acute pollution episode is most likely be inducing its primary effects by stressing already
10     compromised individuals  (rather  than, for example,  inducing chronic respiratory disease from
11     a single air pollution exposure episode), the above results indicate that persons with
12     pre-existing respiratory disease represent a population especially at risk to the mortality
13     implications of acute exposures to air pollution, including PM.
14           In overall summary, the time-series  mortality studies reviewed in this and the previous
15     1982 and 1986 PM  criteria assessments provide reasonably strong evidence that increases in
16     daily human mortality are associated with short-term exposures to air pollution mixes
17     containing elevated  PM levels. Recent studies provide indications that small increases in
18     such risk occur in association with air pollution indexed by moderate increases of 24-h PM10
19     (~50 Mg/m3) above routine ambient levels averaging around 50 to 100 /-tg/m3.  Overall, the-
20     PM10 relative risk estimates derived from the most recent PM10 total mortality studies
21     suggest an acute exposure effect  on  the order of RR =  1.025 to  1.05 in the general
22     population for increases in ambient air pollution indexed by a 24-h average 50 /*g/m3 PM10
23     increment, with higher (30-40%) relative  risks indicated for the elderly sub-population and
24     for those with pre-existing respiratory conditions.
25
26     Fine Particles/Acid Aerosols Relative Risks
27           As noted earlier  and in both Chapters 11 and 12, some epidemiologic and experimental
28     toxicology data point toward fine particles as  a class or certain constituents (e.g.,  acidic
29     aerosols) as possibly being key contributors to observed PM-mortality and/or PM-morbidity
30     associations.  Only a few epidemiologic studies provide direct comparisons between various
31     PM indices, including fine particle and acidity measurements.

       April 1995                                 1-41       DRAFT-DO NOT QUOTE OR CITE

-------
 1           One such study investigated relationships between multiple air pollutants (including
 2      PM2.s) and total daily mortality during a one year period in St. Louis,  MO and
 3      Kingston/Harrirman, TN and surrounding counties.  In Poisson regressions controlling for
 4      weather and season, previous day's PM10 was the only significant predictor of daily mortality
 5      (B  = 0.00175 ± 0.00067), but the association dropped off at 3 days (fl = 0.00042 +
 6      0.00063).  Size-fractionated PM data were examined to determine whether this association
 7      could be attributed to either the fine (PM2 5, aerodynamic diameter d&  < 2.5 ju.m) or the
 8      coarse (2.5 /mi  < Ja < 10 ^im) component of the PM10 mass.  The fine fraction (PM2 5)
 9      was positively associated with mortality (6 = 0.00171  + 0.00096, P = 0.075).  Coarse
10      particles were also positively associated (6 = 0.00247  + 0.00129, P = 0.056).  Neither fine
11      nor coarse particles showed a stronger association than the other when  considered
12      simultaneously.
13           Both daily SO4 and H+ concentrations were significantly correlated with PM10 (Pearson
14      correlations 0.52 and 0.76, respectively).  Sulfate  (SO42~) as measured  by the sulfur fraction
15      of PM10 (/? = 0.00608  + 0.00577) and H+ (6 =  0.00086 + 0.00118) were positively, but
16      not significantly, associated with daily mortality.  Among other PM elements measured,
17      those correlated with PM10 concentrations were also associated with mortality.  In particular,
18      aluminum, calcium, chromium,  iron, and  silica all had correlations with PM10 of 0.5 or
19      higher and had positive associations with mortality.  Neither SO2, NO2, nor O3 was
20      significantly associated  (P > 0.30) with total mortality.
21
22      Long-Term PM10/PM2 5 Exposure Mortality  Studies
23      Population-Based Cross Sectional Mortality Studies
24           Ecological cross-sectional studies employing averages across various geopolitical units
25      (cities,  SMS As, etc.) present data that examine relationships between community-wide PM
26      levels and mortality.  Such community-based studies seek to define (average) community
27      characteristics associated with overall average health status-in this case, annual mortality rate.
28           One study analyzed 1980  total mortality in 98 SMSAs, using data on PM15 and PM2 5
29      from the EPA inhalable particle (IP) monitoring network for 38 of these locations, ranked the
30      importance  of the pollutants by relative statistical significance in separate regressions, and
31      concluded that the results were "suggestive" of an effect of particles on mortality decreasing

        April 1995                                1-42      DRAFT-DO NOT QUOTE OR CITE

-------
  1      with particle size;  although in the basic model only SO42" was statistically significant.  In
  2      some other models tested, PM2 5 was also significant, and PM15 nearly so.   However, if the
  3      effects are judged by elasticities rather than significance levels, SO42"'  PM2 5, and PM15
  4      would be judged as equivalent, with TSP ranking somewhat lower.  Also, based on source
  5      apportionment techniques, particles from coal combustion and from the metals industry
  6      appeared to be the most important.  The specific coefficients and significance levels obtained
  7      for TSP may be the result of the particular TSP data used, being based on a  single
  8      monitoring station in each SMSA and thus not clearly fully representative of  population
  9      exposures.  Thus, alternative interpretations of these findings are certainly possible.  In
 10      addition, because  smoking, diet, and other socioeconomic or lifestyle variables were not
 11      considered  in the regression model, the pollution coefficients may have been  biased.  Finally,
 12      the study did not specifically address the question of acute vs. chronic  responses by exploring
 13      lagged pollution variables.
 14           Data from up to 149 metropolitan areas  (mostly SMSAs) were  analyzed in another
 15      study of relationships between  community air  pollution and "excess"  mortality due to various
 16      causes for the year 1980.  Several socioeconomic models were used  in cross- section multiple
 17      regression analyses to account  for non- pollution effects. Two different sources of
 18      (measured) air quality data were used: data from the EPA AIRS database (TSP, SO42, Mn,
 19      and ozone) and data from the inhalable paniculate (IP; PM15) network; the latter data
 20      (PM15,PM2.5 and SO4= from the IP filters) were only available for 63  locations.  All PM
 21      data were averaged across all monitoring stations available for each SMSA, with TSP data
 22      restricted to 1980  and based on an average of about 10 sites per SMSA.  Using these models,
 23      statistically significant associations were found between TSP and mortality due to  non-
 24      external causes with the log-linear models evaluated, but not with a linear model.  Sulfates,
 25      manganese, inhalable particles  (PM15), and fine particles (PM2 5) were  not significantly (P <
 26      0.05) associated with mortality with any  of the parsimonious models, although PM2 5 and
27      manganese were close with linear models (p=0.07) and significance  may have been affected
28      by the use of smaller data sets.  This study found PM2 5 to be the  "strongest" PM index with
29      linear models, but TSP with log-linear models.  This  study support previous findings of
30      associations between TSP and premature mortality.
31

        April 1995                                1_43      DRAFT-DO  NOT QUOTE OR CITE

-------
 1     Prospective Mortality Studies
 2          Prospective studies consider data on the relative survival rates of individuals, as
 3     affected by age, sex, race, smoking habits, and certain other individual risk factors.  This
 4     type of analysis has a substantial advantage over the above population-based studies,  because
 5     identification of actual decedents allows stratification according to important risk factors such
 6     as smoking.  However, since none of the prospective cohort studies had data on personal
 7     exposures to air pollution, these studies are also considered to be "ecological."
 8          Several newer prospective studies are of most interest.  One such study followed
 9     approximately 6,000 white,  non-hispanic, nonsmoking, long-term California residents for 6
10     to 10 years, beginning in 1976. The study was designed to test the use of cumulative
11     exposure data as an explanatory factor for disease incidence and chronic effects.  Only TSP
12     and ozone data were used to index pollution exposures;  with reliance  on oxidant
13     concentrations  in the early part of the monitoring record.  In a follow-up analysis, SO42",
14     PM10 (estimated from site-specific regressions on TSP), PM2 5 (estimated from visibility),
15     and visibility per se (extinction coefficient) were used to index PM exposure.  No significant
16     associations with nonexternal mortality were reported, and only high levels of TSP or PM10
17     were associated with symptoms of asthma, chronic bronchitis or emphysema.  The finding of
18     no association between long-term cumulative exposure to TSP or O3 and all natural-cause
19     mortality may be interpreted as showing the absence of chronic responses after 10 years but
20     not necessarily the absence  of (integrated) acute responses, since coincident air pollution
21     exposures were not considered.
22          Another prospective study analyzed survival probabilities among 8,111 adults first
23     recruited in the mid-1970s in six cities in the eastern portion of the United States. The cities
24     are: Portage, WI, Topeka, KS; St. Louis, MO; Steubenville, OH, Watertown,  MA,  and
25     Kingston-Harriman,  TN, two small towns southwest of Knoxville.  These locations thus
26     comprise a transect across Northeastern and Northcentral United States, from suburban
27     Boston, through Appalachia, and into the upper Midwest.  The adults were white, aged 25 to
28     74 at enrollment, and the final cohorts numbered 1,400 to 1,800 persons in each  city.
29     Follow-up periods ranged from 14 to 16 years, during which from 13 to 22% of the
30     enrollees died.  Of the 1,430 death certificates, 98% were located, including those for
31     persons who had moved away  and died elsewhere.  The bulk of the analysis was based on

       April 1995                                 1-44      DRAFT-DO  NOT QUOTE  OR CITE

-------
  1     all-cause mortality and included individual characteristics of the members (and thus of the
  2     decedents), e.g. smoking habits, an index of occupational exposure, body mass index, and
  3     completion of a high school education.  Air pollution effects were evaluated in two ways: by
  4     evaluating the relative risks of residence in each city relative to Portage (the city with the
  5     lowest pollution levels for most indices), and by including the community-average air quality
  6     levels directly in the models.
  7          Based on statewide mortality data, substantial differences in survival rates would be
  8     expected across this transect of the Northeastern U.S. and were in fact observed.  The long-
  9     term average mortality rate  in Steubenville was 16.2 deaths per 1,000 person-years; in
 10     Topeka, it was 9.7, yielding a  67% variation in the range of annual average (crude) relative
 11     risk across the six cities.  After individual adjustment for age, smoking status, education, and
 12     body-mass index, the range in average relative risk was reduced to 26%.  The authors
 13     reported that "mortality was more  strongly associated with the levels of fine, inhalable, and
 14     sulfate particles" than with the  other pollutants (e.g., NOX, SO2, TSP), which they attributed
 15     primarily  to factors of particle size.  For those three PM indices, relative risk estimates and
 16     confidence limits based on the differences between air quality in Steubenville and in Portage
 17     were calculated.  Only small differences were found between many pollutants, including S02
 18     and NO2,  owing in part to the strong collinearity present.  Neither mortality associations with
 19     TSP nor with coarse particle fractions created by subtracting PM15 from TSP or PM2 5 from
 20     PM15 were significant, suggesting that particles >  15 um may be less important.  This
 21     outcome may reflect in part greater spatial variability within the communities for these
 22     measures.   The non-sulfate portion of PM2 5 had the tightest confidence limits (SO42~ was
 23     multiplied by 1.2 before subtraction, assuming an average composition of NH4HSO4).
 24     However,  all of the differences in relative risks and their confidence limits could have
 25     occurred due to chance, given the availability of only 6 observations.  No relationship was
 26     found for  aerosol acidity  (H+), but only limited data were available.
 27          The authors of this study appear to have made the most of the available individual data
28     on some of the most important mortality risk factors. They were quite cautious in their
29     conclusions, stating that the results only suggest that fine-particulate air pollution "contributes
30     to excess mortality in certain U.S. cities."  There are several other important outcomes-
31

        April 1995                                !_45       DRAFT-DO NOT QUOTE OR CITE

-------
 1           •  None of the population subgroups examined appeared to be stastically significantly
 2              more sensitive to air pollution than any other, although smokers had notably higher
 3              relative risk estimates (with wide confidence intervals) than non-smokers.
 4
 5           •  The  implied regression coefficients are much larger (about an order of magnitude)
 6              than those found in either type of population-based study noted earlier.  This could
 7              be interpreted as evidence that the chronic effects of air pollution far exceed the
 8              acute effects, or that not all of the spatial confounding has been controlled.  Use of
 9              linear models for non-linear effects (body-mass index) and failure to control for
10              alcohol consumption, diet,  exercise and migration may have contributed to the
11              relatively large effects indicated  for air pollution.
12
13           •  If the measured responses to air  pollution truly  are due to chronic PM exposure it is
14              logical to expect that cumulative exposure would be the preferred metric.  However,
15              pollution levels 10 years before this study began were much higher  in Steubenville
16              and St. Louis, as indexed by TSP from routine monitoring networks.  For example,
17              annual average TSP in 1965 in Steubenville was about three times the value used to
18              index chronic exposure in the  study.  Estimates of previous levels of fine  particles
19              are more difficult,  but atmospheric visibility data suggest that previous levels may
20              have been higher in winter, but not necessarily  in summer.   These uncertainties
21              make it difficult to  accept quantitative regression results based solely on coincident
22              monitoring data.
23
24      Because it seems unlikely that any of the  above-noted shortcomings of this study could have

25      resulted in bias sufficient to  reduce the risk estimates to  levels less than those found in acute
26      mortality studies, the Six City study appears to provide support for the hypothesis that effects

27      indexed by results of long-term air pollution studies must also reflect the presence of acute
28      effects on mortality  as integrated over the long term. Or, it may also be concluded that
29      support has been shown for the existence  of chronic effects; these two possibilities  are not
30      mutually exclusive.  However, these conclusions must be qualified by the realization that not
31      all of the relevant socioeconomic factors may have been properly controlled  in this study.
32           In a very recently reported  1995 study, 7-year survival data (1982 to 1989) for about

33      550,000 adult volunteers obtained by the  American Cancer Society (ACS) were analyzed.

34      The Cox proportional hazards model was used to define individual risk factors for  age, sex,

35      race, smoking (including passive smoke exposure), occupational exposure, alcohol
36      consumption, education, and body-mass index.  The deaths, about 39,000 in all,  were

37      assigned  to geographic locations using the 3-digit zip codes listed at enrollment into the  ACS
38      study in 1982.  Relative risks were  then computed for 151 metropolitan areas defined by

39      these zip codes and  were compared to the corresponding air quality data (ca. 1980) derived

        April 1995                                1-46        DRAFT-DO NOT QUOTE OR CITE

-------
 1      from the EPA AIRS system for sulfates, as obtained from high-volume sampler filters for
 2      1980, and the Inhalable Particulate Network for fine particles (PM2 5), measured by
 3      dichotomous samplers during  1979-81.  Causes of death considered included:  all causes,
 4      cardiopulmonary causes, lung cancer, and  all other causes.
 5           The adjusted total mortality risk ratios for the ACS study (computed for the range of
 6      the pollution variables) were 1.15  (95% CL =  1.09 to 1.22) for sulfates and 1.17 (95% CL
 7      = 1.09 to 1.26) for PM2 5. When expressed as log-linear regression coefficients, these
 8      values were quite similar for both pollution measures: 0.0070 (0.0014) per /ig/m3 for SO42
 9      and 0.0064 (0.0015) for PM2 5, suggesting that particle chemistry may be relatively
10      unimportant as an independent risk factor (it is possible that the SO42" results have been
11      biased high by the presence of filter artifacts).  However, the pollution coefficients were
12      reduced by 10 to 15% when variables for climate extremes were^dded to the model.
13           This study took great care to control  for those potential confounding factors for which
14      data were  available.  Several different measures of active smoking were considered,  as was
15      the time exposed to passive smoke.  The occupational exposure variable was specific to any
16      of:  asbestos, chemicals/solvents,  coal or stone dusts, coal tar/pitch/asphalt, diesel exhaust,
17      or formaldehyde.  The education variable was an indicator for having less than a high-school
18      education.  However,  the possible influences of other air pollutants were not discussed,  and
19      other risk  factors not considered included income, employment status, dietary factors,
20      drinking water hardness  and physical activity levels, all of which have been shown to affect
21      longevity.  Another important caveat is that the ACS cohort is by no means a random sample
22      of the U.S. population; it is 94% white and better educated than the general public, with a
23      lower percentage of smokers than in the Six City Study. The (crude) death rate during  the
24      7.25 years of follow-up was just under  1% per year, which is about 20% lower than
25      expected for the white population  of the U.S. in 1985, at the average age reported for the
26      study cohort.  In contrast,  the corresponding rates for the Six-City study discussed above
27      tended to be higher than the U.S.  average.
28           The results of the long-term  prospective cohort studies are compared in Table 1-6.  The
29      results of the American Cancer Society  (ASC) prospective study were qualitatively consistent
30      with those of the Six City  study with regard to their findings for sulfates and fine particles;
31      but relative standard errors were smaller, as expected because of the substantially larger

        April 1995                               1-47       DRAFT-DO NOT QUOTE OR CITE

-------
                                           TABLE 1-6.  PROSPECTIVE COHORT MORTALITY  STUDIES

§
u\
Source
Abbey et
al. (1991)



Dockery et
al. (1993)
p. 1758







Health
Outcome
Total
mortality
from
disease

Total
mortality









Population
Calif. 7th
Day
Adventist


White
adult
volunteers
in 6 U.S.
cities3




Time
Period/ PM
No. Units Indicators
1977-82 24 h TSP
Defined by > 200
air
monitoring
sites
1974-91 PM15
PM2.5
SO4






PM
Mean
(Mg/m3)
102




29.9
18
7.6







PM Range/
(Std. Dev.)
25-175
(annual avg)



18-47
11-30
5-13






Sites
Per Total
City Deaths Model Type
NA 845 Cox
proportional
hazards


1 1429 Cox
proportional
hazards







PM Lag Other
Structure Pollutants Other Factors
10 yrs none age, sex, race,
smoking,
education,
airway disease

none none age, sex,
smoking,
education,
body mass,
occup.
exposure
hypertension4.
diabetes4
Relative
Risk1 at
S04 = 15,
PM15 = 50,
PM2 5 = 25
0.99 TSP1




1.42 PM15
1.31 PM25
1.46SO/






RR.
Confidence
Interval
(0.87-1.13)'




(1.16-2.01)
(1.11-1.68)
(1.16-2.16)








Elasticity
NS2




0.25
0.22
0.23





OO
 Popeetal.  Total     American  1982-89   PM25     18.2     9-34
 (1995)     mortality  Cancer    PM25 50
 Table 2              Society,    cities
                    adult      SO4 151   SO4       II5      4-24
                    volunteers  cities
                    in U.S.
                                                                                20,765 Cox
                                                                                       proportional
                                                                                       hazard
                                                                                38,963
age, sex, race,   1.17PM25   (1.09-1.26)  0.117
smoking,
education,
body mass,     1.10SO4    (1.06-1.16)  0.077
occup.
exposure,
alcohol
consumption,
passive
smoking,
climate"
O
o
o
H
O
O
3
o
o
»—<
a
'For l,000h/yr > 200pg/m3.
2NS = non significant, confidence limits not shown.
3
Portage, WI; Topeka, KS; Watertown, MA; Harrisman-Kingston, TN; Steubenville, OH.
4Used in other regression analyses not shown in this table.
5Value may be affected by filter artifacts.

-------
  1     database.  However, no other non-PM pollutant measures were investigated in the ACS
  2     analysis, so that no further progress was made in attempting to clarify relative contributions
  3     of PM or other possible "responsible" pollutants.  Also of note were that the ACS regression
  4     coefficients were about 1/4 to 1/2 of the corresponding Six City values and were much closer
  5     to the corresponding values obtained in various acute mortality studies.  Thus it is not clear
  6     to what extent chronic effects (as opposed to integrated acute effects) are indicated by these
  7     results and to what extent the limited air quality data base used was responsible  for this
  8     outcome.
  9          The  California and Six-City studies both suffer from small sample sizes and inadequate
 10     degrees of freedom, which partially offset the specificity gained by considering individuals
 11     instead of population groups.  All of them may have neglected some important risk factors.
 12     The studies of California nonsmokers by Abbey et al.  (1991, 1994) that had the best
 13     cumulative exposure estimates found no significant mortality effects of previous  air pollution
 14     exposure.   The Six Cities  and ACS studies agree in  their findings of strong associations
 15     between fine particles and excess mortality.  At this time, the long-term studies appear
 16     mainly to  provide support for the existance of short-term PM-related mortality increases,
 17     which are not subsequently offset by decreases below normal rates.  However, they do not
 18     exclude the possible existence of additional chronic exposure effects; nor do they provide
 19     convincing evidence as to  the specific pollutant(s) involved; and they do not rule out the
 20     possible existence of pollutant thresholds.
 21
 22     Morbidity Outcomes Associated With PM Exposure
 23           Dockery and Pope (1994) reviewed the effects  of PM on  respiratory mortality and
 24      morbidity.  The authors considered five primary health endpoints: mortality, hospital usage,
 25      asthma attacks, respiratory symptoms and lung funtion. In order to include as many studies
 26      as possible, they converted both British smoke and TSP measurements to PM10.  Results
27      from each study were converted to an estimated percent change in the health endpoint per
28      10 /xg/m3 PM10.  These converted results were then  combined across studies of similar
29      endpoints using the standard inverse variance weighted method (fixed effects model).  The
30      authors concluded that there was  a coherence of effects across the endpoints, with most
31      endpoints showing a one to three percent change per 10 fig/m3  PM10. Pulmonary function

        April 1995                               ^49       DRAFT-DO NOT QUOTE OR CITE

-------
 1     showed a smaller change of 0.15 percent for FEV and 0.08 percent for PEFR. These
 2     smaller percent changes are to be expected because there is much less variation in pulmonary
 3     function measurements than in the other measures.  The limitations of the methodological
 4     considerations as they pertain to quantitative assessment of the subject individual studies are
 5     discussed in Chapter  12.  Dockery and  Pope (1994) also noted such limitations in their
 6     review.
 7           The primary difficulties in combining studies can be summarized  as follows.  Most
 8     studies used several endpoints and it is  not clear that results for all of the different endpoints
 9     were reported. Most studies used different lag times or moving averages for the pollutants,
10     and in some cases reported only those which gave positive results.  For those studies which
11     did report results for similar endpoints, many were analyzed with different statistical models.
12     The short-term studies must take into account serial correlation,  and this was done in a
13     variety of ways in those studies which did adjust for it. For these reasons, key findings from
14     most of the studies are only summarized here rather than combined formally.
15
16     Short-Term PM Exposure Hospital Admission Studies
17           Hospitalization data  can provide a measure of the morbidity status of a community
18     during a specified time frame. Hospitalization data specific for respiratory illness  diagnosis,
19     or more specifically for COPD and pneumonia, index respiratory health status and provide
20     outcome measures which  relate to mortality studies for total and specified respiratory
21     measures.  Tables 1-7 through 1-10 summarize studies that associate hospitalization data with
22     various measures of PM.  Some  of the  same factors and concerns related to the mortality
23     studies are  at issue for these studies also.
24           Both COPD and pneumonia hospitalization studies show moderate, but statistically
25     significant relative risks in the range of 1.06 to 1.25 in association with an increase of 50
26     Mg/m3 m PMIO or its equivalent. There are also indications of a relationship with heart
27     disease,  admissions, but the evidence is less clear.  Overall, these hospitalization studies are
28     indicative of health outcomes related to PM.  They are also supportive of the mortality
29     studies, especially with the more specific diagnosis relationships.
        April 1995                                1-50      DRAFT-DO NOT QUOTE OR CITE

-------
H

6
o

2
o
H

O
e;
o
H
w
o
Study
Burnett et al. (1994)
All ages in Ontario,
Canada, 1983-1988

Thurstonet al. (1994)
All ages in Ontario,
Canada, July and August,
1986-1988


Thurston et al. (1992)
All ages in Buffalo,
Albany, New York City,
July and August, 1988-1989

Schwartz (in press)
Elderly in New Haven,
1988-1990



Schwartz (in press)
Elderly in Tacoma, 1988-
1990



PM Type &
No. Sites
9 monitoring
stations
measuring
sulfate
3 monitoring
stations
measuring
sulfate, TSP,
and PM10

3 monitoring
stations (one
per city)
measuring
sulfate, H +
PM10
monitoring
stations
averaged, no.
of stations not
given
PM10
monitoring
stations
averaged, no.
of stations not
given
PM Mean Ave. Count
& Range per Day
sulfate means 108
ranged from 3 . 1
to 8.2 jtg/m3

mean sulfate 14.4
ranged 38 to 124
(nmole/m3), PM10
30 to 39 jig/m3,
TSP 62 to 87
/ig/m3
(values not given) Buffalo, 24
Albany, 12,
New York,
137

mean = 41, 8.1
10% tile = 19,
90% tile = 67



mean = 37, 4.2
10% tile = 14,
90% tile = 67



Model Type
&Lag
Structure
Lin. regress.
on filtered
data, 1-d lag
best
Linear
regression on
filtered data,
0-d lag best


Linear
regression on
filtered data


Poisson log-
linear
regression,
19 day mov.
ave. filter,
0-d lag best
Poisson log-
lin. regress.
19 day mov.
ave. filter,
0-d lag best

Other pollutants
measured
Ozone



Ozone, H + , SO2,
NO2




Ozone, H +
Weather &
Other Factors
Temperature



Temperature





Temperature
Result*
Pollutants (Confidence
in model Interval)
none 1 .03
(1.02,


none PM10
1.09
(0.96,
ozone PM10
1.01
(0.87,

1.04)




1.22)


1.15)
ozone (not given for
PM measures)



Ozone (ppb): mean
= 29; 10% tile =
16; 90% tile = 45;
SO2 (ppb): mean =
30; 10% tile = 9;
90% tile = 61
Ozone (ppb): mean
= 25; 10% tile =
13; 90% tile = 36;
SO2 (ppb): mean =
17; 10% tile = 6;
90% tile = 28



Temperature
and dew point
adjusted for in
the moving
average

Temperature
and dew point
adjusted for in
the moving
average




none 1 .06
(1.00,

SO2(2day 1.07
lag) (1.01,

none 1.10
(1.03,

SO2 (2 day 1.11
lag) (1.02,





1.13)


1.14)


1.17)


1.20)

     * Relative risk calculated from parameters given by author assuming a 50 /ig/m3 increase in PM10 on 100 /xg/m3 increase in TSP.

-------
TABLE 1-8. HOSPITAL ADMISSIONS STUDIES FOR COPD
l-t
vo
vo






>_>
N>



0
!>
Tl
O1
o
2
o
H
0
o
H
W
O
Q
W
Study
Sunyer et al. (1993)
Adults in Barcelona,
1985-1989

Schwartz (1994f)
Elderly in Minneapolis,
1986-1989


Schwartz (1994e)
Elderly in Birmingham,
1986-1989


Schwartz (1994d)
Elderly in Detroit
1986-1989



* Relative risk calculated






PM Type &
No. Sites
15 monitoring
stations measuring
black smoke

6 monitoring
stations measuring
PM10


1 to 3 monitoring
stations measuring
PM10


2 to 11 PM10
monitoring stations,
data available for
82% of possible
days

from parameters given






Ave.
PM Mean Count
& Range per Day
winter 33% tile 12
= 49, 67% tile
= 77, summer
33% tile = 36,
67% tile = 55

mean = 36, 10% 2.2
tile = 18, 90%
tile = 58


mean = 45, 2.2
10% tile = 19.
90% tile = 77


mean = 48, 5.8
10% tile = 22,
90% tile = 82



by author assuming a 50 ^ig






Model Type
&Lag
Structure
Autoregressive
linear
regression
analysis, 0-d
lag best

Autoregressive
Poisson model,
1 -d lag best


Autoregressive
Poisson model,
0-d lag best


Poisson auto-
regressive
model using
GEE, 0-d lag
best

;/m increase in PM






Other
pollutants
measured
Sulfur dioxide,
winter 33% tile
= 49/tg/m3,
67%tile = 77,
summer
33%tile = 36,
67%tile = 55

Ozone, mean
= 26 ppb,
10%tile =11,
90% tile = 41

Ozone, mean
= 25 ppb,
10%tile = 14,
90% tile = 37

Ozone, mean
= 21 ppb,
10%tile = 7,
90% tile = 36


10 or 100 £ig/m3






Weather & Result*
Other Pollutants (Confidence
Factors in model Interval)
min temp, none winter: 1.15
dummies for day (1.09,1.21)
of week and summer: 1.05
year (0.98,1.12)
SO2 winter: 1.05
(1.01,1.09)
summer:
1.01
(0.97,1.05)
8 categories of none 1.25
temp. & dew (1.10,1.44)
pt., month,
year, lin. &
quad, time trend
7 categories of none 1.13
temp. &dew (1.04,1.22)
pt., month,
year, lin. &
quad, time trend
Dummy vars. ozone 1.11(1.04,
for temp, 1.17)
month, lin. &
quad, time trend


increase in TSP.







-------
                             TABLE 1-9.  HOSPITAL ADMISSIONS STUDIES FOR PNEUMONIA
U)
Tl


U
o

2!
o
H

O
d
o
H
w
Study
Schwartz (1994f)
Elderly in Minneapolis,
1986-1989
Schwartz (1994e)
Elderly in Birmingham,
1986-1989
Schwartz (1994d)
Elderly in Detroit
1986-1989
PM Type & PM Mean
No. Sites & Range
6 monitoring stations mean = 36,
measuring PM10 10% tile =18,
90% tile = 58
1 to 3 monitoring mean = 45,
stations measuring 10% tile = 19.
PM10 90% tile = 77
2 to 11 PM10 mon. mean = 48,
stations, data for 10% tile = 22,
82% of possible days 90% tile = 82
Ave.
Count
per Day
6.0
5.9
15.7
Model Type
& Lag Other pollutants
Structure measured
Autoregressive Ozone: mean 26
Poisson mod., ppb; 10% tile 11;
1-d lag best 90% tile 41
Autoregressive Ozone: mean 25
Poisson mod., ppb; 10% tile 14;
0-d lag best 90% tile 37
Poisson auto- Ozone: mean 21
regress, mod. ppb; 10% tile 7;
using GEE, 90% tile 36
0-d lag best
Weather &
Other Pollutants
Factors in model
8 categories of temp. &none
dew pt., month, year,
lin. & quad, time trend
7 cat. of temp. & dew none
pt., month, year, lin.
& quad, time trend
Dummy variables for ozone
temp, month, lin. &
quad, time trend
Result*
(Confidence
Interval)
1.08
(1.01,1.15)
1.09
(1.03, 1.15)
1.06
(1.02, 1.10)
TABLE 1-10. HOSPITAL ADMISSIONS STUDIES FOR HEART DISEASE
Study
Schwartz and Morris
(in press)
Elderly in Detroit
1986-1989
Ischemic Heart Disease
Burnett et al. (in press)
All ages in Ontario,
Canada, 1983-1988
Cardiac disease admission
PM Type & PM Mean
No. Sites & Range
2 to 11 PM10 mean = 48,
monitoring 10% tile = 22,
stations, data 90% tile = 82
available for
82% of possible
days
22 sulfate station means
monitoring ranged from 3.0 to
stations 7.7 in the summer
and 2.0 and 4.7 in
the winter
Ave.
Count
per Day
44.1
14.4
Model Type
& Lag Other pollutants
Structure measured
Weather &
Other Pollutants
Factors in model
Poisson auto- SO2, mean = 25 Dummy vars. for none
regressive ppb, 10% tile = temp, month, lin.
model using 11, 90% tile = 44 & quad, time trend
GEE, 0-d lag CO, mean 2.4 ozone,
best ppm, 10% tile 1.2, CO, SO2
90% tile = 3.8
Linear Ozone averaged 36 Temperature none
regression on a ppb included in
19 day linear separate analyses
filter, 1-d lag by summer and ozone
best winter
Result*
(Confidence
Interval)
1.06
(1.02, 1.10)
1.06
(1.02, 1.10)
1.04
(1.03, 1.06)
1.04
(1.03, 1.05)
2  * Relative risk calculated from parameters given by author assuming a 50 /ig/m3 increase in PM10 on 100 ^g/m3 increase in TSP.

O

3

-------
 1      Short and Long-Term Exposure Respiratory Disease Studies
 2                 Respiratory illness and the factors determining its occurrence and severity are
 3      important public health concerns. This effect is  of public health importance because of the
 4      widespread potential for exposure to PM and because of the very common occurrence of
 5      respiratory illness.  Of added importance is the fact that recurrent childhood respiratory
 6      illness may be a risk factor for later susceptibility to lung damage.  The occurrence of lower
 7      respiratory morbidity in early childhood may be  associated with impaired lung function and
 8      growth that appears to persist through adolescence, and certain physicians assert that
 9      infections, reactive airways, and inhaled pollutants (mostly cigarette smoke) are the most
10      important risk factors in the development of chronic lung disease.  Thus, factors such as the
11      presence  of PM  (which increases the risk for respiratory  symptoms and related respiratory
12      morbidity) are important because of associated public health concern with regard to both the
13      immediate symptoms produced and the longer term potential for  increases in the development
14      of chronic lung disease.
15
16      Acute Respiratory Disease  Studies
17           Acute respiratory disease studies include several different endpoints, but the majority of
18      authors reported results on  at least two of:  (1) upper respiratory illness;  (2) lower
19      respiratory illness; or (3) cough (See Table 1-11).  These relative risks are all estimated for
20      an increase of 50 /ig/m3 in  PM10 (24-h) or its equivalent.  The results for upper respiratory
21      illness are very inconsistent: two studies estimate a relative risk near 1.00 whereas four
22      others obtain estimates between 1.14 and 1.55.  The relative risks for lower respiratory
23      illness are spread between 1.01  and 2.03, but  all are positive.  The relative risks for cough
24      include two below 1.0 and  go as high as 1.51. All of these are generally suggestive of a PM
25      effect, but whereas the hospital  admission studies were all done in a similar manner and
26      resulted in very similar results,  these studies used different designs and yield very
27      inconsistent results.
28
29      Chronic Respiratory Disease Studies
30           The three studies  listed (Table 1-12) are  based on a similar type of questionnaire but
31      were done by two different groups of researchers.  All three studies suggest a chronic effect

        April 1995                                 1-54      DRAFT-DO NOT QUOTE OR CITE

-------
                          TABLE 1-11.  ACUTE RESPIRATORY DISEASE STUDIES
H

6
o

z
o
H

O
G
O
H
W
Study
Schwartz et al. (1994) 300
elementary school children
in Six-Cities in U.S.,
1984-1988

Popeet al. (1991),
students in the Utah
Valley, winter 1989-1990






Popeet al. (1991),
asthmatic children in the
Utah Valley, winter 1989-
1990



Pope and Dockery (1992),
symptomatic children in
the Utah Valley, winter
1990-1991


PM Type &
No. Sites
PM10
monitoring in
each city


PM10
monitoring
stations at 3
sites





PM10
monitoring
stations at 3
sites



PM10
monitoring
stations at 2
sites


Ave.
PM Mean Rate
& Range per Day
median 30 /ig/m3 3.1
10th percentile =
13, 90th
percentile = 53

mean = 46 (not given)
Hg/m3,
range = 11 to
195





mean = 46 (not given)
^g/m3,
range = 11 to
195



mean = 76 (not given)
fig/m3,
range = 7 to 251



Model Type
& Lag Structure
Autoregressive
logistic
regression using
GEE

Fixed effects
logistic
regression






Fixed effects
logistic
regression




Autoregressive
logistic
regression using
GEE


Other
pollutants
measured
Ozone, NO2,
S02



Limited
monitoring of
NO2, SO2, and
ozone. Values

were well
below the
standard

Limited
monitoring of
NO2, SO2, and
ozone. Values
were well
below the
standard
none





Weather & Other
Other pollutants
Factors in model
Temperature none

S02

ozone
Variables for none
temperature and
time trend






Variables for none
low
temperature and
time trend



Variable for none
low
temperature



Result*
(Confidence
Interval)
1.51 (1.12,

1.39(0.98,

1.49(1.10,
Upper resp.
1.20(1.03,


Lower resp.
1.28(1.06,


Upper resp.
0.99(0.81,


Lower resp,
1.01 (0.81,


Upper resp.
1.20(1.03,
Lower resp
1.27(1.08,
Cough
1.29(1.12,

2.05)

1.96)

2.01)

1.39)



1.56)



1.22)



1.27)



1.39)

1.49)

1.48)
n
HH

m

-------
                        TABLE 1-11 (cont'd). ACUTE RESPIRATORY DISEASE STUDIES
2:
VO
S












H-
Ov



O
£
H
6
o
z
o
H
0


H
W
O


Study
Pope and Dockery (1992),
asymptomatic children in
the Utah Valley, winter
1990-1991





Hoek and Brunekreef
(1993), respiratory disease
in school children aged 7
to 12 in Wageningen,
Netherlands, winter
1990-1991



Schwartz et al. (1991)
Study of acute respiratory
illness in children in
5 German communities,
1983-1985








PM Type &
No. Sites
PM10
monitoring
stations at
2 sites





Two to 4
monitoring
stations
measured
PM10




Two to 4
monitoring
stations in
each area
measured
TSP






Ave.
PM Mean Rate
& Range per Day
mean = (not given)
76 /ig/m3,
range = 7 to
251





max = (not given)
1 10 /ig/m3







medians ranged 0.5 to 2.9
from 17 to
56 /ig/m3,
10% tiles from
5 to 34, 90%
tiles from 41 to
118





Model Type
&Lag
Structure
Autoregressive
logistic regression
using GEE






Autoregressive
logistic regression
using GEE






Autoregressive
Poisson regression
using GEE









Other
pollutants
measured
none








Max SO2 = 105
ftg/m3, max NO2
= 127 /tg/m3






median SO2
levels ranged
from 9 to
48 /tg/m3,
median NO2
levels ranged
from 14 to
5 /xg/m3




Weather & Other
Other pollutants
Factors in model
Variable for none
low temperature







Variable for none
ambient
temperature and
day of study





Most none (TSP
significant was not
terms of day significant
of week, time when NO2
trend, and added to model)
weather
(terms not
listed)




Result*
(Confidence
Interval)
Upper resp.
0.99
(0.78, 1.26)
Lower resp.
1.13
(0.91, 1.39)
Cough
1.18
(1.00, 1.40)
Upper resp.
1.14
(1.00, 1.29)
Lower resp.
1.06
(0.86, 1.32)
Cough
0.98
(0.86, 1.11)
1.26
(1.12, 1.42)










n
HH
H
W

-------
TABLE 1-11 (cont'd).  ACUTE RESPIRATORY DISEASE STUDIES
d:
h-t
u?













K-t
1




O
>
2
O1
0
^
o


o
H
W
O
n
H


Study
Schwartz et al. (1994)
Study of respiratory
symptoms in 6 U.S.
cities, 1984-1988








Braun-Fahrlander et al.
(1992)
Study of preschool
children in four areas
of Switzerland
Roemer et al. (1993)
Study of children with
chronic respiratory
symptoms in
Wageningen, The
Netherlands
Dusseldorf et al.
(1994)
Study of adults near a
steel mill in The
Netherlands





PM Type &
No. Sites
Daily
monitoring
of PM10,
PM2 5 at
each city







Daily
monitoring of
TSP


Daily
monitoring of
PM10



Daily
monitoring of
PM[0, iron,
sodium,
silicon, and
manganese



Ave.
PM Mean Rate
& Range per Day
median PM10 (not given)
= 30 iKg/m3,
10% tile = 13,
90% tile = 53
median PM, 5
= 18 /ig/m ,
10% tile = 7,
90% tile = 37




(not given) 4.4




6 days above .094
110 ng/m3 incidence
rate



mean PM10 (not given)
= 54 pig/m3,
range =
4 to 137)





Model Type
&Lag
Structure
Autoregressive
logistic
regression
using GEE








Logistic
regression



Autoregressive
logistic
regression



Logistic
regression







Other Weather &
pollutants Other
measured Factors
SO2, median = 4 temperature,
ppb, 10% tile = day of week,
1, 90% tile = 18 city or
NO2, median = residence
13 ppb, 10% tile
5, 90% tile = 24,
ozone





SO2, NO2, and city, risk
ozone levels not strata, season,
given temperature


SO2 and NO2 (not given)
means not given




Geometric mean (not given)
iron = 501
ng/m3, manganese
= 17 ng/m3,
silicon =
208 ng/m3



Other Result*
pollutants (Confidence
in model Interval)
all two Cough
pollutant (PMIO lag 1)
models were 1.51
fitted with (1.12,2.05)
minimal effect Upper resp.
on PM (PMIO lag 2)
1.39
(0.97, 2.01)
Lower resp.
(PM10 lag 1)
2.03
(1.36,3.04)
none Upper resp.
1.55
(1.10,2.24)


none Cough
(not given,
probably less than
one)


none Cough
1.14
(0.98, 1.33)







-------
                                   TABLE 1-11 (cont'd).  ACUTE RESPIRATORY DISEASE STUDIES
Ui
oo
o
O
2J
O
H
O
d
o
H
W


Study

PM Type &
No. Sites

PMMean
& Range
Ave.
Rate
per Day
Model Type
&Lag
Structure
Other
pollutants
measured
Weather &
Other
Factors
Other
pollutants
in model
Result*
(Confidence
Interval)
     Ostro et al. (1991)
     Study of adult asthmatics
     in Denver, Colorado
     November 1987 to
     February  1988
     Ostro et al. (1993)
     Study of non-smoking
     adults in Southern
     California
Two monitors
provided daily
measurements of
PM2.5
22 /ug/m3, range 15 (out of 108)  Autoregressive nitric acid,
= 0.5 to 73                    logistic        sulfates,
                              regression     nitrates, SO2,
                                            andH+
day of survey,    none
day of week, gas
stove, minimum
temperature
Apparently one    mean sulfate = 84.2/person for
site (Azusa).      /*g/m3, range =  lower,
PM measurements 2 to 37 mean    10.2/person,
included sulfate    COHS = 12 per  upper
fraction and       100 ft, range
COHS           = 4 to 26
                              Logistic       ozone, mean    temperature, rain none
                              regression     = 7 pphm,     humidity
                                            range = 1 to 28
Cough
1.09
(0.57,2.10)
                            Sulfates:
                            Upper resp.
                            0.91
                            (0.73, 1.15)
                            Lower resp.
                            1.48
                            (1.14, 1.91)
H-   * Relative risk calculated from parameters given by author assuming a 50 /zg/m3 increase in PM10 on 100 /*g/m3 increase in TSP.
n
HH
H
W

-------
                         TABLE 1-12. CHRONIC RESPIRATORY DISEASE STUDIES
eL
Co
Ul














1
Ui
VO


O
g
^
1
O
O
z
o
H
O
C!
0
H
W
O


Study

Ware et al. (1984)
Study of respiratory
symptoms in children in 6
cities in the U.S. Survey
done 1974-1977

Dockery et al. (1989)
Study of respiratory
symptoms in children in 6
cities in the U.S. Survey
done 1980-1981

Chapman et al. (1985)
Study of persistent cough
and phlegm (bronchitis) in
adults in four
communities in Utah.
Survey done in 1976
Neas et al. (1994)
Study of children aged
7 to 1 1 from six cites in
U.S. Survey done
1983-1986.







* Estimates calculated from

PM Type &
No. Sites

Daily monitoring
of TSP, SO2,
NO2, and ozone at
each city


Daily monitoring
of PM15, sulfate
fraction at each
city


Daily monitoring
of TSP, and
sulfate fraction at
each city


PM2.5












PM Mean
& Range

City TSP means
ranged from 39
to 1 14 /tg/m3



City PM15
means ranged
from 20 to 59
/*g/m3


Previous 5 year
TSP ranged from
11 to 115 /xg/m3



Not given











Overall
Symptom
Rate

Cough, .08,
Bronchitis
.08,
Lower resp.
.19

Cough, .02 to
.09, Bronchitis
.04 to .10,
Lower resp.
.07 to .16

.02 to .05 by
city




Not given











data tables assuming a 50 /*g/m3 increase in PM]0 on
Model
Type
&Lag
Structure
Logistic
regression




Logistic
regression




Logistic
regression




Logistic
regression










100 /ig/m3
Other
Other Other pollutants
pollutants Covariates in model
measured
SO2, NO2, age, gender, none
and ozone parental
education,
maternal
smoking

S02, NO2, age, gender, none
and ozone maternal
smoking



SO2, NO2 smoking none





NO2 household none
smoking, gas
stove, age,
gender none


none





increase in TSP.
Result*^
(Confidence
Interval)

Cough
2.75 (1.92,
Bronchitis
2.80(1.17,
Lower resp.
2.14(1.06,
Cough
5.39(1.00,
Bronchitis
3.26(1.13,
Lower resp.
2.93 (0.75,
Mothers
1.75(1.21,
Fathers
1.94(1.16,


Cough
1.08(0.76,
Bronchitis
1.32(0.98,
Lower resp
1.23 (0.98,












3.94)

7.03)

4.31)

28.6)

10.28)

11.60)

2.54)

3.25)



1.53)

1.79)

1.55)







O
a

-------
 1     of paniculate matter on respiratory disease,  but the studies suffer from the usual difficulty of
 2     cross sectional  studies.  The PM effect estimates are based on variations in exposure which
 3     are determined by different numbers of locations.  In the first two studies there were  six
 4     locations and, in the second, four.  The results seen were consistent with a PM gradient, but
 5     it is impossible to separate out the effect of PM any other factors or pollutants which have
 6     the same gradient.
 7
 8     Short and Long-Term Exposure Pulmonary Function Studies
 9           Pulmonary  function studies are part of a comprehensive investigation of the possible
10     effects of any air pollutant. Measurements  can be made in the field, they are noninvasive,
11     and their reproductibility  has been well documented; and guidelines for reference values and
12     interpretative strategies of lung function tests have been prepared.  Various factors are
13     important determinants of lung functions. Lung function in children has been related to
14     genetic factors that exert their greatest influence through general stature as measured  by
15     height and age.  Growth patterns in children differ by gender and lung function declines with
16     age among adults.  Studies of the  growth of pulmonary function and generalized growth
17     models consider  factors of how growth is statistically dependent on initial measures of
18     function, and how it is related to respiratory illness in childhood.  The effects of active
19     smoking and passive smoking are  also considered.  Epidemiological studies relating ambient
20     PM measures to  decrements in pulmonary function represent a potentially important health
21     effect.
22           The acute pulmonary function studies  (Table 1-13) are suggestive of a short term effect
23     resulting from  paniculate pollution. Peak flow rates  show decreases in the range of 30 to 40
24     ml/sec to be associated with an increase of  50 ^g/m3 in PM10 (24-h) or its equivalent.  The
25     results appear to be larger in symptomatic groups  such as asthmatics.  The effects are seen
26     across a variety of study designs,  authors, and analysis methodologies.  Effects using FEVj
27     or  FVC as endpoints are  less  consistent.
28           The chronic pulmonary function studies are less numerous than the acute studies
29     (Table 1-13).  The one study  with good monitoring showed no effect from paniculate
30     pollution.  Cross sectional studies require very large  sample sizes to detect differences
31     because the studies cannot eliminate person to person variation which is much larger  than the

       April 1995                                1-60       DRAFT-DO NOT QUOTE OR CITE

-------
                  TABLE 1-13. ACUTE AND CHRONIC PULMONARY FUNCTION CHANGES
3:
vJD

3
a
%^
o
z
g
r— 5
O

|
w
o

Study Period,
Population
Dockery et al. (1982)-
School age children in
Steubenville, OH,
measured at three times
between 1978 and 1980
Dassen et al. (1986)
School age children in
The Netherlands,
measured in November,
1984 and January, 1985
Quackenboss et al.
(1991) - Asthmatic
children aged 6 to 15
years in Tuscon, AZ,
measured in May and
November, 1988
Pope et al. (1991) -
Study of asthmatic
children in the Utah
Valley
Pope and Dockery
(1992) - Study of non-
asthmatic symptomatic
and asymptomatic
children in the Utah
Valley





PM Type & PM Mean
No. Sites & Range
Single station up to 455 /Kg/m3
measuring TSP



Six station network TSP and RSP both
measuring TSP, exceeded 200
RSP (PM10) /ig/m3


Individual
monitoring of
homes of PM2 5,
PM10


PMIO monitors in PM10 ranged from
Orem and Lindon, 11 to 195 /ig/m3
Utah

PM10 monitors in PM10 ranged from
Orem and Lindon, 11 to 195 pig/m3
Utah







Model Type Other
& Lag pollutants
Structure measured
individual SO2
regression analyses
for each child,
coefficients pooled
across time
multiple linear SO2
regression



Random effects NO2
linear model




Weighted least SO2, NO2, ozone
squares regression


Weighted least SO2, NO2, ozone
squares regression








Weather &
Other Pollutants
Factors in model
average TSP
temperature



technician, RSP
appliance,
presence of
colds

temperature, PM2 5
wind speed,
dew point



low PM10
temperature


low PM10
temperature








Decrease*
(Confidence
Interval)
FVC: 8.1 ml
FEV075: 1.8ml
Note: decreases
were statistically
significant
slopes not given but
FVC, FEVj, and
PEFR were
significantly reduced
during episodes
PEFR: 375 ml/s
Note: these are
diurnal rather than
daily changes


PEFR: 55 ml/s
(24, 86)


Symptomatic
PEFR
30 ml/s
(10, 50)
Asympto-
matic PEFR
21 ml/s
(4, 38)


O

H
W

-------
               TABLE 1-13 (cont'd). ACUTE AND CHRONIC PULMONARY FUNCTION CHANGES
i













H- '
OS



O
!>
3
a'
0
2
O
H
G
O
H
W
O
^-\
Study Period,
Population
Koenig et al. (1993) -
Study of asthmatic and
non-asthmatic elementary
school children in Seattle,
WA in 1989 and 1990





Hoek and Brunekreef
(1993) - Study of children
aged 7 to 12 in
Wageningen, Netherlands

Roemer et al. (1993) -
Study of children with
chronic respiratory
symptoms in The
Netherlands
Pope and Kanner (1993) -
Study of adults in the Utah
Valley from 1987 to 1989


Neas et al. (1994) -
Study of lung function in
children in 6 cities in the
U.S. Data collected from
1983-1988.

PM Type &
No. Sites
PM2 5 calibrated
from light
scattering







Single site
measure black
smoke. PM10 was
measured during
episodes
Single site
measure black
smoke. PM10 was
measured using an
Anderson dichot
PM10 was collected
daily from the north
Salt Lake site


Daily monitoring of
PM2 5, sulfate
fraction at each city



PM Mean
& Range
PM2 5 ranged
from 5 to
45 /tg/rn







range of PM10
was 30 to
144 ftg/m3


range of PM10
was 30 to
144 /ig/rn3


PM10 daily
mean =
55 /ig/m ,
ranged from 1
to 181 /ig/m3
not given





Model Type Other
& Lag pollutants
Structure measured
Random effects none
linear regression








SAS procedure SO2, NO2
AUTOREG



multiple linear SO2, NO2
regression
analysis


Linear regression Limited
on difference in monitoring of
PFT as a function SO2, NO2, and
of PM10 ozone

Linear regression SO2, NO2, and
using logarithm of ozone
PFT value



Weather &
Other Pollutants
Factors in model
height, PM25
temperature








day of study PM10




none PM10




low PM2 5
temperature



city, gender PM2 5
parental
education,
history of
asthma, age,
height, weight
Decrease*
(Confidence
Interval)
Asthmatics
FEVj 42 ml
(12, 73)
FVC 45 ml
(20, 70)
Non-asthmatics
FEVj 4 ml
(-7, 15)
FVC -8 ml
(-20, 3)
PEFR
41 ml/s (-8, 90)



PEFR
34 ml/s (9, 59)



FEV!
29 ml (7, 51)
FVC 15 ml (-
15, 45)

FVC and FEVi not
changed. Values
could not be
converted to mis.


Decreases in lung function calculated from parameters given by author assuming a 50
increase in PM10 or 100 /ig/m3 increase in TSP.

-------
 1      within person variation.  Thus the lack of statistical significance cannot be taken as proof of
 2      no effect.
 3
 4      Comparison of Effects of PM10 Versus PM2>5 on Respiratory Disease and Pulmonary
 5      Function
 6           The most direct comparison of effects of PM10 versus PM2 5 are possible when studies
 7      include both exposure measures in their analyses.  This occurred in the Six City study for
 8      Steubenville children, the Tuscon study, and the Uniontown study  for pulmonary function
 9      changes (Table 1-13). None of these studies directly showed that one of these PM measures
10      was a significantly better predictor than the other for pulmonary function effects. The study
11      did suggest that PM2 5 may be better predictor of lung function change.
12           A few other studies used PM2 5 as a measure of paniculate exposure.  One on lung
13      function in asthmatic and non-asthmatic school children in Seattle (Table 1-13) found a
14      slightly larger effect of PM2 5 for asthmatics, but a slightly smaller effect for non-asthmatics
15      when compared with the PM10 studies.  Also,  the Ostro et al. (1991) study of respiratory
16      disease in Denver found an effect that was in the middle of the range of effects found by the
17      PM10 studies.
18           Based on the above information, there is currently no obvious way by which to clearly
19      distinguish morbidity effects of PM10 versus PM2 5.  Even the suggestive evidence leaves the
20      scales in a balanced  position.
21
22      Mortality Effects of Acid Aerosols
23           Few epidmiological studies have examined mortality data for an association with
24      ambient paniculate strong acid aerosol  (H+) exposures.  The scarcity of the analyses is due
25      to the absence of adequate ambient acid measurement techniques in the past, and to the lack
26      of routine acid aerosol monitoring in more recent years.  Some studies now exist which
27      suggest that human health effects may be associated with expsoures to ambient acid aerosols,
28      both:  (1) as derived from reexamination of older, historically important data on air pollution
29      episode events in North America and Europe and (2) as can be deduced from limited recent
30      epidemiology studies carried out in the U.S., Canada, and Europe.
       April 1995                                1_63       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Historical and present-day evidence suggests that strongly acidic PM can be associated
 2     with both acute and chronic human health effects.  Evidence from historical pollution for
 3     episodes, notably the London Fog episodes of the 1950's and early  1960's, indicates that
 4     extremely elevated daily acid aerosol concentrations (on the order of 400 /xg/m3 as H2SO4, or
 5     roughly 8,000 nmoles/m3 H+) may be associated with excess acute  human mortality when
 6     present as a copollutant with elevated concentrations of PM and SO2. In addition, Thurston
 7     et al. (1989) and Ito et al. (1993) both found significant associations between acid aerosols
 8     and mortality in London during non-episode pollution levels (< 30  /xg/m3 as  H2SO4, or
 9      <  approximately 600 nmoles/m3 H+), although these associations could not be separated
10     from those for BS or SO2.  The only attempts to date to associate present-day levels of acidic
11     aerosols with acute and chronic mortality (Dockery et al., 1992; Dockery  et al., 1993b,
12     respectively) were unable to do so, but weaknesses  in these analyses (in particular, too
13     limited H+ data for the analysis) may have made associations undetectable. At very high
14     concentrations that do not occur in the ambient air, mortality in laboratory animals can occur
15     following acute exposure, due primarily to laryngeal or bronchoconstriction; larger particles
16     are more effective in this regard than are smaller ones.
17
18     Respiratory Illness  Effects of Acid Aerosols
19           Historical and present-day evidence suggests that there can be  both acute and chronic
20     effects of strongly acidic PM on human health.  Increased hospital admissions for respiratory
21     causes were documented during the London Fog episode of 1952, and this association has
22     now been observed under present-day conditions, as well. Thurston et al.  (1992) and
23     Thurston et al. (1994) have noted  associations between ambient acidic aerosols and
24     summertime  respiratory hospital admissions in both New York State and Toronto, Canada,
25     respectively, even after controlling for potentially confounding temperature effects.  In the
26     latter of these studies, significant independent H+ effects remained even after simultaneously
27     considering the other major copollutant, O3, in the regression model. In these studies, H+
28     effects were  estimated to be the largest during acid aerosol episodes (H+  > 10 )ug/m3 as
29     H2SO4, or =200 nmoles/m3 H+), which occur roughly 2 to 3 times per year in eastern
30     North America.  These studies provide evidence that present-day  strongly acidic aerosols
        April 1995                               1-64       DRAFT-DO NOT QUOTE OR CITE

-------
  1      may represent a portion of PM which is particularly associated with significant acute
  2      respiratory disease health effects in the  general public.
  3           Results from recent acute symptoms studies of healthy children indicate the potential for
  4      acute acidic PM effects in this population.  While the 6-City study of diaries kept by parents
  5      of children's respiratory and other illness did not demonstrate H+ associations with lower
  6      respiratory symptoms except at H+ above 110 nmoles/m3 (Dockery et al.,  1994), upper
  7      respiratory symptoms in two of the cities were found to be most strongly associated with
  8      daily measurements of H2SO4 (Schwartz, et al. 1991).
  9           Studies of the effects of chronic H+ exposures on children's respiratory health and lung
10      function are generally consistent with effects as a result of chronic H+ exposure.
11      Preliminary analyses of bronchitis prevalence rates as reported across the 6-City study locales
12      were found to be more  closely associated with average H+ concentrations than with PM in
13      general (Speizer, 1989).  A follow-up analysis of these cities and a seventh locality which
14      controlled the analysis for maternal smoking and education and for race, suggested
15      associations between summertime average H+ and chronic bronchitic and related symptoms
16      (Damokosh et al.,  1993). The relative  odds of bronchitic symptoms with the highest acid
17      concentration (58 nmoles/m3 H+) versus the lowest concentration (16 nmoles/m3) was 2.4
18      (95% CI: 1.9 to 3.2).  Furthermore, in a follow-up study of children in 24 U.S. and
19      Canadian communities (Dockery et al.,  1993a) in which the analysis was adjusted for the
20      effects of gender,  age, parental asthma, parental education, and parental allergies, bronchitic
21      symptoms were confirmed to be significantly associated with strongly acidic PM (relative
22      odds = 1.7, 95%  CI: 1.1 to 2.4).  It was also found in the 24-Cities study that mean FVC
23      and FEV1-0 were lower in locales having high particle strong acidity (Raizenne et al., 1993).
24      Thus, chronic exposures to strongly acidic PM may have effects on measures of respiratory
25      health in children.
26          The respiratory tract has an array  of defense mechanisms to kill,  detoxify,  and
27      physically remove inhaled material, and these defenses may  be altered  by exposure to H2SO4
28      at levels < 1,000 /xg/m3.  Acid aerosols alter mucociliary clearance in  human and laboratory
29      animals, with effects dependent on exposure concentration and the region of the lung being
30      studied.  For example 1- to 2-h resting  exposures of humans to 100 ^ig/m3  accelerate
31      clearance in large  bronchi, but slows clearance in smaller more peripheral airways.

        April 1995                                1-65      DRAFT-DO NOT QUOTE OR CITE

-------
 1     Clearance in asthmatics is also affected, but the results are not clearly interpretable.  Long-
 2     term exposure also affects mucociliary clearance in animals.  For example, in rabbits
 3     exposed intermittently for 125 /ug/m3 H2SO4 for 1 year, clearance was accelerated during
 4     exposure but was depressed 6 months after exposure ceased.  These responses are complex
 5     and are accompanied by histological and chemical changes in mucus and epithelial secretory
 6     cells.  Defenses, such as resistance to bacterial infection, may be altered by acute exposure
 7     to concentrations of H2SO4 around 1,000 jug/m3.
 8          Severe morphologic alterations in the respiratory tracts of animals occur at high acid
 9     levels.  At low levels and with chronic exposure, the main response seems to be hypertrophy
10     and/or hyperplasia of mucus secretory cells in the epithelium; these alterations may extend to
11     the small bronchi and bronchioles, where secretory cells are normally rare or absent.
12          Limited data also suggest that exposure to acid aerosols may affect the phagocytic
13     functioning of alveolar macrophages; the lowest level examined to date is 500 /ig/m3 H2SO4.
14     Alveolar region particle clearance is accelerated by repeated H2SO4 exposures to  as low as
15     250 jug/m3; higher levels retard clearance.  Acute exposure of rabbits to  lower concentrations
16     (e.g., 75 /ig/m3 H2SO4)  can  affect other alveolar  macrophage functions.
17
18     Pulmonary Function Effects of Acid Aerosols
19          Both acute and chronic exposure of laboratory animals to H2SO4 at levels well below
20     lethal ones will produce  functional changes in the  respiratory tract.  The  pathological
21     significance of some of these are greater than for others.  Acute exposure will alter
22     pulmonary function, largely due to bronchoconstrictive action.  However, attempts to
23     produce changes in airway resistance in healthy animals at levels below 1 mg/m3 have been
24     largely unsuccessful, except when the guinea pig  has been used.  The lowest effective level
25     of H2SO4 producing bronchoconstriction to date in the guinea pig is  100 /ig/m3 (1-h
26     exposure).  In general, smaller size droplets are more effective in altering pulmonary
27     function, especially  at low concentrations.  Yet even in  the guinea pig, there are
28     inconsistencies in the type of response exhibited towards acid aerosols.  Chronic exposure to
29     H2SO4 is also associated with alterations in pulmonary function (e.g., changes in the
30     distribution of ventilation and in respiratory rate in monkeys).  But, in these cases, effective
31     concentrations are >500 jug/m3.   Hyperresponsive airways have been induced with repeated

       April 1995                               1-66       DRAFT-DO NOT QUOTE OR CITE

-------
 1      exposures to 250 ptg/m3 H2SO4 in rabbits, and have been suggested to occur following single
 2      exposures at 75 /xg/m3.
 3           Ten human clinical studies since 1988 have confirmed previous findings that healthy
 4      subjects do not experience decrements in lung function following single exposures to H2SO4
 5      at levels up to 2,000 ^tg/m3 for 1 h, even with exercise and use of acidic gargles to minimize
 6      neutralization by oral ammonia.  Mild lower respiratory symptoms occur at exposure
 7      concentrations in the mg/m3 range, particularly with larger particle sizes.
 8           There is no clearly established exposure-response relationship across studies.  Asthmatic
 9      subjects appear to be more sensitive than healthy subjects to the effects of acid aerosols on
10      lung function,  but reported effective concentrations differ widely among studies.  Adolescent
11      asthmatics may be more sensitive than adult asthmatics, and may experience small
12      decrements in lung function in response to H2SO4 at exposure levels only slightly above peak
13      ambient levels (e.g., less than 100 /xg/m3).  Although the reasons for the inconsistency
14      among studies remain largely unclear, individual variability in sensitivity and subject
15      selection may be an important factors.  Even  in studies reporting an overall absence of
16      effects on lung function, occasional asthmatic subjects appear to demonstrate clinically
17      important effects. Two studies from different laboratories have suggested that responsiveness
18      to acid aerosols may correlate with the degree of baseline airway hyperresponsiveness.
19      However, based on  very limited studies, the elderly and individuals with chronic obstructive
20      pulmonary disease do not appear to be particularly more susceptible to the effects of acid
21      aerosols on lung function than healthy adults.
22           Two recent studies have examined the effects of exposure to both H2SO4 and ozone on
23      lung function in healthy and asthmatic subjects. Both studies found evidence that 100 /xg/m3
24      H2SO4 may potentiate the response to ozone,  in contrast with previous studies.  Recent
25      summer camp (and schoolchildren) studies of lung function have also indicated a significant
26      association between  acute exposures to acidic  PM and decreases in the lung function of
27      children independent of those associated with  O3 (Studnicka et al., 1995; Neas et al, 1995).
28           In view of uncertainties about differences between high acid concentrations needed to
29      produce effects in animal studies and low concentrations found in the human environment,
30      the epidemiologic evidence does not establish  a clear role for acid aerosols as a primary
31      agent contributing to ambient PM exposure effects on pulmonary function.

        April 1995                               1_67       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Coherence of Epidemiologic Findings
 2           Factors involved in evaluating both the data and the entire group of epidemiological
 3      studies, include the strength of association, the consistence of the association, as evidenced
 4      by its repeated observation by different persons,  in different places, circumstances and time,
 5      and the consistency with other known facts (Bates, 1992). One can look for
 6      interrelationships between different health indices to provide a stronger and more consistent
 7      synthesis of available information. The various findings that support a picture of coherence
 8      would provide a stronger case with quantitative studies as opposed to  qualitative studies.
 9      Other studies may be inappropriate to use in such a discussion, and the quality of the study
10      should be considered.  Bates (1992) states that the difficulty with discussing any index of
11      internal coherence  is that this requires a series of judgements on the reliability of the
12      individual findings and observations.  The outcome of a coherence discussion then is a
13      qualitative presentation in the end, not quantitative.  Thus, coherence  cannot be formally
14      measured.
15           Bates  (1992) also noted  that the strength of different health indexes are important as are
16      difficulties in assessing exposure.  Bates (1992) also suggests three areas to  look for
17      coherence:  (1) within epidemiological data, (2) between epidemiological and animal
18      toxicological data,  and (3) between epidemiological, controlled human and animal  data.
19           Coherence by its nature considers biological relationships of exposure to health
20      outcome. The biologic mechanism underlying an acute pulmonary function  test reduction in
21      children is most likely not part of the  acute basis for a change  in the mortality rate of a
22      population exposed in an older group of individuals.   In looking  for coherence one should
23      compare  outcomes that look at similar time frames—daily hospitalizations compared  to daily
24      mortality rather than monthly hospitalizations.  Overall the data indicates that PM  has a
25      relationship with a continuum of health outcomes, but the studies may not establish a
26      coherence between them.  The underlying mechanisms may be different.
27           The principal health outcome for which coherence is desirable is mortality, the death
28      rate in a  population.  This can be considered within the  endpoint and/or in other endpoints.
29      Of the various morbidity outcomes studied and discussed in the earlier part of the  chapter,
30      hospitalization studies  reviewed in the chapter support this notion. The  mortality studies
31      suggest that these specific causes provide stronger relationships (i.e., larger  RR estimates)

        April 1995                                 1_68       DRAFT-DO NOT QUOTE  OR CITE

-------
 1     than total mortality.  The outcome potentially most related is hospital admission for
 2     respiratory or cardiovascular causes in the older age group (i.e., >  65 years old).  In a
 3     qualitative sense, the increased mortality found in that age group should also be paralled by
 4     increased hospital admissions.
 5          Partial coherence is established by those studies in which increased incidence of
 6     different health outcomes associated with PM are found in the same population, as is the case
 7     for the following examples, based on currently published studies:
 8          •Detroit:  Mortality mainly in elderly populations, hospital admissions for respiratory
 9                     causes and for cardiovascular causes in the elderly;
10          •Birmingham: Mortality mainly in the elderly, hospital admissions for the elderly;
11          •Philadelphia: Mortality and hospital admissions  for pneumonia in the elderly;
12          •Utah Valley: Mortality and hospital admissions  for respiratory  causes in adults.
13
14     Also, pulmonary function,  respiratory symptoms, and medication use in asthmatic subjects of
15     all ages; hospital admissions for respiratory symptoms, pulmonary function, respiratory
16     symptoms, and medication use in healthy school children, pulmonary function in
17     symptomatic and asymptomatic children; and elementary school absences in children were
18     found to be associatied with PM exposures in Utah Valley.  A similar  study found a PM
19     effect on pulmonary function in smokers with COPD in Salt Lake Valley.  The Utah Valley
20     population was largely non-smoking,  so smoking was not likely to be a source of
21     confounding.
22          While these multiple  outcomes did not occur in strictly identical subgroups of each
23     population, there was probably a sufficient degree of overlap to indicate that PM was a
24     significant predictor of a wide range of health outcomes within a specific community. The
25     symptoms serious enough to warrant hospitalization and the major part of the excess
26     mortality occurred in the elderly sub-group of the population.  However, a significant
27     decrement in pulmonary function and increased incidence of symptoms associated with daily
28     increases in PM occurred in children in Utah Valley, along with a "quality of life" effect
29     measured by lost school days.  Thus, there is evidence for increased risk of health effects
30     related to PM exposure ranging in seriousness from asymptomatic pulmonary function
31     decrements, to respiratory  symptoms  and cardiopulmonary symptoms sufficiently serious to

       April 1995                               1-69       DRAFT-DO NOT  QUOTE OR CITE

-------
 1      warrant hospitalization, and to excess mortality from respiratory and cardiovascular causes,
 2      especially in those older than 65 years of age.
 3          Children may also be at increased risk of pulmonary function changes and increased
 4      incidence of symptoms associated  with PM exposure.  While we have arrayed these health
 5      outcomes in order of increasing severity, there is as yet little indication that there is a
 6      progression of effects in any single individual associated with increasing exposure to PM.
 7      The "exposure-response"  relationship that is derived in most studies must be understood as
 8      characterizing population risk from population exposure.  Additional studies are needed to
 9      define the relationship(s)  among individual exposure to PM and other stress factors,
10      individual risk,  and individual progression among disease states.  Differences in PM
11      dosimetry in the developing, aged, or diseased respiratory tract may also contribute to
12      increased susceptability.
13          The coherence of the various health effects in humans could  be established more
14      conclusively from epidemiology studies if there were better evidence. We cannot prove that
15      the people that suffered respiratory symptoms in response to PM exposure were among the
16      same people who suffered pulmonary function decrements from PM exposure in the past,
17      that those who were admitted to hospital for respiratory or cardiopulmonary causes in
18      response to PM exposure were among those who had suffered  respiratory symptoms or
19      pulmonary function decrements from earlier PM exposures, nor that those who died from
20      PM exposure were among those who had earlier shown other health endpoints associated
21      with PM exposures.  Such information could, in principle, be extracted from longitudinal
22      data bases such as those collected  by health care providers; however, although some such
23      efforts are now being considered,  the preferred design for such a study  is a prospective
24      design rather than a retrospective design.  If and when these studies are completed,  they
25      could be useful in future PM health assessments.
26
27
28      1.13  BIOLOGICAL PLAUSIBILITY:  POTENTIAL  MECHANISMS  OF
29            ACTION
30          Chapter 13, the Integrative Health Synthesis Chapter, incorporates key information  of
31      the types summarized in the several preceeding sections of this chapter.  It also importantly

        April 1995                               1-70       DRAFT-DO NOT QUOTE OR CITE

-------
  1      discusses key points relating to evaluation of the biological plausibility of the new
  2      epidemiologic findings, the  identification of special risk groups, and the interpretation of
  3      implications of reported relative risk estimates for associations between ambient PM exposure
  4      indices and mortality/morbidity effects. Each of these latter three topics are accorded
  5      separate sections in this Executive Summary,  starting with this one on biological plausibility.
  6           Epidemiologic studies  have suggested that  ambient paniculate exposure may be
  7      associated with increased mortality and morbidity at PM concentrations below those
  8      previously thought to affect human health (Section 1.12 and Chapter  12).  However, the
  9      biological plausibility of a causal relationship  between low concentrations of PM and daily
10      mortality and morbidity rates is  neither intuitively obvious nor expected based on
11      experimental studies of the toxicity of inhaled particles.  As indicated in Chapter 11, chronic
12      toxicity from poorly soluble particles has been observed based on the  slow accumulation of
13      large lung burdens of particles, not due to small daily fluctuations of one or another of the
14      specific PM constituents discussed in that chapter.  Two possible exceptions can be noted.
15      Acute toxicity from inhaled particles has been demonstrated with acidic particles, but only at
16      much higher particle concentrations than those observed in the recent  epidemiology studies
17      reporting an association between low-level PM concentrations and morbidity/mortality.
18      Acute toxicity resulting in death has also been reported in rats inhaling singlet ultrafine
19      particles (<0.05 jum) formed in the pyrolysis of perfluorinated compounds at concentrations
20      of 60 to 200 /ig/m3' (Oberdorster et al., 1995; Warheit et al.,  1990), but the significance of
21      these findings for ambient human exposures is yet to be determined.
22           To approach the difficult problem of determining if reported associations between low-
23      level PM concentrations and daily morbidity and mortality are biologically plausible, one
24      must consider:  the chemical and physical characteristics of the particles in the inhaled
25      atmospheres; the characteristics of the morbidity/mortality observed and the affected
26      population; as well  as potential mechanisms that might link the two.  Several salient
27      considerations related to the evaluation of biological plausibility of the epidemiology findings
28      are discussed below.
29
30
        April 1995                                1_71       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Characteristics of Observed Morbidity and Mortality
 2           If daily mortality rates are increased in association with elevated ambient paniculate
 3      concentrations, what are the people dying of? Schwartz (1994) addressed this question by
 4      comparing causes of death in Philadelphia on high pollution days (average = 141 /*g/m3)
 5      with causes of deaths on lower pollution days ( average = 47 /^g/m3).  On the high pollution
 6      days there was a higher relative increase in deaths due to  chronic obstructive pulmonary
 7      disease (COPD) (RR = 1.25), pneumonia (RR = 1.13), cardiovascular disease (RR = 1.09)
 8      and stroke (RR  =  1.15).  There was also an increase in reports  of respiratory factors being
 9      contributing causes in the  deaths and a higher relative age of those dying.  The patterns of
10      causes of death and age of those dying were found to be similar to the patterns observed in
11      the London smog deaths of 1952.
12           Other studies on associations of morbidity with paniculate pollution noted small
13      decreases (2 to 2.5%) in pulmonary function (FVC or FEVj) in  smokers on high pollution
14      days (100 /ig/m3; Salt Lake City; Pope and Kanner, 1993) and in nonsmokers (>60 /zg/m3;
15      NHANES I data, Chestnut et al., 1991).  An increased  number of asthma attacks among
16      working age adults was correlated with increases in  paniculate pollution over a  3-year period
17      (average particle level  = 76 jug/m3) in Helsinki  (Ponka, 1991). Thus, the characteristics of
18      health effects on high particle pollution days are mainly cardiopulmonary in  nature and are
19      the types of effects that can be considered plausibly  related to airborne toxicants.
20           It is also of interest to consider the health status of the people  affected.  People with
21      previously existing health  conditions  (such as COPD, asthma, or other chronic debilitating
22      conditions) are logically likely to be more susceptible to effects from exposure to paniculate
23      pollutants than would be healthy persons. Such a situation might result in an increased daily
24      mortality rate  on days with higher PM10, followed by a decreased daily mortality rate so that
25      the average mortality rate  over a longer time period would not be affected.
26           Data on the relative effect of particle exposures in persons  with pre-existing pulmonary
27      disease compared to healthy persons  do not yield a clear picture. Pope and Kanner (1°93)
28      reported an approximate 2% decline  in FEVj in smokers with mild  to moderate COPD
29      during an increased concentration in  PM10 of 100 fig/m3 in Salt  Lake City.  However,
30      persons with severe COPD (average  FEVj equal to 50% of predicted) had no further
31      reduction in pulmonary function upon acute (2 h) exposure to 90 fig/m3 H2SO4 in clinical

        April 1995                                1-72       DRAFT-DO NOT QUOTE OR CITE

-------
  1      studies (Morrow et al., 1994). Exercising asthmatics experienced mild bronchoconstriction
  2      following the same exposures. In a separate study, exercising adolescent asthmatics exposed
  3      to 68 £tg/m3 H2SO4 experienced reduced pulmonary function (average of 6% decrease in
  4      FEVt)  (Koenig et al., 1989),  but in another study, exercising asthmatics did not respond to
  5      exposures to as high  as 130 jug/m3  H2SO4 (Avol et al., 1990). Using an elastase-induced rat
  6      model of emphysema, Mauderly et al. (1990) found that exposure to diesel exhaust, which
  7      contains aggregates of ultrafine soot particles, resulted in less particle deposition in the lungs
  8      of emphysematous rats than in normal rats, thus sparing the emphyseniatous rats the health
  9      effects  induced by the soot particles in normal animals.
 10
 11      Influence of Particle  Size,  Chemical Composition, and Respiratory Tract
 12      Deposition/Clearance
 13           The PM10 standard is the only U.S. national ambient air quality standard that is not
 14      chemical-specific.  The chemical composition of a particle will greatly affect its toxicity and,
 15      if possible,  should be considered in determining if the observed associations between
 16      atmospheric PM concentrations and increases in morbidity /mortality are causal. For
 17      example, alpha-quartz particles are  more toxic than TiO2 particles (Driscoll and Maurer,
 18      1991); and acid sulfate aerosols are more likely to cause acute health effects than are neutral
 19      sulfate aerosols (Fine et al., 1987).
 20           Size is also important in defining the  toxicity of particles.  Recent studies indicate that
 21      ultrafine particles (<20 nm) are much more toxic than larger inhalable particles (Oberdorster
 22      et al., 1992; Driscoll and  Maurer, 1991).  The ultrafine particles have a greater number and
 23      surface area per unit  mass than fine or coarse particles, which may account, in part, for their
 24      greater toxicity.  Fine particles tend to have a different chemical composition than larger
 25      particles, because their source is often combustion processes.  A study of the chemical
 26      composition of PM2 5 particles versus PM10 particles in Los Angeles indicated that nitrates,
27      sulfates, ammonium and organic and elemental carbon were the most abundant species in the
28      PM2 5 fraction, while the coarser particles contained soil-related species, such as aluminum,
29      silicon, calcium, and  iron (Chow et al., 1994).  Chemical composition of PM10 is discussed
30      in Chapter 3 and summarized  earlier in this chapter (Section 13.2).
       April 1995                                1_73       DRAFT-DO NOT QUOTE OR CITE

-------
 1           In a few epidemiology studies, the investigators attempted to determine what size
 2      and/or chemical form of particles had the strongest association with health effects.  For
 3      example, in the Harvard 6-cities study (Dockery et al.,  1993), the excess chronic mortality
 4      was most strongly associated with the ambient fine particles, including sulfates.  However, in
 5      a study of daily air pollution in St. Louis and eastern Tennessee by Dockery et al. (1992),
 6      the strongest association of particulate pollution with daily mortality rates was PM10, with
 7      progressively weaker associations with PM2 5, sulfate, and aerosol acidity.  This is the
 8      opposite of what one would expect if aerosol acidity were the main cause of increased
 9      mortality, as has been suggested (Lippmann,  1989). The Six Cities study investigators state,
10      however, that the low daily death counts,  the short study period,  and  the large geographic
11      areas considered in the St.  Louis/Eastern Tennessee study limited the  statistical power of the
12      study, and they could not conclude that the acidity of the aerosol was not associated with
13      mortality.
14           If the chemical and physical forms of the PM are important in determining the health
15      effects induced  by PM, one would expect different concentration-response curves to be
16      observed in different epidemiology studies, depending on the type of aerosol present in the
17      atmosphere. Spurny (1993) in his analysis of studies conducted in the south-western part of
18      Germany found differences in concentrations, composition, and cell-toxic effects among
19      urban, residential, and remote areas.  The different slopes of the  concentration-response
20      curves for the different cities could be due to several  factors, including differences in
21      physicochemical properties and resultant potency of the  PM in the different cities.
22           It is also worth noting that considerations of dosimetry could potentially provide insight
23      on plausible mechanisms or alter the exposure-dose-response relationships evaluated.  To
24      date, most analyses have used the exposure concentration (/xg/m3) of particles. Because
25      deposition of particles in the respiratory tract are determined by particle  diameter and
26      distribution, calculation  of the RR estimates based  on various internal dose metrics (e.g.,
27      deposited dose (mass) rate per tracheobronchial or alveolar surface area  or deposited particle
28      number rate per surface area), could alter some of these relationships. Different dose
29      metrics may be more appropriate to characterize acute effects (e.g., mortality) versus chronic
30      effects (e.g., morbidity). Certainly dosimetry can provide insight on  the variability of
31      inhaled dose due to differences in airway morphometry  and  ventilation rates among species,

        April  1995                                1-74       DRAFT-DO NOT QUOTE OR  CITE

-------
  1      age, genders, and disease status of the respiratory tract.  For example, it has been shown that
  2      patients with COPD have increased deposited particle burdens when compared to healthy
  3      subjects (Anderson et al., 1992).  To the extent that particle composition alters the particle
  4      diameter and distribution of a given aerosol, dosimetry will also be effected. Solubility of an
  5      aerosol influences clearance rates and subsequent retained dose estimates.  The potential for
  6      dosimetry to influence the exposure-dose-relationship should be considered to the extent that
  7      mathematical modeling andd norphometric data allow.
  8
  9      Potential Mechanisms  of Causality Between Low Levels of Particulate
 10      Pollution and Health Effects
 11           Pathophysiologic mechanisms by which various specific PM constituents can cause
 12      health effects are discussed  in detail in Chapter 11.  Here, the focus is on mechanisms by
 13      which airborne particles are known to cause health effects and the extent to which such
 14      mechanisms provide plausible evidence or explanations for the reported epidemiologic
 15      findings of increases  in morbidity and daily mortality rates at low PM concentrations.  For
 16      purposes of this discussion, health effects of particle inhalation are discussed below in terms
 17      of:  clinical considerations,  acute lung injury, chronic pulmonary toxicity from accumulation
 18      of particles in the lung, effects on pulmonary function, effects on pulmonary defense
 19      mechanisms, and pathophysicologic mechanisms.  Also considered in this section is the
20      potential for interactive mechanisms among air pollutants that might influence the health
21      effects induced by airborne  particles. This is an area in which there is little information;
22      most studies have been directed toward determining the toxicity of single compounds.
23
24      Clinical Considerations
25           Potential mechanisms  which might help explain the phenomenon of particle related
26      mortality have been considered by Frampton and Utell (1995).  These mechanisms include:
27      (1) "premature" death,  that is the hastening of death for individuals already  near death (i.e.,
28      hastening of an already certain death by  hours or days); (2) increased susceptibility to
29      infectious disease; and (3) exacerbation of chronic underlying cardiac or pulmonary disease.
30           Particulate pollution could contribute to daily mortality rates by affecting those at
31      greatest risk of dying, i.e.,  those individuals for whom death is already  very imminent.

        April 1995                               1-75      DRAFT-DO NOT QUOTE OR CITE

-------
 1      Acute exposure to moderately elevated concentrations which might only be a minor irritant to
 2      healthy people could be the "last straw" that tips over the precariously balanced physiology
 3      of a dying patient.  Other studies suggest that the full effect of particles on mortality cannot
 4      be explained solely by acute PM exposure death-bed effects (Frampton and Utell, 1995), i.e.,
 5      some studies also indicate an effect on annual mortality  rates which cannot be explained
 6      simply by the hastening of death for individuals already near death.
 7           Particle exposure could also increase susceptibility to respiratory infection with bacteria
 8      or viruses, leading to an increased incidence of (and death from) pneumonia in susceptable
 9      members of the population.  However, pneumonia rarely results in death within 24 h of
10      onset; serious infections of the lower respiratory tract generally develop and evolve over
11      weeks, and would not explain effects on daily mortality. If pollutant exposure increased
12      susceptibility  to infectious disease, it should be  possible to detect differences in the incidence
13      of such diseases in communities with low vs. high particle concentrations.  Emergency room
14      visits and hospitalizations for pneumonia caused by the relevant agent should also be
15      measurably higher on days with elevated ambient particle concentrations.  However, no such
16      relationship has been observed, and laboratory  animal data to support such a mechanism are
17      weak.
18           Particulate air pollution might also aggravate the severity of underlying chronic lung
19      disease.  This mechanism could explain increases in daily mortality (through effects on those
20      near death from their disease) and  longitudinal  increases in mortality (if individuals with
21      chronic airways disease experienced more frequent or severe exacerbation of their disease, or
22      more rapid loss of function as a result of paniculate exposure).
23           What chronic disease processes are most likely to be affected by inhaled paniculate
24      matter?  To explain the daily mortality statistics, there must be common conditions that
25      contribute significantly to overall mortality from respiratory causes.  The most likely
26      candidates are the chronic airways diseases, particularly chronic  obstructive pulmonary
27      disease (COPD).  This group of diseases encompasses both emphysema and chronic
28      bronchitis, however, information on death certificates does not allow differentiation between
29      these diagnoses.   The pathophysiology includes  chronic  inflammation of the distal airways  as
30      well as destruction of the lung parenchyma. There is loss of supportive elastic tissue,  so that
31      the airways collapse more easily during expiration, obstructing outflow of air.  Processes that

        April 1995                                1-76       DRAFT-DO NOT QUOTE OR CITE

-------
  1     enhance airway inflammation or edema, increase smooth muscle contraction in the
  2     conducting airways, or slow  mucociliary clearance could adversely affect gas exchange and
  3     host defenses.  Moreover, the uneven ventilation-perfusion matching characteristics of this
  4     disease, with dependence on  fewer functioning airways and alveoli for gas exchange,  means
  5     inhaled particles may be directed to the few remaining functioning lung units in higher
  6     concentration than in normal  lungs (Bates,  1992)
  7          Particulate pollutants have been associated  with increases in cardiovascular mortality
  8     both in the major air pollution episodes and in the more recent time-series analyses.  Bates
  9     (1992) has postulated three ways  in which pollutants could affect cardiovascular mortality
 10     statistics.  These include: (1) acute airways disease misdiagnosed as pulmonary edema;
 11     (2) increased lung permeability, leading to pulmonary edema in people with underlying heart
 12     disease and increased left atrial pressure and (3) acute bronchiolitis or pneumonia induced by
 13     air pollutants precipitating congestive heart failure in those with pre-existing heart disease.
 14     Moreover, the pathophysiology of many lung diseases is closely  intertwined with cardiac
 15     function.  For example, one postulated cause of the increasing mortality rate in asthma is
 16     overuse of adrenergic agonist medications leading to fatal  cardiac arrhythmias.  Many
 17     individuals with COPD also have cardiovascular disease caused by smoking, aging, or
 18     pulmonary hypertension accompanying COPD.  Terminal  events in patients with end-stage
 19     COPD are often cardiac complications, and  may therefore be misclassified as cardiovascular
 20     deaths.  Hypoxemia associated with abnormal  gas exchange can precipitate cardiac
 21      arrhythmias and sudden death.
 22
 23      Acute Lung Injury
 24           The acute toxicity of particles in the respiratory tract has been the topic of numerous
 25      studies to determine the potential pulmonary toxicity of dusts, particularly those of concern in
 26      industrial processes.  Toxic particles  that deposit in the lung can induce an inflammatory
 27      response that, if it persists,  may lead to pulmonary fibrosis and impaired pulmonary function.
28      The response  of the respiratory  tract to such particles includes the release of numerous
29      cytokines from alveolar macrophages and epithelial lining  cells that promote  healing and
30      repair or, if healing does not  occur because of the persistence of toxic particles, may
31      promote development of fibrosis.  Although such acute responses are well known, they

        April 1995                                ^77      DRAFT-DO NOT QUOTE OR CITE

-------
 1      typically only occur after several days or weeks of exposure to airborne particle
 2      concentrations many fold higher than those that have been shown to be associated with
 3      increased mortality and morbidity in epidemiology studies.  Recently, however, it has been
 4      observed  in experimental animal studies that certain types of particles are acutely toxic to the
 5      lung at low exposure concentrations.  A half-hour exposure of rats to freshly generated
 6      ultrafine polytetrafluoroethylene particles at a concentration of 64 ^g/m3 resulted in severe
 7      pulmonary inflammation and death (Oberdorster et al, 1995).  Warheit et al. (1990) also
 8      found that fresh ultrafine aerosols resulted in mortality in rats by causing severe lung injury.
 9      The significance for environmental exposures of the highly toxic fresh aerosols formed from
10      pyrolysis  of perfluorinated materials is unknown at this time, because of the rapid loss of
11      toxicity of the aerosols with time and the lack  of  information on the concentration of those
12      specific aerosols in the ambient atmosphere. Although it is known that combustion processes
13      emit ultrafine aerosols into the environment (Cantrell and Whitby, 1978), it is not clear how
14      much ultrafine particulate matter  is present as the product of pyrolysis of perfluorinated
15      compounds.  Nor is there much information on typical ambient concentrations of other
16      ultrafine particles (e.g., metals from high temperature smelting)  or their persistance as
17      ultrafines in urban aerosol mixes.
18
19      Toxicity Resulting from Accumulation of Particles in the Lung
20           The accumulation of large lung burdens of poorly soluble particles can lead to
21      decreased clearance of subsequently  inhaled particles and an enhanced rate  of accumulation
22      of particles in the lung (Morrow, 1992).  Large lung burdens of particles of even relatively
23      low inherent toxicity have been shown to induce lung cancer in animal models such as the rat
24      (Mauderly et al., 1994).  But how much  prior exposure to particles is required to accumulate
25      enough particles to impair clearance of subsequently inhaled particles? Rats exposed to 350
26      Mg/m3 diesel soot (aggregated ultrafine carbon particles) for 24 months did  not accumulate
27      enough particles to induce pulmonary inflammation (as measured by both histopathology and
28      analysis of lung lavage fluid) or to impair particle clearance, but rats exposed to 3500 ^g/m3
29      for the same length of time did.  Rats that inhaled carbon black  particles at an 8-h
30      time-weighted concentration of 10,000 jig/m3 5 days a week for  12 weeks also accumulated
31      enough particles to induce an inflammatory response by 6 weeks (Henderson et al., 1992).

        April 1995                               1-78      DRAFT-DO NOT  QUOTE OR CITE

-------
  1           In general, the toxicity resulting from accumulation of large burdens of particles in the
  2      lung does not likely provide a plausible biological basis for reported associations between
  3      acute exposures to low level PM concentrations (ca, 30 to 200 Mg/m3) of inhalable particles
  4      (PM10) and daily mortality/morbidity rates.  One possible exception that stands out as a
  5      relatively sizeable segment of the general population would be smokers or former smokers
  6      among the elderly.  In such individuals, particle overload from 40 to  60 years of directly
  7      inhaled tobacco smoke particles could make them more vulnerable to  the impacts of
  8      relatively small additional acute increments in their lung particle burdens, as would the
  9      preexisting chronic  cardiorespiratory diseases caused by smoking.
 10           A second possible exception might be elderly persons who experienced notable past
 11      exposures over many years to very high ambient or workplace PM concentrations, as would
 12      be the  case for individuals who resided  or worked in heavily industrialized cities before
 13      effective occupational and air pollution control measures were introduced in the  1950s to
 14      1970s to reduce such exposures.  For example, in the Harvard 6-cities study, an association
 15      was found between  daily mortality rates and PM levels across a  few rural communities,
 16      lightly  industrialized cities, and some heavily industrialized cities.  Because the ranking of
 17      the cities in terms of air-pollution levels did not change during the study period, it is  not
 18      possible to distinguish completely between effects due to past historical exposures and those
 19      due to  recent exposures.  Therefore, the elevation in daily mortality rates in industrialized
 20      cities such as Steubenville compared to less industrialized cities (such as  Topeka or Portage)
 21      may  be in part  based on accumulated past exposures to higher particle levels and
 22      consequently  larger  lung particle burdens in the former.
 23
 24      Impaired Respiratory Function
 25           Very few  of the specific PM  constituents discussed in Chapter 11 have acute exposure
 26      effects  on respiratory function, except possibly at very high concentrations (in the /ng/m3
27      range).  One  possible exception is acid aerosols, which appear to have acute effects on
28      pulmonary function  among some sensitive individuals at levels below  1,000 Mg/m3.
29      Exposures to  acid particles are known to induce hyperreactive airways and in some cases,
30      bronchoconstriction, but at concentrations in the mg/m3  range, well above peak U.S. ambient
31      acidity  levels  of 50  to 75 /xg/m3. In  healthy humans, inhalation of 1,000 Mg/m3  H2SO4

        April 1995                                1_79      DRAFT-DO NOT QUOTE OR CITE

-------
 1     aerosol for 3 h did not cause any influx of inflammatory cells into the lung based on analysis
 2     of lung lavage fluid obtained 18 h after the exposures (Frampton et al., 1992). However,
 3     mild bronchoconstriction has been reported after brief exposures to as low as 68 jug/m3
 4     H2SO4 in exercising adolescent asthmatics and 90 /ig/m3 in excersing adult asthmatics
 5     (Morrow et al., 1994); (Koenig et al.,  1989), although this has not always been observed
 6     (Avol, et al., 1990).  Also of interest is the finding that hyperresponsive airways developed
 7     after exposure of healthy rabbits to as little as  75 jug/m3 H2SO4 for 3 h (El-Fawal and
 8     Schlesinger, 1994).  Additional studies have also found that acid-coated particles were  more
 9     potent than the acid or particles alone.  Therefore, under  some circumstances, one possible
10     mechanism for increased mortality among some elderly persons with a debilitating disease
11     (asthma) on days with moderately high PM pollution might be that acid aerosols place  a
12     stress on their cardiopulmonary system, leading to death.
13
14     Impaired Pulmonary Defense Mechanisms
15          The ability of paniculate exposures to reduce pulmonary defense mechanisms has been
16     documented for aerosols of H2SO4 and trace metals.  Trace metals have been shown to be
17     cytotoxic to alveolar macrophages (AMs) and immunosuppressive,  but only  at much higher
18     concentrations than encountered in ambient atmospheres (Zelikoff et al., 1993).   Sulfuric
19     acid aerosols have also been shown to alter resistance to bacterial infection in mice after
20     acute exposures to 1,000 fj.g/m3; repeated exposures to 100 /jg/m3 reduced mucociliary
21     transport rates in animals.  Even these  levels of H2SO4 are much higher than have been
22     reported in atmospheres of cities evaluated in the recent epidemiology studies.  Also, one
23     would expect effects from impaired pulmonary defense mechanisms to develop over  an
24     extended period of continuing exposure, not within a few days.
25
26     Synergistic Effects
27          An area for which there is little information is the potential interactive effects of
28     mixtures of air pollutants and/or with other factors (e.g.,  aging).  The potential significance
29     of mixtures is illustrated by the studies of Amdur and Chen (1989), in which a repeated daily
30     3-h exposure for 5 days of guinea pigs to 20 /ng/m3 of H2SO4 coated on metal particles
31     resulted in decrements in lung volume  and pulmonary diffusing capacity and elevations of

       April 1995                                1-80      DRAFT-DO NOT QUOTE  OR CITE

-------
 1     lung weight/body weight ratio, protein, and number of neutrophils in pulmonary lavage fluid.
 2     For example,  A 1-h exposure to 20 /ag/m3  H2SO4 coated on metallic particles increased
 3     bronchial reactivity in guinea pigs; a 10-fold higher concentration of H2SO4 alone was
 4     required to produce the same response (Chen et al., 1992b).  However,  such synergistic
 5     effects were not observed by Anderson et al.  (1992), who studied the effects on 15 healthy
 6     and 15 asthmatic volunteers of 1-hr exposures to 100 /ng/m3 H2SO4 (0.5 /^m) or 250 /-ig/m3
 7     carbon black (0.5 /xm) separately or with the  H2SO4 coated on the particles.  The exposures
 8     did not result  in changes in symptoms  or pulmonary function, except for an equivocal
 9     response in one person.
10           The population segment most susceptible to elevations in ambient PM are the elderly
11     (> 65 years old) with preexisting respiratory disease.  Aging, in the absence of pathology, is
12     an extremely complex biological phenomenon and is described as being  a multifactorial
13     process composed of both genetic and  environmental components (Cristofalo et al., 1994).
14     While the physiological characteristics of the healthy older population is an area of active
15     research, significant decrements in key physiological parameters including lung volumes,
16     FEVj, flow velocity/volume curves,  resting cardiac output, and cardiac  output reserve with
17     age have been reported (Kenney, 1989). However, there is controversy concerning
18     decrements in physiological function associated with the aging process alone as well as with
19     accompanying disease processes or with other environmental stressors.   Moreover, there is
20     little  information on the extent to which an older population might be more susceptible to the
21     effects of ambient paniculate pollution (Cooper et al., 1991).  It is possible that the elderly
22     are more susceptible to ambient particles because of numerous changes in  the body's
23     protective mechanisms and protracted exposures to particles over a life time.  This could
24     allow time for latent effects from earlier life time exposures to manifest themselves, and for
25     potential cumulative effects to emerge.  Virtually nothing is known of the  possibilities for
26     interaction among toxicants over a long life time or the possibilities for interaction between
27     medications and ambient pollutants.
28
29
        April 1995                                1-81       DRAFT-DO NOT QUOTE OR CITE

-------
 1     Pathophysiologic mechanisms
 2           The respiratory system may be compromised and become less efficient in older people
 3     or as a result of disease, and inhaled particles could, conceivably, further compromise their
 4     respiratory function.  Because small increases in environmental particle concentrations would
 5     not be lethal to most people,  the effect must result from initiating or promoting a lethal
 6     failing of a critical function, such as ventilation, gas exchange, pulmonary circulation, or
 7     cardiorespiratory control in subjects brought to  the limits of tolerance by preexisting
 8     conditions (Mauderly, 1995).
 9           Inhaled particles or their pathophysiological reaction products could  further impair
10     ventilation in the chronically  ill individual by further reducing airway caliber.  For example,
11     particles may activate airway smooth muscle, constricting airways,  or may influence various
12     airway secretions which could add to and thicken the mucous blanket.  Inhaled particles or
13     their pathophysiological reaction products could decrease the diffusing capacity of the lungs
14     by decreasing the area of the respiratory membrane available for diffusion, by  increasing
15     diffusion distances across the respiratory membrane, and/or by causing abnormal ventilation-
16     perfusion ratios in some parts of the lung.  Particles or their products could also act at the
17     level of the pulmonary vasculature to elicit changes in pulmonary vasculature resistance,
18     which could further alter ventilation-perfusion abnormalities in people with respiratory
19     disease.  Furthermore, particles  could conceivably alter respiratory and cardiovascular
20     control by affecting local control mechanisms located in the endothelial cells or other sites.
21     This could produce changes in peripheral and central control mechanisms and directly affect
22     the respiratory and cardiovascular control centers. Little evidence is currently  available that
23     directly addresses the above speculative possibilities.
24
25     Biological Plausibility  Conclusions
26           Having considered the characteristics  of the paniculate exposure  atmospheres and the
27     types of morbidity and mortality associated with the polluted atmospheres, what can be
28     concluded about the biological plausibility of the  epidemiological results?  The types of
29     morbidity and mortality reported to be associated with increased  ambient  particle
30     concentrations are  consistent with the types of health effects that one might expect from
31     exposures to high levels of PM.  Therefore, the type of response seems plausible, if one

       April 1995                                 1-82      DRAFT-DO NOT QUOTE OR CITE

-------
 1     accepts the temporal relationships modeled in the epidemiological studies. The analyses
 2     found associations with 1-day or multi-day (usually 3 to 5 day) lags.  The concentrations of
 3     paniculate matter reported to be associated with such health responses, however, are much
 4     lower than would be expected based on animal and human clinical studies of responses to
 5     single particulate pollutants. This is true even when one considers that there is evidence that
 6     the people who make up the excess mortality population may be susceptible  subpopulations.
 7     Moreover,  it is not clear what portion of the inhalable particulate matter constitutes the
 8     delivered dose that is associated with the observed morbidity or mortality. There are
 9     suggestions from both animal toxicology data and epidemiology data that ultrafine acid
10     aerosols may be of greater health significance than the rest of the particulate mass.  Finally,
11     the potential for interactive effects between PM of different types and 'PM and other air
12     pollutants is not known.
13          Thus, although there are several hypotheses as outlined above, little clear or convincing
14     evidence is available at this time to support the biological plausibility of a causal relationship
15     for the reported epidemiologic associations between low ambient concentrations of PM and
16     daily mortality and morbidity rates.
17
18
19     1.14 IDENTIFICATION OF POPULATION  GROUPS POTENTIALLY
20           SUSCEPTIBLE TO HEALTH  EFFECTS FROM PM EXPOSURE
21          Certain groups within the population may be more susceptible to the effects of PM
22     exposure, including persons with preexisting  respiratory disease, children, and the elderly.
23     The reasons for paying special attention to these groups  is that (1) they may be affected by
24     lower levels of PM than other subpopulations and (2) the impact of an effect of given
25     magnitude may be greater.  Some potential causes of heightened susceptibility are better
26     understood than others.  Subpopulations that  already have reduced ventilatory reserves (e.g.,
27     the elderly and persons with asthma, emphysema, and chronic bronchitis) would be expected
28     to be more impacted than other groups by a given decrement in pulmonary function.  For
29     example, a healthy young person may not even notice a small percentage change in
30     pulmonary function, but a person whose activities are already limited by reduced lung
31     function may not have the reserve to compensate for the same percentage change.

       April 1995                               1-83       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Based on Chapter 12 discussions, it is clear that the bulk of the total mortality effects
 2      suggested by the epidemiology studies discussed earlier are among the elderly.  During the
 3      historic London, 1952 pollution episode the greatest increase in the mortality rate was among
 4      older citizens and those having respiratory diseases.  An analysis by Schwartz (1994c) of
 5      mortality in Philadelphia, PA found the greatest increase in risk of death in those aged 65 to
 6      74 and those  >74 year of age (mortality risk ratios =  1.09 and  1.12, respectively, between
 7      high and low TSP days).  Other studies also suggest that the elderly experience a higher
 8      excess risk from exposure to PM air pollution than the  population overall.
 9           Other potentially susceptible groups include patients with COPD, such as emphysema
10      and chronic bronchitis. Some of these patients have airway hyperresponsiveness to physical
11      and chemical stimuli.  A major concern with COPD patients is the absence of an adequate
12      ventilatory reserve, a susceptibility factor described above.  In addition, altered distribution
13      of respiratory tract ventilation in COPD may lead to a greater delivery of PM to the segment
14      of the lung that is well ventilated, thus resulting in a  greater regional tissue dose. Also, PM
15      exposure may alter already impaired defense mechanisms, making this population potentially
16      more susceptible to respiratory infection. It is estimated (U.S. Department of Health and
17      Human Services,  1990; Collins, 1988) that  14 million persons (-6%) suffer from COPD in
18      the United States.  Bronchial  mucous transport clearance may be impaired in people with
19      chronic bronchitis, asthma, and in association with various acute infections.  Rates of
20      alveolar region clearance  appear to be reduced in humans with chronic obstructive lung
21      disease.
22           Throughout the results and discussions presented above and in Chapter 12 regarding the
23      effects of acute  PM exposure on human mortality,  a consistent trend was for the effect
24      estimates to be higher for the respiratory mortality category.  This lends support to the
25      biological plausibility of a PM air pollution effect,  as the breathing of toxic particles would
26      be expected to most directly affect the respiratory tract, and these results are consistent with
27      this expectation.  For example,  the estimates of relative risk for PM-induced mortality due to
28      respiratory causes discussed in Chapter 12 are all higher than the risks for the population as
29      a whole and for other causes.  More specifically, the PM RR for respiratory diseases  ranged
30      from 50 to more than 400% higher for respiratory  disease categories than for all causes of
31      death, indicating that increases in respiratory deaths are a key major contributor to the

        April 1995                                1-84      DRAFT-DO NOT QUOTE OR CITE

-------
  1     overall PM-mortality associations noted previously.  PM relative risk estimates for
  2     cardiovascular causes were also notably elevated.  Moreover,  since evidence suggests that an
  3     acute pollution episode is most likely be inducing its primary effects by stressing already
  4     compromised individuals (rather than, for example, inducing chronic respiratory disease from
  5     a single air pollution exposure episode), the above results indicate  that persons (especially the
  6     elderly) with pre-existing cardiovascular or respiratory disease constitute a population
  7     segment especially at risk for mortality implications of acute ambient PM exposures.
  8          Apropos  to the identification of individuals with pre-existing  respiratory and
  9     cardiovascular diseases as being at special  risk for ambient PM exposure effects, it is
 10     important to highlight smoking as a key etiological  agent for such  diseases. The U.S.
 11     Environmental Protection Agency (1992) report on  environmental tobacco smoke indicates
 12     that smoking is the major cause of chronic obstructive pulmonary disease (COPD), which
 13     includes emphysema, and is thought to be responsible for approximately 61,000 COPD
 14     deaths yearly,  i.e., about 82% of U.S. COPD deaths  (U.S. DHHS, 1989).  Tobacco use is
 15     also a major risk factor for cardiovascular  diseases, the  leading cause of death in the United
 16     States. It is estimated that each year 156,000 heart disease deaths  and 26,000 deaths from
 17     stroke are attributable to smoking (CDC, 1991).  Smoking is also a risk factor for various
 18     respiratory infections, such  as influenza,  bronchitis, and pneumonia. An estimated 20,000
 19     influenza and pneumonia deaths per year are attributable to smoking (CDC, 1991).
 20          The U.S. Environmental Protection Agency report also indicates that in children, ETS
 21      exposure is causally associated with an increased risk  of lower respiratory tract infections
 22      such as bronchitis and pneumonia.  It is estimated that 150,000 to 300,000 cases annually in
 23      infants and young children up to  18 months of age are attributable  to ETS.  ETS exposure is
 24      also causally associated with additional episodes and increased  severity of symptoms in
 25      children with asthma.  It is estimated that 200,000 to 1,000,000 asthmatic children have  their
 26      condition worsened by exposure to ETS.  ETS is also a  risk factor for new cases of asthma
 27      in children who have not previously displayed symptoms (U.S.  Environmental Protection
 28      Agency).
29           Lastly,  the EPA report also indicates that environmental tobacco smoke (ETS) is a
30      human lung carcinogen, responsible for approximately 3,000 lung cancer deaths annually in
31      U.S.  nonsmokers (U.S. Environmental Protection Agency, 1992).

        April 1995                                !_85       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Overall, then, the most susceptible population segment that can be most clearly
 2      identified as being at likely increased risk for low-level ambient PM exposure-induced
 3      mortality or morbidity are elderly individuals with pre-existing cardiovascular respiratory
 4      diseases,  the majority of which are likely either current or former smokers.  Smoking may
 5      also be a key ancillary contributor to any low-level PM exposure-induced exacerbation of
 6      respiratory  infections among other adults and children and to any increased cancer mortality
 7      attributable to chronic ambient PM exposures.
 8           Asthmatic subjects appear to be more sensitive than healthy subjects to the effects of
 9      acid aerosols on lung function, but the effective concentration differs widely  among  studies.
10      Adolescent  asthmatics may be more sensitive than adults, and may experience small
11      decrements in lung function in response  to H2SO4 at exposure levels only slightly above peak
12      ambient levels.  Although the reasons for the inconsistency among studies remain largely
13      unclear, subject selection may be an important factor.  Even in studies reporting an  overall
14      absence of effects on lung function, occasional asthmatic subjects appear to demonstrate
15      clinically important effects. Studies from different laboratories suggest that responsiveness to
16      acid aerosols may correlate with degree  of baseline airway hyperresponsiveness.  On the
17      other hand, based on very limited studies, elderly and individuals with chronic obstructive
18      pulmonary disease do not appear to be particularly susceptible to the effects of acid  aerosols
19      on lung function.
20           Alveolar deposition at different flow rates was  lower (26%  versus 48% thoracic
21      deposition)  in subjects after induced bronchoconstriction.  In asthmatics,  thoracic deposition
22      of particles was higher than healthy subjects  (83% versus 73% of total deposition).
23      Trachial/bronchial deposition was also found to be higher in asthmatics.  The results are
24      similar to those found in subjects with obstructive lung disease.   The buffering capacity of
25      mucus may be altered in persons with compromised lungs.  For example, sputum from
26      asthmatics had a lower pH than that from normals and a reduced buffering capacity, and so
27      may represent a population segment especially sensitive to inhaled acidic particles.
28           The National Institutes of Health (1991) estimates that approximately 10 million persons
29      in the United States have asthma.  In the general population, asthma prevalence rates
30      increased by 29% from 1980 to 1987. For those under 20 years old, asthma rates increased
31      from approximately 35 to 50 per 1,000 persons, a 45% increase.  The airways of asthmatics

        April 1995                                 1-86       DRAFT-DO NOT QUOTE  OR CITE

-------
  1      may be hyperresponsive to a variety of inhaled materials, including pollens, cold-dry air,
  2      allergens, and air pollutants. The potential addition of an PM-induced increase in airway
  3      response to the already heightened responsiveness to other substances  raises the possibility of
  4      exacerbation of this pulmonary disease by PM.
  5
  6
  7      1.15  IMPLICATIONS OF RELATIVE RISK ESTIMATES
  8           Preceding sections of this Chapter concluded that the newly emerging epidemiologic
  9      data base on PM-mortality/morbidity effects provides reasonably consistent results indicative
10      of increased risk of mortality and morbidity effects being associated with exposures  of the
11      general population  to ambient air pollutant mixes containing PM concentrations currently
12      found in many U.S. urban areas.  This  includes effects associated with ambient air exposures
13      to pollutant mixes having 24-h PM10 concentrations falling in the range of 30 Mg/m3 to 200
14      /*g/m3, including evidence suggestive of effects below 150 /*g/m3 (the level of the current 24-
15      h U.S. PM10 NAAQS).
16           It was also  noted in Chapter 12 that the relative risk (RR) estimates for both the
17      mortality and morbidity effects associated with short-term (ca. 24-h or a few days) exposures
18      to ambient PM are very small compared to RR values typically viewed in epidemiologic
19      literature as providing strong evidence for a likely causative association.  Section 1.12 further
20      noted  the relatively limited evidence directly demonstrating coherence  between the mortality
21      and morbidity effects findings from epidemiologic studies, with the most  compelling evidence
22      for coherence now  being findings of both increased hospital  admissions (for cardiopulmonary
23      endpoints) and increased mortality in relation to increments in 24-h PM concentrations in the
24      same population group (the elderly) within several  U.S. urban areas (Detroit, Birmingham,
25      Philadelphia) and the Utah Valley.  However, only very limited evidence for the biological
26      plausibility of acute low-level PM exposure effects at the above-stated PM10 concentration
27      range  now appears  to exist to support several hypotheses discussed in  Section 1.13 with
28      regard to possible mechanisms of action.  A key point emerging from  the plausibility
29      discussion and the ensuing section (1.14) was the identification of elderly individuals (65 yr.
30      old) with preexisting chronic cardiovascular and respiratory disease conditions (the majority
31      of whom are likely current or former smokers) as being the  most susceptible general

        April 1995                                1-87      DRAFT-DO NOT QUOTE OR CITE

-------
 1      population segment most clearly at special risk for mortality and morbidity effects associated
 2      with exposures to ambient air mixes containing moderately elevated PM concentrations.
 3           The  meaning or interpretation of quantitative estimates of PM-related effects (i.e.,
 4      relative risk estimates) discussed earlier as having been generated by the newly available PM
 5      epidemiology studies remains a subject of controversy, with divided opinions still existing in
 6      the scientific community as noted earlier in this chapter.  Thus, in attempting to  interpret
 7      such risk estimates, several caveats should be kept in mind.  First,  caveats analogous to those
 8      made in point  (4) at the top of page  1-33 for key conclusions drawn from the last previous
 9      PM criteria review  still apply.  That is, although new evidence has emerged which points
10      toward very small,  but statistically significant increases in risk  of human mortality and
11      morbidity  effects being associated with exposures to ambient air mixes containing moderately
12      elevated PM (with no evident thresholds being identified in the studied range of PM
13      concentrations), precise quantitative specification of relative contributions of such low-level
14      concentrations of ambient PM to reported mortality and morbidity effects is not possible at
15      this time.  Nor can one now separate out with confidence potential  relative contributions to
16      the reported PM effects of several other likely  important confounding or interacting
17      variables.
18           With regard to the latter, it is as of yet very difficult, for example, to sort out with
19      confidence relative  contributions of weather versus PM per se.   It is clear that temperature
20      extremes (very hot  or very cold days in relation to typical ranges of temperature for any
21      given locale) have notable effects on variations in daily mortality, with temperature or other
22      combinations of variables indexing weather impacts usually being found to be significant
23      predictors  of daily human mortality in modeling of PM effects  and  to account for distinctly
24      larger proportions of the variance in daily mortality than do indices of PM pollution.  On the
25      other hand, in most of the newer PM studies, small elevations in relative risk attributable  to
26      PM still remained even after control for temperature extremes and/or other weather indices;
27      and PM effects were found to be significant in several analyses (e.g., for London) restricted
28      to days not involving wide variations in temperature that would constitute geographic-specific
29      extremes.  It is also not yet clear to what extent any given relative risk estimate derived from
30      any of the newer analyses represent actual risk due to an  increase in ambient PM or to what
31      extent the  elevations in risk attributed to modeled PM indices more broadly represent

        April 1995                                1-88       DRAFT-DO NOT QUOTE OR CITE

-------
  1      increased mortality or morbidity risks due to human exposure to the overall pollutant mix in
  2      the particular airshed evaluated (including not only the ambient PM aerosols present but
  3      other copollutants, such as SO2, CO, O3, NOX or non-particulate organic air toxics).
  4           Other caveats bear on the issue of how generalizable the  reported PM relative risk
  5      estimates are.  It is not yet possible to determine the extent to  which the risk estimates for
  6      PM mortality or morbidity effects are generalizable to other geographic areas or are highly
  7      site-specific, i.e., narrowly applicable to the specific cities from which they were derived or,
  8      at  least, most credibly confined for use in projecting any estimates of likely PM risk to other
  9      airsheds with fairly similar ambient aerosol mixes in terms of particle size distribution and
10      chemical composition.  Thus, it is not clear, for example, how credible the use of PM-
11      related relative risk estimates derived from Philadelphia, St. Louis, or other midwestern or
12      eastern U.S. conurbanations  (or foreign cities such as Sao Paulo, Santiago, or  Athens) with
13      high percentages of particles  from combustion processes might be in attempting to estimate
14      PM-related risks for other cities, e.g., in the western U.S., with much greater proportions of
15      crustal materials in the ambient air pollutant mix.  Use of presently available PM-related risk
16      estimates to attempt to quantify potential PM-related risks across various seasons in locales
17      where widely varying seasonal mixes of particles of different sizes/chemical composition may
18      also be of dubious scientific credibility at this time.
19           Another issue of much  interest and debate has been that of "threshold" for the estimated
20      PM effects derived from the  newly available analyses. As noted earlier and discussed  in
21      Chapter 12, no evident thresholds have yet been demonstrated  for reported PM-related
22      mortality or morbidity effects, based on the presently available published analyses.  On the
23      other hand, as also discussed in Chapter 12, only very limited  efforts have been made to date
24      to  undertake statistical analyses by which to more definitively address the issue;  and serious
25      doubt exists as to whether any thresholds, if they do exist, even in the range of the  observed
26      data (i.e., roughly from 30 to 200 /ug/m3 PM10) can be demonstrated, given notable
27      statistical power limitations associated with necessary breaking  down of data into more
28      refined intervals as part of any threshold "search".  Nor is there now any scientifically
29      credible basis by which to make a "no-threshold" argument in  support of extrapolating
30      currently available PM relative risk estimates to ambient PM concentrations below the  range
31      of  observed data used in the reported analyses. This is especially true in view of the lack of

        April 1995                                 1-89       DRAFT-DO NOT QUOTE  OR CITE

-------
 1      any well demonstrated evidence for one or another hypothesized potential mechanisms of
 2      action that might plausibly explain the elevated risk of mortality or morbidity at the very low
 3      PM concentrations implied by the results of the newly available epidemiology studies.
 4           It is also clear from the available analyses that the occurrence of any increased risk of
 5      mortality or morbidity due to short-term moderate elevations in PM (either alone or in
 6      concert with other copollutants) likely represents the outcome of a combination of risk factors
 7      culminating in relatively rare health events (as clarified further by the ensuing quantitative
 8      discussion below). By far the greatest risk is posed for the elderly over 65 years old and
 9      especially those with preexisting cardiopulmonary diseases, with very distinctively lower risk
10      estimates having been derived for younger individuals and those without chronic respiratory
11      or cardiovascular  diseases.  Thus, in order for notable health effects  to occur in association
12      with short-term exposures to ambient PM (and/or copollutants), it appears that other
13      predisposing conditions and/or contributing risk factors must be present, as well.  That is,
14      low-level ambient PM exposures alone do not typically appear to be  sufficient per se to
15      induce increased morbidity or mortality, but may contribute to such health outcomes under
16      conditions when one  or more other contributing risk factors also converge.  Thus, for
17      example, short-term low-level exposures to ambient PM at concentrations in the ranges
18      evaluated in most of the newer epidemiology studies are extremely unlikely alone to cause
19      lung function decrements or  respiratory symptoms of much note (except possibly for some
20      highly sensitive asthmatic patients), based on currently available epidemiologic and controlled
21      human exposure study results. Nor are such exposures likely to markedly reduce or impair
22      respiratory tract defenses (e.g., alveolar macrophage numbers or function, retrociliary
23      clearance mechanisms, lung  immune response, etc.) sufficiently so as to cause increased
24      susceptibility  to respiratory infections, based on available experimental toxicology findings.
25      On the other hand, once a respiratory infection were to occur due to other causes, then  it is
26      conceivable that added stress due to low-level PM exposure in terms of small further
27      decrements in pulmonary function or exacerbation of respiratory symptoms could lead to
28      worsening of the acute illness and, possibly, the need  for medical attention and/or hospital
29      admission in some cases.
30           Still additional converging risk factors appear to be necessary for exposures  to ambient
31      air pollution mixes containing low  concentrations of typical outdoor urban aerosols to

        April 1995                                1-90       DRAFT-DO NOT QUOTE OR CITE

-------
  1      contribute to increased mortality. By far the most important are the cooccurrence of
  2      advanced age (>  65 yr old) and already compromised cardiopulmonary function.  In older
  3      individuals with preexisting COPD, emphysema, chronic heart disease, etc. resulting from
  4      other predisposing risk factors (e.g., long-term earlier high-level particle exposures from
  5      smoking or past occupational or ambient PM exposures before effective control measures
  6      were introduced), it appears conceivable that additional stress from low-level ambient PM
  7      exposures  might cause further complications that might lead to terminal consequences in
  8      some cases.  Several possibilities were discussed earlier as  having been hypothesized, e.g.,
  9      increased air flow to and consequent greater particle deposition/retention in remaining
 10      functioning areas of the compromised lung, possible tipping over by small  additional particle
 11      burdens of already saturated lung defenses due to particle overloads from past long-term high
 12      level particle exposures, and/or the induction of cascading inflammatory or other immune
 13      responses (due to  particularly toxic specific PM constituents e.g., possibly  certain transition
 14      metals) that overwhelm remaining lung reserves and oxygen exchange mechanisms.
 15      However,  at this time, no clearly convincing  scientific evidence has yet been reported by
 16      which to either compellingly substantiate or refute such hypothesized possibilities.  Thus,
 17      considerable uncertainties still exist with regard to what the relative risk estimates from the
 18      newly  available epidemiologic studies may imply.
 19          In evaluating the potential public health  significance of the relative risk  increases for
 20      mortality or morbidity effects reported in the  newer PM epidemiology studies, much recent
 21       interest has focussed on use of such relative risk estimates to generate projections of numbers
 22      of excess deaths or morbidity events likely to be associated with ambient PM exposures at
 23      concentrations currently found in the United States or other countries.  Given the above-
 24      noted caveats and uncertainties pertaining both to the derivation of the relative risk estimates
 25      and their interpretation, there are substantial reasons to caution against attempting such
 26      calculations at this time and to have  major reservations about accepting any such projections
 27      as credible quantitative estimates of additional deaths or morbidity events likely to actually
 28      occur with current or future exposures in the United States or elsewhere.  At best, such
 29      projections might be associated with exposures to PM-containing  ambient air mixes in cities
 30      with closely similar particle size/chemical composition characteristics and population
31      demographics to those cities from which the relative risk estimates  were derived.  It is

        April 1995                                j_91       DRAFT-DO NOT QUOTE OR CITE

-------
 1      currently questionable as to whether widely generalizable, broadly applicable projections can
 2      be made based on some single "best estimate" of PM-related relative risk and, also, whether
 3      such projections can be credibly aggregated across PM exposure variations during different
 4      seasons and/or across geographic locales with widely disparate mixes of PM aerosols and/or
 5      other copollutants.
 6          Despite the above caveats and reservations, however, it may be useful  to provide
 7      illustrative examples of possible quantitative implications of relative risk estimates of the type
 8      generated by the recent PM epidemiology analyses. Table 1-5 earlier showed that total acute
 9      mortality relative risk estimates associated with exposure to ambient air pollution having a 50
10      Mg/m3 increase in one-day 24-h average PM10 can range from  1.015 to 1.085, depending
11      upon the site (i.e., the PM10 composition and population demographics) and  also upon
12      whether PM10 is modeled as the sole index of air pollution or not. Relative  Risk  estimates
13      with PM10 as the only pollutant index in the model range from RR = 1.025 to 1.085, while
14      the PM10 RR with multiple pollutants in the model range from 1.015 to 0.025.  The former
15      range, as noted earlier, might be viewed as approximating an upper bound of the best
16      estimate, as any mortality effects of co-varying pollutants may be "picked up" by the PM10
17      index, whereas the latter multiple pollutant model range might be viewed as  approximating a
18      lower bound of the best estimate, as the  inclusion of highly correlated covariates might
19      weaken the PM10 estimate.  Thus, "typical"  total mortality effect estimates (per 50 /xg/m3
20      PMio increase) most likely fall within an approximate RR = 1.025 to 1.06 range, based on
21      the various coefficients reported in the published studies.  Formal EPA meta-analyses results
22      discussed in Chapter 12 yielded a best estimate of 1.031 with 95% confidence intervals (CI)
23      of 1.025 to 1.038 for PM10 studies using models that include 0-1 day lags but no
24      copollutants. For those analyses  with longer lag times (3-5 days) and no copollutants in the
25      models,  the EPA meta-analyses yielded a best estimate of 1.064 (CI = 1.047 to 1.082).
26      Thus, the very small increased risks of about 3.1 to 6.4% over baseline mortality levels (per
27      50 Mg/m3 increase in 24-h PM10 concentration in the 30 to 200 /xg/m3 range) derived from
28      the EPA meta-analyses probably represent currently best available upper bound estimates for
29      reported PM10-related total mortality effects. Lower bound estimates, from analyses that
30      included other copollutants in the models for acute PM-mortality effects,  could be as much as
        April 1995                               1-92       DRAFT-DO NOT QUOTE OR CITE

-------
  1      50% lower than the above upper bound estimates or, possibly, even include zero (i.e.,
  2      represent no increased risk) especially during some seasons in different locales.
  3           To help place these findings into a context by which to better understand the potential
  4      implications of such relative risk estimates, Table 1-14 summarizes important information by
  5      which to project potential increases  in excess mortality in a city of one million people that
  6      might be associated with exposure to ambient air mixes in which an increment in 24-h PM10
  7      exposure of 50 /zg/m3 may be a contributing factor.  First it is important to note that the
  8      typical general population baseline or background risk of death on any given day in the U.S.
  9      is only about 23.6 in a million (23.6 x 10"6) or 23.6 deaths per day in a city of 1 million
 10      people.  If the 24-h PM10 concentration increased by 50 /xg/m3 on a given day (e.g., from a
 11      usual level of about 50 /ug/m3 to around 100 /xg/m3) then risk for mortality in the total
 12      general population would be expected to increase by about 3.0 to 6.0% over baseline, i.e., to
 13      increase from 23.6 in a  million to about 24 or 25 in a million as an upper bound estimate.
 14      In other words, exposure to the ambient mix of pollutants indexed by  the 50 jwg/m3 increase
 15      in 24-h PM10 levels might contribute to as much as an additional 0.7 to  1.5 deaths per
 16      million people exposed,  as shown in the Table 1-14 far right column.
 17           Of the 23.6 baseline deaths per day expected in a city of 1  million, about  17 would be
 18      attributable to elderly individuals (aged 65 or over), who only constitute about 12.6% of the
 19      1991 U.S. population but for whom the background risk of dying on a given day is much
20      higher than for the total general population.  For such individuals, the upper bound estimate
21      for increased numbers of excess deaths possibly contributed to by the 50 /jg/m3 PM10
22      increase would be projected to be approximately 1.0 (more than half of the higher estimate
23      for total mortality among the  entire population), assuming that the specific city has a typical
24      demographic distribution of percentages of people in different age brackets.  In other cities
25      or locations with notably higher elderly populations (e.g., some  retirement communities or
26      cities left with higher percentages of the elderly possibly due to  outmigration of younger
27      people), then the overall risk  and expected deaths per day would be higher.  Conversely, in
28      other locations with much younger than average populations and lower percentages of elderly
29      residents, the risk and expected numbers of PM-related excess deaths would be lower.
30           If the increment in PM10 concentration continued  to average about 50 /xg/m above
31      routine ambient levels for 3-5 days in the given city of 1.0 million people, then relative risk

        April 1995                               1.93       DRAFT-DO NOT QUOTE OR CITE

-------
 1     estimates derived from PM]0-mortality models using 3-5 d lags might more appropriately
 2     apply.  Then, the expected number of deaths to which the 3-5 day 50 pig/m3 PM elevation
 3     might be projected to contribute could range up to about 1.5 deaths per day among the total
 4     general population per 1 million people exposed;  or up to about 4.5 to 7.5 deaths during the
 5     full 3-5 days of elevated PM10.  Of the 1.50 excess death per  day attributed to the 3-5 day 50
 6     /xg/m3-increment in PM10 24-h concentrations, an estimated 0.34 would likely be due to
 7     respiratory causes and about 0.91 to cardiovascular causes.  Obviously, both the increased
 8     deaths due to respiratory and cardiovascular causes would mainly occur in elderly persons
 9     having preexisting  chronic respiratory or cardiovascular disease conditions.  Note that small
10     numerical inconsistencies in Table 1-14 and in succeeding tables on morbidity arise from the
11     fact that the excess risk estimates are based on different studies in a number of different
12     populations,  with different baseline death or hospital admissions rates.
13           Table 1-15 simply takes the information from the far right column of Table 1-14 on
14     upper bound estimates of the number of possible  PM-contributory deaths per day (for the
15     total population, for the elderly 65 + , and for respiratory and cardiovascular causes),  and
16     depicts  ranges of lower and upper bound estimates for comparable numbers of estimated
17     possible deaths per day contributed to by exposure to ambient air pollution mixes containing
18     50 jug/m3 increments in PM10 concentrations in cities ranging  from 10 thousand to 10 million
19     in size.  Table 1-14 is extremely informative in showing that no appreciable risk for
20     mortality is expected to occur with exposure to such ambient air mixes for cities less  than 1
21     million population, even if the PM10 elevation lasts for 3-5 days or occurs several times  a
22     year; nor is there much appreciable risk for the elderly in smaller population cities, unless
23     perhaps  a particular city with less  than  1 million  population has an extraordinarily high
24     percentage of elderly residents.  This applies even for days when 100 /xg/m3-increments  in
25     PM10 might occur  for 3-5 days in a row.  Even for cities of 1 million population, the
26     projected upper bound risk may be of dubious public health significance unless 50-100 /xg/m3
27     PM10 elevations were to occur numerous times per year, especially in view of such tiny
28     increased risk likely mainly being posed for elderly individuals with preexisting
29     cardiopulmonary disease conditions that predominantly arise from voluntary smoking. Any
30     risk of excess mortality associated with short-term, acute exposures to ambient air pollutant
        April  1995                                1-94       DRAFT-DO NOT QUOTE OR CITE

-------
 TABLE 1-14.  ESTIMATED EXCESS MORTALITY PER DAY IN A POPULATION
     OF ONE MILLION FOR WHICH AN INCREASE OF 50 jtg/m3 PM10 (24-h)
                    COULD BE A CONTRIBUTING FACTOR


Health
Outcome
Total
Mortality
Total
Mortality
Respiratory
Mortality
Cardiovascular
Mortality


Age
Group
All
65 +
All

All

All

Population
Baseline
Annual
Mortality
S^OB1
6,2013
8,603

6761

3,635

Population
Baseline
Daily
Mortality
23.6
17.0
23.6

1.85

10.0


PM10
Lag
Time
< 2d
2d
3-5d

3-5d

3-5d

Upper Bound
Excess Risk
Per PM,r,
50 pg/rr?
0.032
0.064
0.062

0.195

0.095

Possible
Number of
PM-Related
Deaths/Day
0.7
1.0
1.5

0.3

0.9

'From Monthly Vital Statistics Report for 1991 (U.S. CD 1993).
2From EPA meta-analyses, Table 12-25; all models without co-pollutants
3Elderly as 12.6% of 1991 U.S. population
4From Saldiva and Bohn (1994) and Ostro et al. (1995), variance-weighted average (TWA); Section 12.3.1.3
5From Pope, et al. (1991), Schwartz (1993) for Utah Valley and Birmingham TWA, Table 12-4
    TABLE 1-15. ESTIMATED NUMBER OF DEATHS PER DAY IN CITIES OF
       10,000 to 10 MILLION^OR WHICH AN INCREASE OF 50 jtg/m3 PM10
                     COULD BE A CONTRIBUTING FACTOR
Expected Number of PM-Related Excess Deaths Per Day
Population
of City
10 Million
5 Million
1 Million
500,000
100,000
50,000
10,000
Whole Pop.
All Causes
< 2d Lag
~4 -7
~2 -4
-0.4 -0.7
-0.2 -0.4
-0.05
-0.03
-0.005
65+ Pop.
All Causes
< 2d Lag
-5 - 10
-2.5 - 5
-0.5 - 1
-0.3 - 0.5
-0.07
-0.04
-0.01
Whole Pop.
All Causes
3-5 Day Lag
-7 - 15
-4 - 8
-0.8 - 1.5
-0.4 -0.8
-0.01
-0.05
-0.01
Whole Pop.
Respiratory
3-5 Day Lag
-2 - 3
-1 - 2
-0.2 - 0.3
-0.1 - 0.15
-0.03
-0.02
-0.002
Whole Pop.
Cardiovascular
3-5 Day Lag
-5 -9
-2 -5
-0.5-0.9
-0.2-0.5
-0.07
-0.04
-0.008
'Upper end of range for each city size calculated from upper bound estimates in Table 13-17 for population of 1
 million. Lower end of range derived as lower bound estimate roughly 50% less than the upper bound, as per
 text.
April 1995
1-95
DRAFT-DO NOT QUOTE OR CITE

-------
 1     mixes having 50-lOOjug/m3 increments in 24-h PM10 levels would most likely be projected as
 2     possibly causing meaningful numbers of excess deaths mainly as such exposures occur for
 3     large segments of the elderly population (age 65 + yrs) with preexisting cardiopulmonary
 4     diseases in rather large cities exceeding 1-2 million population.  The level of public concern,
 5     however, even in such cases may be tempered by the likelihood that a majority of those at
 6     special risk are  most probably current or former  smokers, given the predominant role
 7     attributed (U.S. EPA,  1992) to smoking in the etiology of preexisting cardiopulmonary
 8     diseases that distinguish those identified as being at greater risk.
 9          There is some limited direct evidence for an interaction between smoking status and
10     excess  mortality attributable to PM exposure.  Based on the Six Cities Study, Dockery, et.
11     al. (1993) reported an increased RR for PM  2 5 between the least polluted city (Portage) and
12     the most polluted city (Steubenville) that is substantially (albeit not  statistically significantly)
13     higher  in individuals who are current or former smokers, compared to never-smokers.  This
14     is shown in Table 1-16. Prospective studies that have individual data on smoking status
15     probably offer the best opportunity for detecting  differential effects  of smoking status on PM-
16     related mortality and morbidity for use in future  criteria assessments.
17
18
         TABLE 1-16.  ASSOCIATION BETWEEN CIGARETTE SMOKING STATUS AND
        EXCESS MORTALITY RISK FROM AIR POLLUTION IN THE SIX CITIES STUDY
Relative Risk for Worst PM2 5 City (Steubenville)
Versus Lowest Best PM2 5 City (Portage)
Smoking Status
Never Smoker
Former Smoker
Current Smoker
M+F
1.19
1.35
1.32
(95% CI)
(0.90,1.57)
(1.02,1.77)
(1.04,1.68)
M
1.29
1.31
1.42
(95% CI)
(0.80,2.09)
(0.96,1.80)
(1.05,1.92)
F
1.15
1.48
1.23
(95% CI)
(0.82,1.62)
(0.82,2.66)
(0.83,1.83)
             on Table 3 from Dockery, et. al (1993)
 1          The prematurity of the excess deaths is also a matter of considerable importance, but
 2     there is as yet little firm evidence from acute mortality studies by which to judge whether
 3     PM-related excess deaths generally represent highly compromised elderly individuals dying a

       April 1995                               1-96      DRAFT-DO NOT QUOTE OR CITE

-------
 1      few days or weeks sooner than they would have otherwise versus several months or years of
 2      prematurity of death for some.
 3           Morbidity  effects demonstrated as likely being associated with short-term exposures to
 4      ambient U.S.  PM exposures include increased hospital admissions for respiratory and
 5      cardiovascular disease conditions, increased respiratory symptoms (including exacerbation of
 6      asthma), and small pulmonary function decrements (e.g., 2-3% decreases in FEVj or FVC).
 7      Probably of most immediate public health concern are the hospital admissions, which are also
 8      more readily quantifiable and understandable as an index of obviously serious health impacts.
 9      Table 13-20 summarizes key types of information by which one might attempt to project
10      increments in hospital admissions for which exposure to an increment in ambient PM10
11      (24-h) of 50 /ig/m3 might contribute per 1 million people exposed.   Table 1-17 can be
12      interpreted in an analogous fashion to Table 1-14.  Note from Table 1-17 that the typical
13      number of PM10-related hospital admissions for cardiovascular causes would be projected to
14      be only about 2.5 times as high as the number of potential deaths during the same event, and
15      the number of respiratory admissions about 6 times as high as the possible number  of deaths
16      from respiratory causes shown in Table 1-14.  However, many deaths from cardiovascular or
17      respiratory causes occur without a prior hospital admission.  There is, nonetheless,  a
18      reasonable numeric consistency between the rough estimates of potential hospital admissions
19      or discharges  and possible total deaths contributed to by exposure in a community to PM-
20      containing ambient-air mixes.
21           Table 1-18 then scales expected daily hospital admissions potentially associated with
22      exposures to ambient air mixes having a 50 ^g/m3 increase in PM10 (24-h) for towns and
23      cities with populations of 10 thousand to 10 million (analogous to what was done earlier in
24      Table 1-15 based on Table 1-14 calculations).  However, in this case, both Tables 1-17 and
25      1-18 provide only upper bound estimates for hospital admissions  based on available analyses,
26      which did not include copollutants in the  models.  Essentially the same types of statements
27      as made with  regard to  the very small increases in excess risk depicted  in Tables 1-17 and
28      1-18 for mortality also generally apply here for hospital admissions, except  to note  somewhat
29      larger projected  numbers for possible hospital admission cases for which the ambient PM
30      exposure might be a contributing factor.
        April 1995                                1-97      DRAFT-DO NOT QUOTE OR CITE

-------
 TABLE 1-17. ESTIMATED HOSPITAL ADMISSIONS PER DAY IN A POPULATION OF
   ONE MILLION FOR WHICH AN INCREASE OF 50 /tg/m3 (24-h) PM10 COULD BE A
                           CONTRIBUTING FACTOR

First
Listed
Diagnosis
All conditions

Respiratory
Conditions (all)
Pneumonia


COPD

Heart
Disease


Age
Group
All
65 +
All
65 +
All
65 +

All
65 +
All
65 +
Population
Baseline
Annual
Discharges
124, HO1
42,8452
12, ISO1
54011
4,3401
2,3352

3,3377
2,5607
12,310
13,502
Population
Baseline
Daily Hospital
Discharges
340.0
117.4
33.4
14.0
11.9
6.4

9.1
7.0
58.4
37.0
Excess Risk
per PMJj?
50 /ig/m3
(Lag <. 1 d)
	
—
0.063
0.083


0.084
0.155
0.163
0.046
0.066
Possible Number
of PM-Related
Hosp. Admissions
Per Day
	
—
2.0
1.1


0.5
1.4
1.1
2.3
2.2
       Table 12-7
   3From Table 12-9, average
   5From Table 12-10, average            6From Table 12-12
   7From 1992 detailed Tables; excludes asthma (ICD9 493-493.9)
2From Table 12-7, assuming 12.6% age 65 +
4From Table 12-11, average
TABLE 1-18. ESTIMATED NUMBERS OF HOSPITAL ADMISSIONS FOR RESPIRATORY
 AND CARDIOVASCULAR CAUSES PER DAY IN CITIES OF 10,000 to 10 MILLION FOR
WHICH AN INCREASE OF 50 jtg/m3 PM10 (24-h) COULD BE A CONTRIBUTING FACTOR
Population
of City
10 Million
5 Million
1 Million
500,000
100,000
50,000
10,000
All Respiratory
Conditions
Whole 65 +
Pop. Pop.
20.0
10.0
2.0
1.0
0.2
0.1
0.02
11.0
5.5
1.1
0.55
0.11
0.05
0.01
Pneumonia
Whole 65 +
Pop. Pop.
5.0
2.5
0.5
0.25
0.05
0.02
0.01
COPD
Whole 65 +
Pop. Pop.
14.0
7.0
1.4
0.7
0.14
0.07
0.02
11.0
5.5
1.1
0.55
0.11
0.05
0.01
Heart Disease
Whole 65 +
Pop. Pop.
23.0
11.5
2.3
1.15
0.23
0.12
0.02
22.0
11.0
2.2
1.1
0.22
0.11
0.02
   April 1995
     1-98
DRAFT-DO NOT QUOTE OR CITE

-------
1           Overall, based on the foregoing discussions, there appears to exits credible evidence for
2     a likely very small, but real PM effect on human health in some susceptible subpopulations
3     (including contributing along with other risk factors to premature  deaths among the elderly
4     with preexisting cardiopulmonary diseases) at PM10 24-h concentrations in the range  of 30 to
5     200 jug/m3.   However, the biological mechanisms by which such effects occur are  as yet not
6     well understood and remain to be delineated, as is the case for clearer characterization and
7     interpretation of relative risk estimates for PM-related effects and their appropriate use in
8     projecting potential public health impacts.
      April 1995                                1-99       DRAFT-DO NOT QUOTE OR CITE

-------
  i                                 2.  INTRODUCTION
  2
  3
  4           This document is an update revision of "Air Quality Criteria for Particulate Matter and
  5      Sulfur Oxides" published by the United States Environmental Protection Agency (EPA) in
  6      1982, an Addendum to that document published in 1986, and an Acid Aerosols Issue Paper
  7      published in 1989, and it will serve as the basis for reevaluating the current National
  8      Ambient Air Quality Standard (NAAQS) for particulate matter (PM). The present document
  9      critically evaluates and assesses the scientific information relative to determining the health
 10      and welfare effects associated with exposure to various concentrations of PM in ambient air.
 11      Although the document is not intended as a complete and detailed  literature review, it is
 12      intended to cover pertinent literature through its publication date.  The literature through that
 13      time is reviewed thoroughly for information relevant to criteria development. Though the
 14      emphasis is on the presentation of data on health and welfare effects, other scientific data are
 15      also discussed in order to provide a better understanding of the pollutants in the environment.
 16
 17
 18      2.1   LEGISLATIVE  REQUIREMENTS
 19           Two sections of the CAA (Sections 108 and 109, U.S. Code, 1991) govern the
 20      establishment, review, and revision of National Ambient Air Quality Standards (NAAQS).
 21      Section 108 directs the Administrator of the U.S.  Environmental Protection Agency (EPA) to
 22      list pollutants that may reasonably be anticipated to endanger public health  or welfare and to
 23      issue air quality criteria for them.  The air quality criteria are to reflect the latest scientific
 24      information useful in indicating the kind and extent of all exposure-related effects on public
 25      health and welfare that may be expected from the  presence of the pollutant in ambient air.
 26           Section 109(a, b) directs the Administrator of EPA to propose and promulgate
27      "primary" and  "secondary" NAAQS for pollutants identified under Section 108. Section
28      109(b)(l) defines a primary standard as a level of air quality, the attainment and maintenance
29      of which, in the judgment of the Administrator, based on the criteria and allowing for an
30      adequate margin of safety,  is requisite to protect the public health.   Section 109(d) of the
31      CAA requires periodic review and, if appropriate, revision of existing criteria and standards.
        April 1995                                2-1       DRAFT-DO NOT QUOTE OR CITE

-------
 1     Under Section 109(b) of the CAA, the Administrator must consider available information to
 2     set secondary NAAQS that are based on the criteria and are requisite to protect the public
 3     welfare from any known or anticipated adverse effects associated with the presence of such
 4     pollutants.  The  welfare effects included in the criteria are effects on vegetation, crops, soils,
 5     water, animals, manufactured materials,  weather, visibility, and climate, as well as damage
 6     to and deterioration of property,  hazards to transportation, and effects on economic value and
 7     personal comfort and well-being.
 8
 9
10     2.2  REGULATORY BACKGROUND
11           "Particulate matter" is the generic term for a broad class of chemically and physically
12     diverse substances that exist as discrete particles (liquid droplets  or solids) over a wide range
13     of sizes. Particles originate from a variety of stationary and mobile sources.  They may be
14     emitted directly or formed in the atmosphere by transformation of gaseous emissions such as
15     sulfur oxides (SOX),  nitrogen oxides (NOX), and volatile organic  substances.  The chemical
16     and physical properties of PM vary greatly with time, region, meteorology, and source
17     category, thus complicating the assessment of health and  welfare effects.  Particles in
18     ambient air usually occur in two somewhat overlapping bimodal  size distributions:  (1) fine
19     (diameter less than 2.5 jum) and (2) coarse (diameter larger than 2.5 ji«n). The two size
20     fractions tend to have different origins and composition.
21           On April 30, 1971 (Federal Register, 1971), EPA promulgated the original primary and
22     secondary NAAQS for paniculate matter (PM) under Section 109 of the CAA.  The
23     reference method for measuring attainment of these standards was the "high-volume" sampler
24     (Code of Federal Regulations,  1986), which collects PM up to a nominal size of 25 to 45 /xm
25     (so-called "total  suspended paniculate," or "TSP").  Thus, TSP was the original indicator for
26     the PM standards. The primary  standards for PM (measured as  TSP) were 260 jwg/m3, 24-h
27     average not to be exceeded more than once per year, and 75 pig/m3, annual geometric mean.
28     The secondary standard (measured as TSP) was 150 ptg/m3,  24-h average not to be exceeded
29     more than once per year.
30           On October 2,  1979  (Federal Register,  1979a), EPA announced that it was in the
31     process of  revising the AQCD and reviewing the existing air quality standards  for possible

       April  1995                                2-2       DRAFT-DO NOT QUOTE OR CITE

-------
  1     revisions. External review drafts of that revised AQCD were made available for public
  2     comment and peer review by the Clean Air Scientific Advisory Committee (CASAC) of
  3     EPA's Science Advisory Board (SAB). CASAC prepared a "closure" memorandum to the
  4     Administrator indicating its satisfaction with the final draft of the AQCD.  After closure,
  5     minor technical and editorial refinements were made to the AQCD  (U.S. Environmental
  6     Protection Agency, 1982).  The final draft (December 1981) of the document was issued
  7     simultaneously with the proposal of revisions to the PM standards.
  8           On March 20, 1984 (Federal Register, 1984), EPA proposed  a number of revisions to
  9     the primary and secondary PM standards.  Following publication of the proposal,  EPA held a
10     public meeting in Washington, DC, on April 30,  1984, to receive comments on the proposed
11     standards revisions. After the close of the original public comment period (June 5, 1985),
12     CASAC  met on December 16 and 17, 1985, to review the proposal and to discuss the
13     relevance of certain new scientific studies on the health effects of PM that had emerged since
14     CASAC  completed its review of the AQCD and staff paper in January 1982.  Based on its
15     preliminary review of these new studies, CASAC recommended that EPA prepare separate
16     addenda  to the AQCD and staff paper for the  purpose of evaluating relevant new studies and
17     discussing their potential implications for standard-setting.  The EPA announced its
18     acceptance of these recommendations on April 1, 1986 (Federal Register,  1986a).  On July
19     3, 1986, EPA announced (Federal Register, 1986b) the availability  of the external review
20     draft  document, entitled Second  Addendum to Air Quality Criteria  for Particulate  Matter and
21     Sulfur Oxides (1982):  Assessment of Newly Available Health Effects Information (U.S.
22     Environmental Protection Agency, 1986).  At the same time, on July 3, 1986, EPA
23     announced a supplementary comment period to provide the public an opportunity to comment
24     on the implications of the new studies and  addenda for the final standards.  The CASAC held
25     a public meeting on October 15  and 16, 1986, to review the AQCD addendum. At this
26     meeting, CASAC members, as well as representatives of several  organizations, provided
27     critical review of the EPA documents.
28           The CASAC sent a closure letter on the AQCD addendum to  the Administrator dated
29     December 15, 1986, which stated that the 1986 addendum and the  1982 AQCD, previously
30     reviewed by CASAC,  represent a scientifically balanced and defensible summary of the
31     extensive scientific literature on PM and SOX (Lippmann, 1986b).

       April  1995                              2-3       DRAFT-DO NOT QUOTE OR CITE

-------
  1          On July 1,  1987 (Federal Register, 1987), EPA published final revisions to the NAAQS
  2      for PM.  The principal revisions in 1987 included (1) replacing TSP as the indicator for the
  3      ambient standards with a new indicator that includes only particles with an aerodynamic
  4      diameter less than or equal to a nominal 10 pirn ("PM10"), (2) replacing the 24-h primary
  5      TSP standard with a 24-h PM10 standard of 150 /xg/m3, (3) replacing the annual primary TSP
  6      standard  with an annual PM10 standard of 50 ^g/m3,  and  (4)  replacing the  secondary TSP
  7      standard  with 24-h and annual PM10 standards identical in all respects to the primary
  8      standards.
  9
10
11      2.3   SCIENTIFIC BASIS FOR THE EXISTING  PARTICIPATE
12            MATTER STANDARDS1
13          The following discussion describes the bases for the existing PM NAAQS set in 1987.
14      The discussion includes the rationale for the primary standards, the pollutant indicator for
15      paniculate matter, the averaging time and form of the standard, and finally a discussion of
16      EPA's assessment that led to the standard set in 1987.
17
18      2.3.1    Rationale  for the Primary Standards
19          In selecting primary standards for PM, the Administrator must specify (1) the particle
20      size fraction that is to be used as an indicator of paniculate pollution, (2) the appropriate
21      averaging times and form(s) of the standards, and (3) the  numerical levels of the standards.
22      Based on the assessment of relevant scientific and technical information in the earlier 1982
23      PM AQCD and addenda, the staff paper and staff paper addendum outlined a number of key
24      factors considered in making decisions in each of these areas.  The following discussion of
25      the 1987  revisions of the standards focuses mainly on the  considerations that were most
26      influential in the Administrator's selection of particular options.
27
28
29     Adapted from Federal Register (1987) National Ambient Air Quality Standard for Paniculate
30     Matter.
31
       April 1995                              2-4       DRAFT-DO NOT QUOTE OR CITE

-------
  1      2.3.2    Pollutant Indicator

  2           Based on the assessment of the available scientific information, EPA concluded in 1987
  3      that (1) a separate PM standard (as opposed to a combination standard for PM and SOX)
  4      remained a reasonable public health policy choice, and (2) given current scientific knowledge

  5      and uncertainties, a size-specific (rather than chemical-specific) indicator should be used.  In
  6      assessing the information in the AQCD, EPA reached several conclusions  summarized here.

  7
  8        (1)  Health risks posed by inhaled  particles are influenced by both the penetration and
  9            deposition of particles in the various regions of the respiratory tract and the biological
10            responses to these deposited materials.  Smaller particles penetrate furthest in the
11            respiratory tract. The largest  particles are deposited predominantly in the
12            extrathoracic (head) region, with somewhat smaller particles depositing in the
13            tracheobronchial region;  still  smaller particles  can reach the deepest portion of the
14            lung, the pulmonary region.
15
16        (2)  The risks of adverse health effects associated with deposition of typical ambient fine
17            and coarse particles in the thoracic region (tracheobronchial and pulmonary deposition)
18            are markedly greater than those associated with deposition in the extrathoracic region.
19            Maximum particle penetration to the thoracic region  occurs during oronasal or mouth
20            breathing.
21
22        (3)  The size-specific indicator for primary standards should represent those particles small
23            enough to penetrate to the thoracic region.  The  risks of adverse health effects from
24            extrathoracic deposition of typical ambient PM are sufficiently low  that particles
25            depositing only in that region  can safely be excluded from the indicator.

26
27           Considering the above conclusions, together with  information on air  quality
28      composition, the need to provide protection for sensitive individuals who may breathe by

29      mouth or oronasally and the similar convention on particles penetrating the thoracic region
30      adopted by  the International Standards Organization (1981), EPA staff recommended that the
31      size-specific indicator include particles of diameters less than or equal to a nominal 10 nm
32      "cut point"  generally referred to as "PM10".  In terms of collection efficiency, this represents

33      a 50%  cut point or diameter (D50) the aerodynamic particle diameter for which the efficiency

34      of particle collection is 50%. With  such a cut point, larger particles are not excluded

35      entirely but are collected with substantially decreasing efficiency, and smaller particles are
36      collected  with increasing (up to 100%) efficiency.  Ambient samplers with this cut point

37      provide a reliable estimate of the total mass of suspended PM of aerodynamic size less than

        April 1995                                 2-5        DRAFT-DO NOT QUOTE OR CITE

-------
 1      or equal to 10 /mi.  Such an indicator (PM10) is conservative with respect to health
 2      protection in that it includes all of the particles small enough to penetrate to the sensitive
 3      pulmonary region and includes approximately the same proportion of the coarse-mode
 4      fraction that would be expected to reach the tracheobronchial region.  It places substantially
 5      greater emphasis on controlling smaller particles than does a TSP indicator, but does not
 6      completely exclude larger particles from all control.
 7           The assessment of then-available information on respiratory tract deposition in  the 1986
 8      AQCD and staff paper addenda reinforced the conclusions reached in the original EPA
 9      assessment.  In particular, (1) the data do not provide support for an indicator that excludes
10      all particles larger than 10 /zm in diameter; (2) the analysis used to support an alternative
11      indicator with a nominal size cut point of 6 /mi (Swift and Proctor, 1982) significantly
12      underestimated thoracic deposition of particles larger than 6 /mi  in diameter under natural
13      breathing conditions; (3) the PM10 indicator generally includes a similar or larger fraction of
14      the range of particles that can deposit in the tracheobronchial region,  although it appears to
15      be somewhat less conservative in this regard than previously thought with respect to large
16      (> 10 /mi) particle deposition under conditions of natural mouthbreathing; and (4) the studies
17      of tracheobronchial deposition generally involved adult subjects (other information indicating
18      even greater tracheobronchial deposition of particles in children than in  adults provides an
19      additional  reason for an indicator that includes particles capable of penetration to the
20      tracheobronchial region).  Consideration of these and the earlier  conclusions led EPA to
21      reaffirm its recommendation for a PM10 indicator.  The CASAC also restated its support for
22      PM10 in its review of the proposal and the closure letter to the Administrator (Lippmann,
23      1986a,c).
24           In 1987  the Administrator accepted the recommendations of the  staff and CASAC, as
25      well as their underlying rationale, and decided to replace TSP as the particle indicator for the
26      primary standards with a new indicator that includes only  those particles less than a nominal
27      10 /xm in diameter (PM10) as specified in the Federal Reference  Method.
28
29      2.3.3   Averaging Time and Form of the Standards
30           The EPA's assessment at that time of scientific information available prior to 1987
31      confirmed the need for both short- and long-term standards for PM.  The alternative of a

        April  1995                                2-6       DRAFT-DO NOT QUOTE OR CITE

-------
  1      single averaging time would not provide adequate protection against potential effects from
  2      both long- and short-term exposures without being unduly restrictive.  The forms for the
  3      24-h and annual standards are discussed below.
  4
  5      24-Hour Standard
  6           The Environmental Protection Agency proposed in 1987 that the 24-h standard be stated
  7      in a statistical form that uses more than 1 year of data and accounts for variations in
  8      sampling frequency in order to predict  the actual number of exceedances to be expected in an
  9      average year.  When used with an appropriate standard level,  the statistical form can provide
10      improved health protection that is less sensitive to changes in sampling frequency than the
11      deterministic form and  also can offer a more stable target for control programs.  Recognition
12      of the limitations of the deterministic form also led EPA to promulgate a statistical form for
13      the ozone standard (Federal Register, 1979b).
14
15      Annual Standard
16           The Administrator decided to change the form of the annual standard in 1987 from the
17      previous annual geometric mean form to a statistical form expressed as an expected annual
18      arithmetic mean.  The expected annual  arithmetic mean is equivalent to  the long-term
19      arithmetic average concentration level,  assuming  no changes in underlying emissions.  The
20      expected arithmetic mean is more directly related to the available health effects information
21      than is the annual geometric mean, which was the previous form of the  standard. Because
22      the arithmetic mean concentration is proportional to the sum of the  daily means, it reflects
23      the total cumulative exposure of PM to which an individual is exposed.  Therefore, it is an
24      appropriate  indicator to protect against  any health effect that depends on chronic, total
25      exposure.  It is also a reasonable indicator for protecting against health effects that depend  on
26      repeated short-term high concentrations (short-term peaks have an influence on the arithmetic
27      mean that is proportional to their frequency, magnitude, and duration).  The geometric mean,
28      on the other hand,  deemphasizes the effect of short-term peak concentrations and is heavily
29      influenced by days of relatively clean air.  For these reasons, EPA  staff and CAS AC
30      recommended the  change to an arithmetic mean.
        April 1995                                 2-7       DRAFT-DO NOT QUOTE OR CITE

-------
 1          Under the statistical form, the expected annual arithmetic average is determined by
 2     averaging the annual arithmetic averages from 3 successive years of data.  The prior
 3     deterministic form of the standard did not adequately take into account the random nature of
 4     meteorological variations.  In general, annual mean PM concentrations will vary from year to
 5     year, even if emissions remain constant, due to the random nature of meteorological
 6     conditions that affect the formation and dispersion of particles in the atmosphere.  If only 1
 7     year of data is considered, compliance with the standard and, consequently, emission control
 8     requirements, may be determined on the basis of a year with unusually adverse  or unusually
 9     favorable weather conditions.  The problem of year-to-year variability is, however, reduced
10     by averaging 3 years of data.
11
12     2.3.4    Level  of the Standards
13          The original Office of Air Quality Planning and Standards  (OAQPS) PM Staff Paper
14     and CASAC recommendations set forth a framework for determining the levels  for the
15     proposed PM standards  that  would protect public health with an adequate margin of safety.
16     The discussion that follows relies heavily on that framework and on the supporting material
17     in the staff paper and its addendum, as well as the CASAC closure letters.  The essential
18     steps in this framework  are summarized here.
19
20     Assessment of the Quantitative Epidemiological Studies
21          The 1982 AQCD and its 1986 addendum identified a small number of community
22     epidemiological  studies that are useful in determining concentrations at which PM is likely to
23     adversely impact public health.  The EPA staff used these quantitative studies to examine
24     concentration-response relationships and to develop numerical "ranges of interest" for
25     possible  PM10 standards.
26          A number  of uncertainties associated with the use of these studies  had to be considered
27     in selecting an appropriate margin  of safety. As discussed in the staff paper and the AQCD,
28     and the addenda to those documents, epidemiological studies generally are limited in
29     sensitivity and are subject to inherent difficulties involving confounding variables.
30     Moreover, many of the  quantitative studies were conducted in times and places  where
31     pollutant composition may have  varied considerably from current U.S. atmospheres.  Also,

       April 1995                                2-8       DRAFT-DO NOT QUOTE OR CITE

-------
  1      most of the studies used British Smoke—British Smoke (BS) is a pseudo-mass indicator
  2      related to small particle (aerodynamic diameter less than a nominal 4.5 /*m) darkness—or
  3      TSP as particle indicators. None of the  published studies used the proposed PM10 indicator.
  4      Thus, assumptions had to be used to convert the various results to common (PM10) units.
  5
  6      Identification of Margin of Safety  Considerations
  7           The 1982 AQCD and its addendum identified an additional substantial body of scientific
  8      literature that, although it did not provide reliable concentration-response relationships for
  9      ambient exposures, did provide important qualitative insights into the health risks associated
10      with human exposure to particles.   This  literature  included both quantitative and qualitative
11      epidemiological studies, controlled  human exposure experiments, and animal toxicological
12      studies.   The EPA staff assessed this literature to identify additional factors and uncertainties
13      that should be considered in selecting the most appropriate margin of safety.
14           Experience has shown that it is difficult to identify, with confidence,  the lowest
15      pollution level at which an adverse  effect will occur.  Moreover,  in cases  such as the
16      present one, the evidence suggests that there is a continuum of effects, with the risk,
17      incidence, or severity of harm decreasing, but not necessarily vanishing, as the level of
18      pollution is decreased.
19           The requirement for an adequate margin of safety for primary  standards addresses
20      uncertainties associated with  inconclusive scientific and technical information available at the
21      time of standard setting.  It also aims to  provide a reasonable degree of protection against
22      hazards that research  has not yet identified.   Both kinds of uncertainties are components  of
23      the risk associated with pollution at levels below those at which human health effects can be
24      said to occur with reasonable scientific certainty.  Thus, by selecting primary standards that
25      provide an adequate margin of safety, the Administrator sought not only to prevent pollution
26      levels that have been demonstrated  to be harmful,  but also to prevent lower pollutant levels
27      that may  pose an unacceptable risk  of harm,  even  if that risk is not precisely identified as to
28      nature or degree.
29           In the absence of clearly identified  thresholds for health effects, the selection of a
30      standard that provides an adequate margin of safety requires an exercise of informed
31      judgment by the Administrator.  The level selected will depend on the expected incidence

        April 1995                                 2-9       DRAFT-DO NOT  QUOTE OR CITE

-------
 1     and severity of the potential effects and on the size of the population at risk, as well as on
 2     the degree of scientific certainty that the effects will in fact occur at any given level of
 3     pollution.
 4           The 1986 EPA staff paper recommended a range of potential standards for the
 5     Administrator's consideration.  The recommended range  was below the levels at which EPA
 6     staff, with the concurrence of CAS AC, had concluded from the available data that adverse
 7     health effects were "likely", but in the domain where  the data suggested that such effects
 8     were "possible".   The Administrator proposed refined ranges of standard levels that were
 9     based on the 1984 staff and CAS AC recommendations.   After consideration of the then new
10     scientific evidence contained in the AQCD addendum, the staff revised its recommended
11     range of standards.  The Administrator considered the revised assessments and the
12     recommendations  of CASAC (Lippmann, 1986b) in making the  final decision on the standard
13     levels.  The rationales for the levels of the 24-h and annual standards are presented below.
14
15     24-Hour Standard
16           The 1987 assessment of the short-term epidemiological data expresses PM levels in
17     both  the BS or TSP and PMIO units.  The term "effects likely"  denoted concentration ranges
18     derived from  the 1982 AQCD and its addendum at or above which a consensus judgment
19     suggests the greatest certainty that the effects studied would occur, at least under the
20     conditions that occurred in the original studies. In the "effects  possible" range, EPA found
21     credible scientific evidence suggesting the existence of adverse health effects in sensitive
22     populations, but substantial uncertainty exists regarding the conclusions to be drawn from
23     such  evidence.
24           The 1987 review of the data did not provide evidence of clear thresholds in exposed
25     populations.   Instead, they suggested  a continuum of response for a given number of exposed
26     individuals, with both the likelihood (risk) of any effects occurring and the extent (incidence
27     and severity)  of any potential effect decreasing with concentration (this is particularly true for
28     the statistical  analyses of daily mortality in London).  Substantial agreement existed that
29     wintertime pollution episodes produced premature mortality in elderly and ill populations, but
30     the range and nature of association provide no clear basis for determining lowest
31     effects-likely  levels or for defining a concentration below which no association remained.

       April 1995                               2-10       DRAFT-DO NOT QUOTE OR CITE

-------
 1      The lung function studies in children also provided evidence of effects at concentrations over
 2      a range, but the relationships were not certain enough to derive effects-likely levels for
 3      PM10.  The lung function studies did, however, suggest levels below which detectable
 4      functional changes were unlikely to occur in exposed populations.  Following CAS AC
 5      recommendations, EPA used the combined range of effects-possible studies as a starting
 6      point for developing alternative standards.
 7           The original range proposed by the Administrator, drawn from the 1984 staff analysis,
 8      was 150 to  250 /xg/m3 PM10 24-h average, with no more than one expected exceedance per
 9      year.  The lower bound of this range was derived from the original assessment of the London
10      mortality studies.  As a result of reanalyses of the London mortality data and the findings
11      from the then current U.S.  morbidity studies, the staff reduced the level of the lower bound
12      of the range of interest to 140 /ig/m3, and noted that the difference between it and the
13      original lower bound (150 /tig/m3) was within the range of uncertainty associated with
14      converting the morbidity study results from TSP to PM10.
15           At that tune the study of Lawther et al. (1970) was judged to provide evidence that
16      health effects  are likely at PM concentrations above 250 Mg/m3 (as BS).  The effects
17      observed in this study (related to aggravation of bronchitis) are of concern because of both
18      their immediate impact and their potential for inducing longer term deterioration of health
19      status in a significant sensitive group.  Based on the uncertain conversion between smoke and
20      PM10, the lowest effects likely level derived from the Lawther study (250 jig/m3 as BS)
21      should be in the range of 250 to 350 /ig/m3 in PM10 units.
22           The 1987 assessment of the Lawther et al. (1970) study formed the basis for the upper
23      bound of the range of PM10 standards proposed by the Administrator in 1984.  Considering
24      this study alone, a PM10 standard of 250 /xg/m3 might appear to contain some margin of
25      safety, even for the sensitive bronchitics studied,  because  it incorporated a conservative PM10
26      conversion factor and because of differences between exposure conditions in the British study
27      and current U.S. air quality.  Because bronchitics are identified as a group particularly
28      sensitive to  particulate pollution, a standard of 250 /xg/m3 (as PM^) also might provide some
29      margin of safety for other, less  sensitive groups.  Nevertheless, this study of bronchitics in
30      London has inherent limitations in sensitivity that preclude derivation of unequivocal "effects
31      thresholds"  at 250 ^ig/m3 as BS and, by extension,  PM10.  The 1982 AQCD noted that
        April 1995                                2-11       DRAFT-DO NOT QUOTE OR CITE

-------
 1     associations between pollution and health status persisted at lower BS concentrations in
 2     selected, more sensitive individuals.  Although the lead author of the study objected to
 3     attaching any importance to these latter findings (Lawther, 1986), EPA, with CASAC
 4     concurrence, found no basis for asserting that this study demonstrated a population threshold
 5     at 250 /zg/m3.
 6           In evaluating the margin of safety for a 24-h standard, it was also important to consider
 7     the London mortality studies.  A standard at the upper portion of the proposed range (250
 8     /^g/m3) would be well below the levels (500 to 1,000 /zg/m3 as BS) of the historical London
 9     episodes in which the scientific  consensus indicated that pollution was responsible for excess
10     mortality.  The portions of the population at greatest risk  of premature mortality  associated
11     with PM exposures in those episodes included the elderly and persons with preexisting
12     respiratory or cardiac disease.  Although the extent of life shortening could not be specified,
13     the seriousness of the effect strongly justified a margin of safety for it (below the consensus
14     effects levels) that was larger than that warranted for the  effects on bronchitis.
15           The staff assessment  at that time of several reanalyses of London mortality  suggested,
16     however, that the risk of premature mortality to sensitive individuals extended to
17     concentrations substantially lower than those that occurred in the "episodes".  Other analyses
18     (Mazumdar et al., 1982; Ostro,  1984; Shumway et al., 1983) provided no objective support
19     for a population threshold  below which such a risk no longer existed.  Although the risk to
20     individuals may be small at concentrations of 250 /xg/m3 and below, the number of people
21     exposed to lower concentrations, given U.S. levels, was substantially larger than the number
22     exposed to higher levels.   The increased number of individuals exposed increased the risk
23     that effects would occur in the total population exposed.
24           Differences in the composition of particles and gases among U.S. cities  and between
25     conditions in the United States and London at the  time the mortality and morbidity data were
26     gathered added to the complexity of assessing the  risk associated with PM in  the United
27     States.  In the case of the  mortality studies, however,  the staff found that at least one of the
28     studies (Ozkaynak and Spengler, 1985) provided qualitative support for an association
29     between daily mortality and particle concentrations in then nearly contemporary U.S.
30     atmospheres.
        April 1995                                2-12      DRAFT-DO NOT QUOTE OR CITE

-------
 1           The 1982 assessment of the mortality studies and related factors prompted the EPA
 2      Administrator to consider standard levels that extended from 250 ^g/m3 to the lower bound
 3      of the original staff range (150 Mg/m3) and even lower.   Reanalyses of the London mortality
 4      data prior to 1987 provided additional evidence that serious adverse health effects may occur
 5      at PM concentrations below 250 /^g/m3.  These analyses addressed a number of the
 6      uncertainties associated with the earlier studies and reinforced the Administrator's concern
 7      that a 24-h standard at the upper end of the proposed range may  not provide an adequate
 8      margin of safety. However, given the uncertainties in converting from BS to PM10
 9      measurements, particularly at lower concentrations, and the possible differences in particle
10      composition between London at the time the data were gathered and the contemporary United
11      States,  it was difficult to use these studies to set a precise level for a PM10 standard.
12           Given these difficulties, it was important to examine studies contemporary with the
13      other studies that utilize gravimetric measurements  of particulate  concentrations. The staff
14      found the studies of Dockery  et al. (1982)  and Dassen et al.  (1986) to be  useful.  The
15      Dockery study observed physiologically small but statistically significant decreases in lung
16      function in a group of children exposed to  peak PM10 levels  of 140 to 250 ^g/m3.  The
17      decrements persisted for 2 to  3 weeks following the exposures.  The study also suggested the
18      possibility of larger responses in a subset of the children, including those  with existing
19      respiratory symptoms.  The Dassen study recorded similar decrements in  children in the
20      Netherlands following exposure to PM10 levels estimated at 200 to 250 /ig/m3, but no
21      observable effects 2 days  after exposure  to PM10 levels  estimated at  125 /xg/m3.  The particle
22      composition, at least in the Dockery study, was more representative of contemporary U.S.
23      cities in that time period,  and the  associated  aerometry provided a more reliable estimate of
24      PMio levels than did the measurements used in the London studies.  It was reasonable  to
25      expect the endpoints observed (small reversible reductions  in lung function in children) to be,
26      in most cases,  more sensitive to air pollution than those observed in the London studies.
27      These effects are, of themselves, of uncertain significance to health,  but might be associated
28      with aggravation of respiratory symptoms in children with  preexisting illness. Long-term
29      examination of respiratory health in the same community studied by Dockery et al.  (1982)
30      suggested that the children in that  community had a higher incidence of respiratory illness
31      and symptoms than children in communities with lower particle levels, but the data showed

        April 1995                                 2-13       DRAFT-DO NOT QUOTE OR CITE

-------
 1      no evidence for any persistent reduction in lung function (Ware et al., 1986).  Uncertainties
 2      with respect to the effects of other pollutants (e.g., sulfur dioxide), the consistency of the
 3      changes, and exposures precluded specifying unequivocal "effects likely" levels based on this
 4      study.  The EPA assessment therefore suggested that short-term lung function effects in
 5      children were possible across a  range of 140 to 250 /^g/rn3 or more as PM10.
 6           In making a decision on a final standard level, the Administrator also considered
 7      information from the more qualitative studies of PM.  These studies suggested increased risks
 8      for sensitive groups (asthmatics) and risks of potential effects (morbidity in adults) not
 9      demonstrated in the more  quantitative epidemiological literature.  The qualitative studies did
10      not provide clear information on effect levels but did justify  consideration of effects of PM
11      that have not been sufficiently investigated.
12           Based on the 1982 assessment of the available scientific data, in 1984, the
13      Administrator expressed an inclination to select a 24-h level  from the lower portion of the
14      proposed range of 150 to 250 /tg/m3.  The addendum to the  1982 assessment supported the
15      original findings and, if anything, suggested an even wider margin of safety was warranted.
16      The Administrator, therefore, decided to set the final standard at the extreme lower bound of
17      the range originally proposed (i.e.,  150 Mg/m3)-  This standard provided a substantial margin
18      of safety below the levels at which there was a scientific consensus that PM caused
19      premature mortality and aggravation of bronchitis.  Such a margin was necessary because of
20      the seriousness of the effects and because of the analyses of  daily mortality studies that
21      suggested that adverse effects may occur at PM levels well below the consensus  levels.  The
22      standard was in the lower portion of the range where sensitive,  reversible physiological
23      responses of uncertain health significance possibly, but not definitely, are observed  in
24      children.  Using a conservative  assessment of the lung function/particle relationship from
25      Dockery et al.  (1982), a change in concentration from background levels ( = 20 /xg/m3) to
26      150 /xg/m3 would produce lung  function changes of at most  10 to 15%  in less than 5% of
27      exposed children.   Based on the results of Dassen et al. (1986), it appeared unlikely that  any
28      functional changes would  be detected 1 or 2 days following such exposures. Thus,  the
29      maximum likely changes in lung function appeared to present little risk of significant adverse
30      responses.  Standards set at a somewhat higher level would,  however, present an
31      unacceptable risk  of premature  mortality and allow the possibility of more significant

        April  1995                               2-14      DRAFT-DO  NOT QUOTE OR CITE

-------
  1     functional changes.  Furthermore, a standard level of 150 /xg/m3 was fully consistent with the
  2     recommendations of CAS AC on the 24-h standard (Lippmann, 1986c).
  3
  4     Annual Standard
  5          The long-term epidemiological studies examined in  1987 were subject to confounding
  6     variables that reduce the studies'  sensitivity and make their interpretation difficult.  No clear
  7     thresholds could be identified for the effects-likely levels, and evidence existed for effects at
  8     lower levels (the effects-possible levels); however, the evidence was inconclusive, and the
  9     effects were difficult to detect.
 10          Based on an EPA assessment of PM10/TSP ratios in areas  with elevated TSP levels, the
 11     effects-likely levels from the Ferris et al. (1973) study were revised to'80 to 90 pig/m3 as
 12     PM10.  Because of limitations in sampling duration and the conversion to PM10, this estimate
 13     was particularly uncertain,  with effects possible at lower concentrations. Of greatest concern
 14     was the possibility of long-term deterioration of the respiratory system in exposed
 15     populations, the potential for which is indicated by lung function (mechanical pulmonary)
 16     changes and increased incidence of respiratory disease. One set of studies  (Ferris et al.,
 17     1973, 1976) provided some evidence for a "no-observed-effect level" for those  effects at or
 18     below 60 to 65 /xg/m3 as PM10 (130 /ng/m3 as TSP),  whereas another study (Bouhuys et al.,
 19     1978) suggested some possibility of symptomatic responses  in adults at long-term median
 20     levels at or below about 50 to 55 /xg/m3 as PM10.  The importance of these symptomatic
 21     responses, which were unaccompanied by lung function changes, to long-term respiratory
 22     health was unclear.
 23          The most important study of long-term effects at that time was an ongoing examination
 24     of six U.S. cities (Ware et al., 1986). The study  indicated the possibility of increased
 25     respiratory symptoms and illnesses in children at multiyear levels across a range of 40 to
 26     more than 58 /^g/m3 as PM10 but found no evidence of reduced lung function at these
 27     concentrations.  This study  did not find similar gradients in symptoms and illness within
 28     some of the cities, which had somewhat smaller localized  pollution gradients.  The results of
29     a  separate series of studies of long- and intermediate-term (2- to 6-week) exposures in a
30     number of U.S. metropolitan areas (Ostro, 1987; Hausman et al.,  1984) were more
31      supportive of the possibility of effects  within cities (respiratory-related activity restrictions in

        April 1995                                2-15       DRAFT-DO NOT QUOTE OR CITE

-------
 1     adults) at comparable U.S. exposure levels.  The results of these studies were generally
 2     consistent with the earlier U.S. studies.  In particular, the finding of symptomatic responses
 3     in children with no change in lung function (Ware et al., 1986) was consistent with similar
 4     findings  in adults (Bouhuys et al.,  1978) at estimated long-term PM10 levels down to 50
 5     Mg/m3.  However, the information available to support the existence of significant adverse
 6     effects at annual PM10 levels below 50 /zg/m3 (especially when 24-h levels are maintained
 7     below 150 jig/m3) was quite limited and uncertain.
 8           Because of the uncertainties and the limited scope and number of long-term quantitative
 9     studies available for review in 1987, it was important to examine the results of qualitative
10     data from a number of epidemiological, animal, and ambient particle composition studies in
11     determining what constitutes an adequate margin of safety for an annual standard. These
12     studies justified concern for serious effects not directly evaluated in the above studies.  Such
13     effects included damage to lung tissues contributing to chronic respiratory disease, cancer,
14     and premature mortality.  Substantial segments of the population may be susceptible to one
15     or more  of these effects.  Although the qualitative data did not provide  evidence for major
16     risks  of these effects at the annual PM levels in most U.S. cities at that time,  the
17     Administrator believed, that the seriousness of the potential effects and the large population at
18     risk warranted caution in setting the standard.
19           Based on the findings discussed in the 1982 AQCD, in 1984, the Administrator
20     proposed to select an annual standard level from a range of 50 to 65  jig/m3.  In the proposal,
21     the Administrator favored a standard in the lower portion of the range.   The evidence
22     discussed in the 1986 addendum, although subject to substantial uncertainty,  served to
23     reinforce this inclination.  In light of the 1986 assessment, and in accordance with the
24     recommendation of CAS AC, the Administrator decided  to set the level of the annual standard
25     at the lower bound of the original range (50 /ng/m3, expected annual arithmetic mean).  This
26     standard provided a reasonable margin of safety against long-term degradation in lung
27     function, which was judged likely to occur at estimated  PM10 levels above 80 to 90 pig/m3
28     and for which there was some evidence at PM10 levels above 60 to 65 /*g/m3. Such a
29     standard also provided reasonable protection against the less serious symptomatic effects
30     (bronchitis) for which only inconclusive evidence was available. Moreover, the  staff and
31     CASAC recommended that the combined protection afforded by both 24-h and annual

       April 1995                               2-16       DRAFT-DO NOT QUOTE OR CITE

-------
  1      standards be considered in selecting the final standard level.  In this regard, analyses of air
  2      quality data showed that implementation of the 24-h standard would reduce substantially the
  3      annual levels in a number of areas  to below 50 /ig/m3, adding to the protection afforded by
  4      the annual standard in areas with higher 24-h peak-to-mean ratios.  Based on the then
  5      available information with respect to risks associated with annual exposures, the
  6      Administrator believed that the annual and 24-h standards provided an adequate margin of
  7      safety.
  8
  9      2.3.5   Welfare Effects
 10           No convincing evidence existed indicating significant adverse  soiling and nuisance at
 11      TSP levels below 90 to 100 ptg/m3, and, on that basis, the Administrator concluded that
 12      secondary standards different from the primary standards were not requisite to protect the
 13      public welfare against soiling and nuisance.  This conclusion was supported by CASAC's
 14      determination that there was  no scientific support for a TSP-based secondary  standard
 15      (Lippmann, 1986c).  Therefore, the Administrator decided to set 24-h and annual secondary
 16      PM10  standards that are equal in all respects to the primary standards.
 17           The other welfare effects of principal interest were impairment of visibility, potential
 18      modification of climate, and  contribution to acidic deposition.  All three of these effects were
 19      believed to be related to regional-scale levels of fine particles, and control programs designed
 20      to ameliorate them would likely involve region-wide reductions in emissions of sulfur oxides.
 21      The Administrator also concurred with the staff suggestions that a separate secondary particle
 22      standard was not needed to protect vegetation or to prevent adverse effects on personal
 23      comfort and well-being.
 24
 25
 26      2.4   TOPICS/ISSUES OF CONCERN FOR CURRENT CRITERIA
 27            DEVELOPMENT
28          Based on the available scientific evidence, several critical topics and associated issues
29      are addressed in this document,  as part of the current CAA-mandated periodic review of
30      criteria and NAAQS for PM.  Some of the most critical topics and  issues addressed are as
31      follows.

        April 1995                               2-17       DRAFT-DO NOT QUOTE OR CITE

-------
 1     2.4.1   Air Quality and Exposure
 2     Physics and Chemistry of Atmospheric Aerosols
 3           The atmospheric aerosols of interest because of their potential health and welfare effects
 4     consist of two principal components:  a gas phase ("air" in this case) and a solid or liquid
 5     particle phase in suspension.  Fine particles in the atmosphere consist mainly of (1) sulfate,
 6     nitrate,  ammonium ions, and water; (2) photochemically formed organic aerosols; and (3)
 7     carbon, organic matter, and metallic components emitted directly into the atmosphere.
 8     Coarse particles in the atmosphere are composed mainly of silica, calcium carbonate, clay
 9     minerals, soot, and, sometimes, organic substances.  A general relationship exists between
10     chemical composition and particle diameter, with particles of <2.5 /mi in diameter
11     containing most of the SO42", H+, and NH4+, as well as a significant fraction of the  NO3"
12     and Cl~.  The particle volume (mass) frequency function is often multimodal.  The fine-
13     volume fraction may have two or more modes below 1.0.  The coarse fraction generally has
14     one mode within the range ~ 5 to 50 /xm. The particle volume frequency functions for the
15     fine and coarse fractions often behave independently, (i.e., vary in relative proportion of the
16     total ambient particle mix from location to location or from one time or season to another at
17     the same location).
18           Previous documentation has shown that  hydroxy, hydroperoxy,  and alkoxy  radicals are
19     probably important in the oxidation of SO2 to SO3", although the rate constants for some of
20     these reactions are not well established. The hydroxy radical dominates the gas-phase
21     oxidation of SO2 in the clean troposphere, and H2O2 is effective in the  formation of SO42~  in
22     particles, mists, fogs, and rain.  Transition metals and soot have been shown to be effective
23     catalysts for atmospheric oxidation of SO2. Oxidation rates for NO  and NO3~ are known but
24     have  been considered too low to be important.  The oxidation rate for NO2~ is known, but
25     the tropospheric concentration of HNO2 is probably too low for this reaction to be important.
26     Except for reactions of carbon (soot), solid surface reactions do not appear to be effective
27     pathways for H2SO4 formation in the troposphere.
28           The physical properties of particles are  physical configuration, bulk material properties,
29     and surface properties. The bulk material properties that affect aerosol behavior include
30     chemical composition, vapor pressure, hygroscopicity and deliquescence, and index of
31     refraction.  These properties control the physical state and growth of particles and result in

       April 1995                               2-18       DRAFT-DO NOT QUOTE OR CITE

-------
 1     scattering and absorption of light by tropospheric particles.  Hygroscopicity, deliquescence,
 2     and efflorescence are critical properties in the growth of particles, but there is a paucity of
 3     thermodynamic data to permit prediction of deliquescence and hygroscopic behavior and
 4     vapor pressures of multicomponent systems, especially for relative humidities below about
 5     90%.  Few studies  of desorption under atmospheric conditions have been reported.  Of
 6     greater concern, desorption may prove to be important in biological systems.  Shape,
 7     structure, and density are physical configuration properties that are important parameters in
 8     the equations of motion for particles.  Because of irregularities in particle geometry or
 9     because the particles are agglomerates, the three  configuration properties are usually defined
10     in terms of an aerodynamic diameter.  Surface properties of importance include electrostatic
11     charging, adhesion,  and the influence  of surface films.
12          The physical properties of particles and their modal distributions are important
13     considerations (1) in the sampling and analysis of atmospheric particles and (2) in predicting
14     or determining the flux to biological and nonbiological materials and deposition in the human
15     and experimental animal respiratory tracts.
16          Advances in understanding  the properties and behavior of atmospheric paniculate matter
17     have been made since publication of the previous criteria document (U.S. Environmental
18     Protection Agency,  1982).  In the current revision of the document, newer literature and data
19     on the above topics  are reviewed and  discussed.  For example, chemical pathways and rates
20     of atmospheric particle formation and  of removal from the atmosphere, by dry deposition and
21     by precipitation scavenging, are examined.  Likewise, the physical processes of nucleation,
22     condensation, and coagulation by which condensible material is converted into particles are
23     discussed, along with the size distribution of the  resulting particles.  The physical properties
24     relevant to sampling considerations and deposition on surfaces, including those of the
25     respiratory tract, are also discussed, including coverage of several newer areas of expanded
26     research: aerosol equilibria, the unique properties of semi-volatile aerosols, and the role of
27     particle-bound water.
28
29     Measurement Methodology
30          Techniques available for measurement of mass and specific components of aerosols are
31     examined.  Special attention is given to the suitability of current technology for the

       April 1995                               2-19      DRAFT-DO NOT QUOTE OR CITE

-------
 1     measurement of aerosol mass with sufficient accuracy and precision to determine compliance
 2     with one or another possible type of a new PM standard (i.e., a PM10 standard with a lower
 3     level or a fine-particle standard). The need for continuous or at least daily PM
 4     measurements, the  difficulty of removing particle-bound water without losing NH4NO3 or
 5     semivolatile organic matter, problems in defining and maintaining a precise cut at 10 ^im or
 6     lower (e.g., at 2.5  pirn), and techniques for maintaining good quality control in monitoring
 7     networks are also addressed.
 8
 9     Ambient Levels
10          The present draft of the revised PM AQCD describes ambient PM data for the United
11     States, with characterization as available by size (fine/coarse) and chemical composition.
12     Data that focus on  the current U.S. PM10 standard are emphasized, but information is also
13     provided on PM2 5, PM2 5.10, and other similar cut points,  as data  are available. Ambient
14     patterns are discussed, to  include daily, seasonal, regional, etc. Acid aerosol data are also
15     described as above as a separate aspect  of PM.  Key questions addressed include: What
16     information is available on the distribution  of PM in regard to:  geographic, seasonal,
17     diurnal, size, composition, sources, and trends?   How important are uncertainties introduced
18     by variations in the position and shape of the 10-/mi cut point in various PM10 monitors?
19     How important are measurement uncertainties due to volatilizable/condensible components
20     (e.g., loss of ammonium nitrate and, possibly, other ammonium salts) and loss of
21     semivolatile organics  or retention of particle-bound water?  How do  these uncertainties vary
22     geographically and seasonally?   How do these uncertainties differ for filter collection and
23     subsequent weighing as compared to continuous indicators?
24
25     Cut Points
26          Information helpful in evaluating the  possible need for a new particle  standard (PM-
27     Fine) in addition to or instead of a PM10 NAAQS is presented. This information includes
28     discussion of sources, composition, lung deposition,  sampling problems, epidemiology,
29     biochemistry, and toxicology of fine and coarse particles.  Other considerations include
30     techniques for separating fine particles from coarse particles. Can fine and coarse particles
31     be separated adequately by  a single size cut-point in all areas of the country or will  the

       April 1995                               2-20      DRAFT-DO NOT  QUOTE OR CITE

-------
  1      optimal cut point differ in time and space, especially between very dry areas where coarse
  2      particles may be found below 2.5 /urn and very humid areas where fine particles may be
  3      found above 1.0 /xm?  If a single fine-particle cut point is chosen, which is best:  2.5, 1.0, or
  4      something in between?  Is size  an adequate separation or will chemical composition
  5      measurements also be needed?
  6
  7      Exposure
  8           Particulate matter exposure estimates for most epidemiology studies are based on data
  9      from ambient monitoring sites.  Relationships between such measurements and personal
 10      exposure are important in evaluating epidemiology studies.   Aspects evaluated  and discussed
 11      in the present draft document include:  urban scale PM exposure models, indoor/outdoor
 12      characteristics and relationships, and the validity of ambient measurements to provide
 13      appropriate estimates to relate to health effect endpoints.  Two exposure estimates are of
 14      concern, individual and population estimates of PM exposure.  The type of epidemiology
 15      study determines which estimate is appropriate.  Additionally, other factors (such as exposure
 16      durations) that may determine health effects are considered.  Human exposure patterns to
 17      ambient and indoor air particles, including consideration of activity patterns and various
 18      microenvironments, are characterized.
 19           Actual human exposure differs from outdoor concentrations due to infiltration of
 20      ambient aerosols indoors, indoor sources, and human activity patterns.  Human exposure can
 21      be determined through measurements and models.  For PM the indoor and personal
 22      monitoring data show both higher than ambient and lower than ambient PM concentrations in
 23      indoor  settings as a function of  varying particle size and human activity patterns.
 24           Coarse-mode particles (>2.5 ^m), which are generally of nonanthropogenic  origin
 25      (windblown dust, etc.), require  turbulence to  provide vertical velocity components greater
26      than their settling velocity to  allow them to remain suspended in the air.  Outdoor  particles
27      enter into an indoor setting either by bulk flow (e.g., through an open window) in which all
28      particles can enter at the inlet condition, or by diffusional flow (e.g., through cracks and
29      fissures in the barrier of the building envelope) in which velocities are relatively lower and
30      therefore capable of settling out the coarser particles in the passage through the barrier.   Fine
       April 1995                                2-21       DRAFT-DO NOT QUOTE OR CITE

-------
 1     particles, however, are not easily removed by settling or impaction and penetrate indoors
 2     with high efficiency.
 3           Indoor settings are usually quiescent, and the larger ambient particles that do enter
 4     indoors quickly settle out, leading to the presence of the familiar dust layers that require
 5     indoor settings to be cleaned constantly. However, human activity in indoor settings does
 6     generate fine particles (<2.5 /xm) from smoking, vacuuming, cooking, etc., and resuspends
 7     coarse particles that previously had settled out.
 8           Two major factors influencing the relationship of ambient to indoor PM air quality are
 9     (1) the variability of indoor concentrations of PM compared to outdoor concentrations  as a
10     function of particle size (e.g., fine indoor  > fine outdoor and coarse indoor  <  coarse
11     outdoor) and (2) the variation of exposures of individuals related to the different activities
12     that are involved with the local generation of particles in their immediate surroundings
13     (smoking, traffic, dusting and vacuuming at home, etc.).
14           Long-term personal exposures  to coarse-fraction PM (>2.5 pirn) can be less than half
15     the ambient concentrations.  Long-term personal  exposures to fine-fraction PM (<2.5  jum) of
16     ambient origin may be estimated by ambient measurements of the <2.5-^tm PM fraction.
17     However,  personal activities and indoor concentrations cause personal exposures to PM to
18     vary substantially raising the issue of how well ambient measurements can  serve as predictors
19     of human exposures, either on an individual personal level or on a community-wide level.
20
21     2.4.2   Health Effects
22           A rapidly growing body of epidemiologic data examines relationships between PM
23     concentrations and  human health effects, ranging from respiratory function changes and
24     symptoms to exacerbation of respiratory disease and excess mortality associated with
25     premature death.  These effects appear to lie along an increasing gradient of severity of
26     effects in different  subpopulations.   Although the exact biological mechanisms underlying
27     such effects are poorly understood, there seems to be an emerging pattern of findings that
28     increases the plausibility that the observed relationships may  reflect a real,  causal relationship
29     between paniculate matter and human health. This revised PM  criteria document assesses
30     evidence suggesting that this overall pattern of effects may extend to concentrations  of PM10
31     below the current NAAQS or may be  associated  with other PM  size fractions (e.g., fine

       April 1995                                2-22      DRAFT-DO NOT QUOTE OR CITE

-------
  1      particles <  2.5 ^m).  Controlled human exposure and animal toxicologic studies are also
  2      evaluated, and the overall coherence/consistency of findings in relationship  to the
  3      epidemiologic database is assessed.  These include, for example:  (1) studies of respiratory
  4      tract disposition and clearance of particles, (2) experimental studies (animal and human)
  5      evaluating mechanisms of action of various particles (by size, chemical composition, etc.) in
  6      order to evaluate  biological plausibility of effects reported by epidemiology studies, and (3)
  7      other experimental studies that demonstrate various toxic  effects of PM in humans  or in
  8      animal models.
  9
10      Respiratory  Tract Dosimetry
11           The biological  endpoint or health effect of an aerosol exposure is likely more directly
12      related to the quantitative pattern of deposition within the respiratory tract than just to the
13      external exposure concentration.  The regional deposition pattern determines not only the
14      initial respiratory tract dose but  also the specific pathways and rates by which the inhaled
15      material is cleared and redistributed.  Thus, in order to evaluate different toxic responses to
16      inhaled particles across species and to accurately extrapolate such laboratory animal data to
17      humans, or to evaluate differences that sex, age, or disease may have on human variability,
18      the various physicochemical, anatomic, and physiologic factors described must be integrated
19      to estimate a deposited dose or perhaps a retained dose (deposition — clearance =  retention).
20      Delineation of the dose to each respiratory tract region (extrathoracic, tracheobronchial, and
21      pulmonary) is desired because each region has different dominant factors controlling
22      deposition and clearance, and different defense mechanisms. A theoretical model to describe
23      particle deposition and clearance would require detailed information on all the influential
24      parameters mentioned above (e.g., respiratory  rates, exact airflow patterns,  complete
25      measurements of the branching structure of the respiratory tract,  pulmonary region
26      mechanics) for men, women, children, and across the various species used in toxicity
27      studies. An empirical  model (i.e., equations fit to experimental data) may adequately
28      describe regional  deposition and require much  less data to develop the model structure.
29           Within the dosimetry chapter (Chapter 10) of this document, the anatomy of the
30      respiratory tract and  the physicochemical, anatomical, and physiological  factors controlling
31      particle deposition, clearance,  and retention are reviewed.  Other factors that modify

        April 1995                                 2-23       DRAFT-DO NOT QUOTE OR CITE

-------
 1      deposition,  including sex, age, disease state, and exposure to irritants also are discussed.
 2      The available human and laboratory data on deposition and clearance and their positive and
 3      negative attributes for use in quantitative model development are discussed.  Available
 4      validated model structures to estimate deposition and clearance in humans and laboratory
 5      animals are described and evaluated.  The application of these models to quantitative
 6      extrapolation of the human and animal toxicity data also  are discussed.  Consideration is
 7      given to uncertainties in input parameters and  the variability of model predictions when
 8      evaluating the usefulness of models for quantitative dose  extrapolation.
 9
10      Epidemiology Studies
11           Epidemiologic analyses are expected to provide some of the most crucial  information
12      useful in deriving health criteria upon which to base Agency decisions regarding possible
13      revision of the current PM  standards, and such studies are accorded extensive attention in
14      this PM criteria document as evaluated in Chapter 12 and elsewhere.
15           One useful distinction is to separate short- and long-term effects. The short-term
16      effects include changes in respiratory function, symptom indicators,  hospital admissions
17      associated with exacerbation of respiratory or  cardiovascular disease, and excesses of daily
18      death rates  in urban areas associated with concurrent 24-h PM measurements on the same or
19      preceding few days.   The short-term effects studies are typically longitudinal in nature and
20      are specific to a  community or metropolitan area with reasonably homogeneous PM
21      exposures.  The  analyses of data in short-term studies use time-series analysis methods.  The
22      long-term or chronic exposure effects studies typically use annual PM concentrations and
23      annual symptom or death rates and are more likely to involve comparisons across several
24      communities rather than within a single community. Although both kinds of epidemiologic
25      analysis are useful, it is important  to assess the consistency of conclusions based on different
26      kinds of studies.  Coherence of effects at lower concentrations is a useful criterion for
27      assessing diverse studies with different endpoints or effects, different populations, and
28      different exposure  metrics (Bates et al., 1990) and is considered as part of the  evaluation of
29      the available epidemiology  literature.
30
        April 1995                                2-24       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Mortality Studies. Studies examining the relationship between ambient measures of PM and
  2      mortality were examined during the last review process (U.S. Environmental Protection
  3      Agency, 1982,  1986) and contributed to the key scientific bases underlying the current PM10
  4      NAAQS.  However, given the uncertainties in converting from British Smoke to PM10
  5      measurements, particularly at lower concentrations, and the possible differences in paniculate
  6      composition between London at the time the data were gathered and the contemporary United
  7      States, it was difficult to determine a precise level for a relationship between PM10 and
  8      mortality.   Since that time, numerous contemporary U.S. mortality studies using either PM10
  9      or TSP measurements have been published that examine short-term measurements.  Also,
 10      long-term PM ambient measurements and mortality have been examined in some recent
 11      studies. These  and other newly emerging PM-mortality studies are summarized and critically
 12      evaluated.
 13           Issues of greatest concern so  far relate primarily to the use and interpretation of the
 14      short-term mortality studies.  Almost all analyses of the relationship between PM and excess
 15      mortality require statistical  adjustment for mortality excesses associated with other potential
 16      confounding factors,  including other environmental stressors such as temperature and relative
 17      humidity or other pollutants (co-pollutants) associated with PM and with mortality.  For
 18      example, weather-related effects may be directly related to excess mortality, but may also be
 19      indirectly related when weather affects PM emissions and atmospheric concentrations.
 20      Statistical and conceptual approaches to estimating the direct and indirect effects of
 21      confounding variables,  and specification of the form of the statistical adjustment for
 22      confounding factors are evaluated in interpreting the PM effects on mortality calculated from
 23      each study.  Comparison of studies using different exposure metrics is considered.  In
 24      characterizing the relationship between excess mortality and PM in  different cities, evaluation
 25      of differences in particle size distribution or particle composition between cities, is done as
 26      the data allow.
27           Specification of "exposure-effect" relationship(s)  between mortality and PM is also
28      important.   A number of studies have reported no evident threshold for effects, even at
29      relatively low concentrations,  but the ability to carry out meaningful threshold evaluations
30      may be greatly limited by the power of the various available studies.  Estimates of the
31      relationship between PM and mortality may depend on differences in model specification.

        April 1995                                2-25      DRAFT-DO NOT QUOTE OR CITE

-------
 1      Even with similar model specifications (exposure-response relationship, adjustment for
 2      weather, copollutants, and other factors) there may be differences in the effects of PM at a
 3      given concentration, possibly related to particle size/composition and/or differences in
 4      climate or demographics among different cities.  An important component of the health
 5      effects assessment in the criteria document is identification of susceptible subpopulations and
 6      other variables such as weather, climate,  or other pollutants, potentially contributing to
 7      increased mortality risk.
 8
 9      Morbidity Studies.  Decreased pulmonary function in predominantly healthy children was
10      been reported in some earlier epidemiology studies. More recent studies add to this
11      database.  Earlier long-term exposure studies provided no evidence for an effect from PM
12      exposure on level of pulmonary function, whereas some recent studies report reductions in
13      pulmonary function associated with chronic exposure to paniculate pollution.  An evaluation
14      of the epidemiologic database relating short-term (24-h) and long-term (annual) ambient
15      measurement of PM10 and other measures of PM to changes in pulmonary function test
16      results in children and adults is presented. The strength and consistency of epidemiologic
17      databases that relate short-term (24-h) and long-term (annual) PMjQ and other ambient PM
18      indicator measurements  to changes in the rate and/or severity of respiratory symptoms and
19      disease are also critically reviewed.  Studies examining exacerbation of respiratory (i.e.,
20      COPD and asthma) and cardiovascular diseases that lead to increased medical care utilization
21      (such as emergency room visits and hospital admissions) in relation to ambient PM exposure
22      are also evaluated.  As appropriate, other factors and copollutants are  also examined  in
23      relation to findings on each of the above types of health endpoints.
24
25      Toxicology of PM Constituents
26           In addition to the chapter evaluating epidemiologic studies of PM differentiated mainly
27      in terms of various size indicators (TSP,  PM10, etc.), the toxicology of various major
28      subclasses of PM constituents  is also summarized and discussed in a separate chapter. That
29      toxicology chapter focuses on  acid aerosols, metals, ultrafine particles, diesel particles,
30      silica, and other types of particles that make up ambient air mixes of particles  in the  broad
31      class designated in toto as "particulate matter".  Animal inhalation toxicology and other types

        April  1995                               2-26      DRAFT-DO NOT QUOTE OR CITE

-------
  1      of studies reviewed are useful in improving understanding of several key overall health
  2      issues, especially:  (1) the influence of particle size, number, and mass on health responses;
  3      (2) the differential influence of varying particle chemistry on the health effects observed;
  4      (3) the array of health effects that can be caused by specific PM constituent; (4) exposure-
  5      response relationships for various exposure durations (acute and chronic); (6) mechanisms of
  6      toxicity; and (6) pollutant interactions.  Some of the information from these studies provides
  7      a background for evaluating the biological plausibility of the mortality and morbidity
  8      associations reported in epidemiological studies.  For example, whether chronic bronchitis
  9      can be caused by sulfuric acid exposure, as hypothesized from animal studies, is a significant
10      issue.  The data on the relationship between particle size, mass, and number elucidates the
11      appropriateness of various exposure indicators of potential human effects.  The document
12      reviews toxicological studies examining hypotheses related to health outcome  and the physical
13      and chemical characterizations of PM.  For chemistry, this includes acidity, surface coatings
14      (i.e., soluble metals),  and particle-bound organics. Particle size is examined  (PM10,  PM2 5,
15      other [fine versus ultrafine]).  Aerosol concentration examines particle number and mass.
16           Evaluation of the controlled human exposure (clinical) studies database concerning PM
17      and health outcomes is presented as a subsection of the overall PM constituent toxicology
18      chapter. This includes critical review of PM effects on pulmonary function in healthy and
19      asthmatic individuals,  pulmonary clearance mechanisms, airway reactivity, and immunologic
20      defense especially in relation to particle size but only to a limited extent in relation to
21      chemical composition.  There remains an almost complete absence of controlled  experiment
22      data on exposure of humans to particles other than acid aerosols.
23           Human clinical studies of PM constituents have been almost completely  limited to
24      measuring effects on symptoms, lung function, and airway reactivity, in addition to a few
25      studies of effects on mucociliary clearance.  Few have used bronchoalveolar lavage to study
26      effects on airway inflammation and host defense; nor have many, if any, examined effects of
27      acid aerosols or other  particle exposures on airway inflammation in asthmatic subjects or on
28      exacerbation of effects of antigen challenge in allergic or asthmatic subjects.
29
30
        April 1995                                2-27      DRAFT-DO NOT QUOTE OR CITE

-------
 1      Sensitive Groups
 2           Available data are also evaluated for insight concerning human population groups
 3      potentially having increased susceptibility to PM exposure.  Preexisting respiratory or
 4      cardiovascular disease, in conjunction with advanced age, appear to be important factors in
 5      PM mortality susceptibility.  For morbidity health endpoints, children and asthmatic
 6      individuals potentially may display increased sensitivity to PM exposure,  and, as such, this
 7      topic is discussed.
 8
 9      2.4.3   Welfare Effects
10      Effects on Materials
11           All manmade materials exposed to the outdoor environment undergo degradation by
12      heat, moisture, and some bacteria and fungi.  For many years, air pollution has been
13      suspected of accelerating the natural degradation processes.  For example, acidic pollutants
14      have been associated with accelerated degradation of paints  such as water-based paint  and
15      alkyd coatings containing titanium dioxide, lead minium, or ferric oxide red.  Other
16      researchers  have reported acidic pollution-related effects on automotive paint and steel
17      coating. Particulate matter has also been reported to produce paint soiling. Also, acid
18      aerosols and other particles containing acids also have been reported to affect building stones,
19      cement, and concrete.  Acidic aerosols  change the physical characteristics of some stones,
20      cement, and concrete by changing the chemical composition.  Studies  examining the effects
21      on materials of PM pollution (primary and secondary particles and aerosol precursor gases)
22      are reviewed and summarized; where possible, changes  in material damage are correlated
23      with changes in PM concentrations.
24
25      Visibility Effects
26           Airborne PM in the form of varying amounts of sulfates, ammonium and nitrate ions,
27      elemental carbon and organic carbon compounds, water and smaller amounts of soil dust,
28      lead compounds, and other trace species reduce visibility, thereby affecting transportation
29      safety and creating a loss in aesthetic appeal. The natural background visibility range is 150
30      ±  45 kilometers for the east and 230 ± 35 kilometers for the west.  When current visibility
31      data are compared to background visibility data, manmade contributions account for

        April 1995                                 2-28       DRAFT-DO  NOT QUOTE OR CITE

-------
  1      approximately one-third of the average extinction coefficient in the rural west and over 80%
  2      of the average extinction coefficient in the rural east.  The effects of aerosol concentration,
  3      composition, and size and pollutant emission trends on visibility are evaluated. Existing PM
  4      models are discussed in relation to how well such models can be used to predict changes  in
  5      visibility.
  6
  7      Climate Change
  8           It has been suggested that fine particles released into the atmosphere may alter the
  9      climate through a reduction in the amount of solar radiation reaching the earth's surface,  thus
 10      cooling the surface while heating the aerosol layer.  The scattering and absorbing properties
 11      of aerosols and their vertical distribution are briefly reviewed and reference made to other
 12      assessments of their effects on radiative balance and how changes in radiative balance may
 13      affect weather and climate.  Aerosols also affect weather and climate through their role as
 14      cloud condensation nuclei.  The concentration, composition,  size,  and number of aerosols can
 15      influence the structure,  stability, and albedo of clouds, possible changing the location and
 16      amount of  rainfall and the rate of global and regional warming due to greenhouse gases.
 17
 18
 19      2.5   DOCUMENT CONTENT AND ORGANIZATION
 20           The present document includes review and critical evaluation of relevant scientific
 21      literature on PM through early 1995.  The material selected for review and comment in the
 22      text generally comes from the more recent literature published since 1982, with emphasis on
 23      studies conducted at or  near PM pollutant concentrations found in ambient air.  Older
 24      literature cited in the previous criteria document for PM and Addendum (U.S. Environmental
 25      Protection Agency, 1982,  1986) is generally not discussed.  However, as appropriate, some
 26      limited discussion is included of older studies judged to be significant because of their
 27      potential usefulness in deriving a NAAQS.  An attempt has been made to discuss  key
28      literature in the text and present it in tables as  well.  Reports of lesser importance for the
29      purposes of this document  are typically only summarized in tables.
30           Generally, main emphasis is placed on consideration of published material that has
31      undergone scientific peer review. In the interest of admitting new and important

        April 1995                                2-29       DRAFT-DO NOT QUOTE  OR CITE

-------
 1      information, however, some material not yet published in the open literature but meeting
 2      other standards of scientific reporting may be included as reviewed by CAS AC.  Emphasis
 3      has been placed on studies in the range of current ambient levels. On this basis, studies in
 4      which the lowest concentration employed exceeded this level have been included if they
 5      contain unique data, such as documentation of a previously unreported effect or of
 6      mechanisms of effects, or if they were multiple-concentration studies designed to provide
 7      information on concentration-response  relationships.  Results of studies conducted at higher
 8      levels have  been included because of the potential importance of these effects to public
 9      health. In reviewing and summarizing the literature, an attempt  is made to present
10      alternative points of view where scientific controversy  exists.  As warranted, considerations
11      bearing on the quality of studies are noted.
12           The present document consists of 13 chapters.  The Executive Summary for the entire
13      document is contained in Chapter  1, followed by this general  introduction in Chapter 2.
14      Chapters 3 through 7 provide background information on physical and chemical properties of
15      PM and related compounds; sources and emissions; atmospheric  transport, transformation,
16      and fate of PM; methods for the collection and measurement of PM; and ambient air
17      concentrations and factors affecting exposure of the general population.  Chapter 8 describes
18      effects on visibility, and Chapter 9 describes damage to materials attributable to PM.
19      Chapters 10 through 13 evaluate information concerning the health effects of PM.  More
20      specifically, Chapter  10 discusses  dosimetry of inhaled particles  in the respiratory tract and
21      Chapter 11  summarizes information on the toxicology of specific types of PM constiuents,
22      including experimental toxicological studies of animals and human clinical studies.
23      Chapter 12  discusses epidemiological studies and Chapter 13 characterizes information on
24      critical health issues derived from studies reviewed in the prior chapters.
25           Neither control techniques nor control  strategies for the abatement of PM are discussed
26      in this document, although some topics covered may be incidentally relevant to abatement
27      strategies.  Technologies for controlling PM emissions are discussed in other documents
28      issued by EPA's Office of Air  Quality Policy and Standards (OAQPS).  Likewise,  issues
29      germane to the scientific basis  for control strategies, but not pertinent to the development of
30      criteria, are addressed in numerous other documents issued by OAQPS.
        April 1995                                2-30       DRAFT-DO NOT QUOTE OR CITE

-------
  1           In addition, certain issues of direct relevance to standard setting are not explicitly
  2      addressed in this document, but are instead analyzed in documentation prepared by OAQPS
  3      as part of its regulatory  analyses materials. Such analyses include (1) discussion of what
  4      constitutes an "adverse effect" and delineation of particular adverse effects  that the primary
  5      and secondary NAAQS are intended to protect against, (2) exposure analyses and assessment
  6      of consequent risk, and (3) discussion of factors to be considered in determining an adequate
  7      margin of safety.  Key points and conclusions from such analyses are summarized in a Staff
  8      Paper prepared by OAQPS and reviewed by CAS AC. Although scientific data contribute
  9      significantly to decisions regarding the above issues, their resolution cannot be achieved
10      solely on the basis of experimentally acquired information.  Final decisions on items (1) and
11      (3) are made by the Administrator, as mandated by the CAA.
12           A fourth issue directly pertinent  to standard setting is identification of populations at
13      risk, which is basically a selection by  EPA of the subpopulation(s) to be protected by the
14      promulgation of a given  standard.  This issue is addressed  only partially in this document.
15      For example,  information is presented on factors, such as preexisting disease, that may
16      biologically predispose individuals and subpopulations to adverse effects from exposures to
17      PM. The identification of a population at risk, however, requires information above and
18      beyond data on biological predisposition, such as information on levels of exposure, activity
19      patterns, and personal habits.   Such information is included in the Staff Paper developed by
20      OAQPS and reviewed by CAS AC as a separate item from  this document.
       April 1995                                2-31       DRAFT-DO NOT QUOTE OR CITE

-------
 1      REFERENCES
 2
 3      Bates, D. V.; Baker-Anderson, M.; Sizto, R. (1990) Asthma attack periodicity: a study of hospital emergency
 4             visits in Vancouver. Environ. Res. 51: 51-70.
 5
 6      Beloin, N. J.; Haynie, F. H. (1975) Soiling of building materials. J. Air Pollut. Control Assoc. 25: 399-403.
 7
 8      Bouhuys, A.; Beck, G. J.; Schoenberg, J. B. (1978) Do present levels of air pollution outdoors affect respiratory
 9             health? Nature (London) 276: 466-471.
10
11      Braun-Fahrlander, C.; Ackermann-Liebrich,  U.; Schwartz, J.; Gnehm, H. P.; Rutishauser, M.; Wanner, H. U.
12             (1992) Air pollution and respiratory symptoms in preschool children. Am. Rev. Respir. Dis. 145: 42-47.
13
14      Brunekreef, B.; Dockery, D. W.; Speizer, F. E.; Ware, J. H.; Spengler, J. D.; Ferris, B. G. (1989) Home
15             dampness and respiratory morbidity in children. Am. Rev.  Respir. Dis. 140: 1363-1367.
16
17      Brunekreef, B.; Kinney, P.  L.; Ware, J. H.; Dockery, D.; Speizer, F. E.; Spengler, J. D.; Ferris, B. G., Jr.
18             (1991) Sensitive subgroups and normal variation in pulmonary function response to air pollution episodes.
19             Environ. Health Perspect. 90: 189-193.
20
21      Chestnut, L. G.; Schwartz,  J.;  Savitz, D. A.; Burchfiel, C. M. (1991) Pulmonary function and ambient
22             paniculate matter: epidemiological evidence from NHANES I. Arch. Environ. Health 46:  135-144.
23
24      Code of Federal  Regulations. (1986) Appendix B—reference method for the determination of suspended
25             paniculate matter in the atmosphere (high-volume  method). C. F. R. 40: § 50.
26
27      Dassen, W.; Brunekreef, B.; Hoek, G.; Hofschreuder, P.; Staatsen, B.; De Groot, H.; Schouten,  E.; Biersteker,
28             K. (1986) Decline in children's pulmonary function during  an air pollution episode. J. Air Pollut. Control
29             Assoc. 36: 1223-1227.
30
31      Dockery, D. W.; Ware, J. H.; Ferris, B. G., Jr.;  Speizer, F. E.; Cook, N. R.; Herman, S. M. (1982) Change
32             in pulmonary function in children associated with air pollution episodes. J.  Air Pollut. Control Assoc.
33             32: 937-942.
34
35      Dockery, D. W.; Speizer, F. E.; Stram, D. O.; Ware, J.  H.; Spengler, J. D.; Ferris, B. G., Jr. (1989) Effects
36             of inhalable particles on respiratory health of children. Am. Rev.  Respir. Dis. 139: 587-594.
37
38      Dockery, D. W.; Schwartz, J.; Spengler, J. D. (1992) Air pollution and daily mortality: associations with
39             particulates and acid aerosols. Environ. Res. 59: 362-373.
40
41      Dockery, D. W.; Pope,  C.  A., Ill;  Xu, X.; Spengler, J. D.; Ware, J. H.; Fay, M. E.; Ferris, B.  G., Jr.;
42             Speizer,  F. E. (1993) An association between air pollution and mortality in six U.S. cities. N. Engl. J.
43             Med. 329: 1753-1759.
44
45      Edney, E. O.; Cheek, S. F.; Stiles, D. C.; Corse, E. W.;  Wheeler, M. L.; Spence, J. W.; Haynie, F. H.;
46             Wilson, W. E. (1988) Effects of acid deposition on paints and metals: result of a controlled field study.
47             Atmos. Environ. 22: 2263-2274.
48
49      Edney, E. O.; Cheek, S. F.; Corse, E.  W.; Spence, J.  W.; Haynie, F. H. (1989) Atmospheric weathering
50             caused by dry deposition of acidic species. J.  Environ. Sci. Health Part A 24: 439-457.
51
52      Federal Register. (1971) National primary and secondary ambient air quality standards. F. R. (April 30)
53             36: 8186-8201.
54

         April  1995                                     2-32       DRAFT-DO NOT QUOTE  OR CITE

-------
  1     Federal Register. (1979a) National ambient air quality standards; review of criteria and standards for paniculate
  2             matter and sulfur oxides. F. R. (October 2) 44: 56730-56731.
  3
  4     Federal Register. (1979b) National primary and secondary ambient air quality standards: revisions to the national
  5             ambient air quality standards for photochemical oxidants. F.  R. (February 8) 44: 8202-8221.
  6
  7     Federal Register. (1984) Proposed revisions to the national ambient air quality standards for paniculate matter. F.
  8             R. (March 20) 49: 10408-10435.
  9
 10     Federal Register. (1986a) National ambient air quality standards; review of criteria and standards for paniculate
 11             matter and sulfur oxides. F. R. (April 1) 51:  11058.
 12
 13     Federal Register. (1986b) National ambient air quality standards: review of criteria and standards for paniculate
 14             matter and sulfur oxides. F. R. (July 3) 51: 24392-24393.
 15
 16     Federal Register. (1987) Revisions to the national ambient air quality standards for paniculate matter. F. R.
 17             (July 1) 52: 24634-24669.
 18
 19     Federal Register. (1994) Air quality criteria for paniculate matter. F. R. (April 12) 59: 17375.
 20
 21     Ferris, B. G., Jr.;  Higgins, I. T. T.; Higgins, M. W.; Peters, J. M. (1973) Chronic nonspecific respiratory
 22             disease in Berlin, New Hampshire, 1961 to 1967: a follow-up study. Am. Rev. Respir. Dis.
 23             107: 110-122.
 24
 25     Ferris, B. G., Jr.;  Chen, H.; Puleo, S.; Murphy, R.  L. H., Jr. (1976) Chronic nonspecific respiratory disease in
 26             Berlin, New Hampshire, 1967 to 1973:  a further follow-up study. Am. Rev. Respir. Dis. 113: 475-485.
 27
 28     Gauri, K. L.  (1979) Effect of acid rain on structures. In:  Acid rain: a session at the national convention; April;
 29             Boston, MA.  New York, NY: American Society of Civil Engineers, Environmental Impact Analysis
 30             Research Council; pp. 55-75.
 31
 32     Hausman, J. A.; Ostro, B. D.; Wise,  D.  A. (1984) Air pollution and lost work. Cambridge, MA: National
 33             Bureau of Economic Research; NBER working paper no. 1263.
 34
 35     He, Q.-C.; Lioy, P. J.; Wilson, W. E.; Chapman, R. S.  (1993) Effects of air pollution on children's pulmonary
 36             function in urban and suburban areas of Wuhan, People's Republic of China. Arch. Environ. Health
 37             48:382-391.
 38
 39     Hoek, G.; Brunekreef, B. (1993) Acute effects of a winter air pollution episode on pulmonary function and
 40             respiratory  symptoms of children. Arch. Environ. Health 48: 328-335.
 41
 42     International Standards Organization. (1981) Size definitions for particle sampling: recommendations of ad hoc
 43            working group appointed by Committee TC 146 of the International Standards Organization. Am. Ind.
 44            Hyg. Assoc. J. 42: A64-A68.
45
46     Johnson, J. B.; Elliott, P.; Winterbottom, M. A.; Wood,  G. C. (1977) Short-term atmospheric corrosion of mild
47            steel at two weather and  pollution monitored sites. Corros. Sci. 17:  691-700.
48
49     Johnston-Feller, R.; Osmer, D. (1977) Exposure evaluation: quantification of changes in appearance of pigmented
50            materials. J. Coat. Technol. 49: 25-36.
51
52     Knotkova-Cermakova, D.; Vlckova, J.; Honzak, J. (1982) Atmospheric weathering steels. In:  Dean, S. W.;
53             Rhea, E.  C., eds. Atmospheric corrosion of metals. Philadelphia, PA: Aerican Society for Testing and
54            Materials; special technical publication 767.


         April  1995                                     2-33        DRAFT-DO NOT QUOTE  OR CITE

-------
  1      Lampa, K.; Saarnak, A. (1986) Acid rain test: short term test on paint systems for corrosion protection. Farbe
  2             Lacke 92: 692-696.
  3
  4      Lawther, P. J. (1986) [Letter to John Bachmann]. Washington, DC: Office of Air Quality Planning and
  5             Standards; August 22. Available for inspection at: U.S. Environmental Protection Agency, Central
  6             Docket Section, Washington, DC; docket no. A-82-37, IV-D-319.
  7
  8      Lawther, P. J.; Waller, R. E.;  Henderson, M. (1970) Air pollution and exacerbations of bronchitis.  Thorax
  9             25: 525-539.
10
11      Lippmann, M. (1986a) [Letter to EPA Administrator Lee Thomas]. Washington, DC: U.S. Environmental
12             Protection Agency, Clean Air Scientific Advisory Committee; January 2.  Available for inspection at:
13             U.S. Environmental Protection Agency, Central Docket Section, Washington, DC; docket no. A-82-37,
14             IV-D-315.
15
16      Lippmann, M. (1986b) [Letter to EPA Administrator Lee Thomas]. Washington, DC, U.S. Environmental
17             Protection Agency, Clean Air Scientific Advisory  Committee; December 15. Available for inspection at:
18             U.S. Environmental Protection Agency, Central Docket Section, Washington, DC; docket no. A-82-37,
19             IV-D-339.
20
21      Lippmann, M. (1986c) [Letter to EPA Administrator Lee Thomas]. Washington, DC: U.S. Environmental
22             Protection Agency, Clean Air Scientific Advisory  Committee; December 16. Available for inspection at:
23             U.S. Environmental Protection Agency, Central Docket Section, Washington, DC; docket no. A-82-37,
24             IV-D-338.
25
26      Malm, W.  C.; Bell, P.; McGlothin,  G. E. (1984) Field testing a methodology for assessing the importance of
27             good visual quality. Presented at: the 77th annual meeting of the Air Pollution Control Association; June;
28             San Francisco,  CA. Pittsburgh, PA: Air Pollution Control Association.
29
30      Mazumdar, S.; Schimmel, H.; Higgins, I. T.  T. (1982) Relation of daily mortality to air pollution: an analysis of
31             14  London winters, 1958/59-1971/72. Arch. Environ. Health 37: 213-220.
32
33      Nriagu, J. O. (1978) Deteriorative effects of sulfur pollution on materials. In: Nriagu, J. O., ed.  Sulfur in the
34             environment: part II, ecological effects. New York, NY: Wiley and Sons.
35
36      Ostro, B. (1984) A search for a threshold in the relationship of air pollution to mortality: a reanalysis of data on
37             London winters. Environ. Health Perspect. 58: 397-399.
38
39      Ostro, B. D. (1987) Air pollution and morbidity revisited: a specification test. J. Environ. Econ.  Manage.
40             14: 87-98.
41
42      Ostro, B. D.; Lipsett, M. J.; Wiener, M.  B.;  Seiner, J. C. (1991) Asthmatic responses to airborne acid aerosols.
43             Am. J. Public Health 81: 694-702.
44
45      Ozkaynak,  H.; Spengler, J. D.  (1985) Analysis of health effects resulting from population exposures to acid
46             precipitation precursors. Environ. Health Perspect. 63: 45-55.
47
48      Pope, C. A., III.; Dockery, D. W. (1992) Acute health effects of PM10 pollution on symptomatic and
49             asymptomatic children. Am.  Rev. Respir. Dis. 145:  1123-1128.
50
51      Pope, C. A., Ill;  Dockery, D. W.; Spengler,  J. D.; Raizenne, M. E.  (1991) Respiratory health and  PM10
52             pollution: a daily time series  analysis. Am. Rev. Respir. Dis.  144: 668-674.
53
         April  1995                                     2-34        DRAFT-DO NOT QUOTE OR CITE

-------
  1     Pope, C.  A., Ill; Schwartz, J.; Ransom, M. R. (1992) Daily mortality and PM10 pollution in Utah valley. Arch.
  2            Environ.  Health 47: 211-217.
  3
  4     Roemer, W.; Hoek, G.; Brunekreef, B. (1993) Effect of ambient winter air pollution on respiratory health of
  5            children with chronic  respiratory symptoms. Am. Rev. Respir. Dis. 147: 118-124.
  6
  7     Ross, D.  M.; Malm, W. C.; Loomis, R. J. (1985) The psychological valuation of good visual air quality by
  8            national park visitors. Presented at: the 78th annual meeting of the Air Pollution Control Association;
  9            June; Detroit, MI. Pittsburgh, PA: Air Pollution Control Association.
 10
 11     Ross, D.  M.; Malm, W. C.; Loomis, R. J. (1987) An examination of the relative importance of park attributes at
 12            several national parks. In: Bhardwaja, P. S.,  ed. Visibility protection: research and policy aspects.
 13            Pittsburgh, PA: Air Pollution Control Association.
 14
 15     Schwartz, J. (1989) Lung function and chronic exposure to air pollution: a cross-sectional analysis of NHANES
 16            II. Environ. Res. 50:  309-321.
 17
 18     Schwartz, J. (1991) Paniculate air pollution and daily mortality in Detroit. Environ. Res. 56: 204-213.
 19
 20     Schwartz, J. (1992) Air pollution and the duration of acute respiratory symptoms. Arch. Environ. Health
 21            47: 116-122.
 22
 23     Schwartz, J. (1993) Air pollution and daily mortality in Birmingham,  Alabama. Am. J. Epidemiol.
 24            137:  1136-1147.
 25
 26     Schwartz, J.; Dockery, D. W. (1992) Increased mortality in Philadelphia associated with daily air pollution
 27            concentrations. Am. Rev. Respir. Dis. 145: 600-604.
 28
 29     Schwartz, J.; Marcus, A. (1990) Mortality and air pollution in London:  a time series analysis. Am.  J. Epidemiol.
 30            131:  185-194.
 31
 32     Schwartz, J.; Wypij, D.; Dockery, D.; Ware, J.;  Zeger, S.; Spengler, J.; Ferris, B., Jr. (1991) Daily diaries of
 33            respiratory symptoms and air pollution: methodological issues and results. Environ. Health Perspect.
 34            90: 181-187.
 35
 36     Shumway, R. H.; Tai, R. Y.; Tai, L. P.; Pawitan, Y. (1983) Statistical analysis of daily London mortality and
 37            associated weather and pollution effects. Sacramento, CA: California Air Resources Board; contract no.
 38            Al-154-33.
 39
 40     Silver-man, F.; Hosein, H. R.; Corey, P.; Holton, S.; Tarlo, S. M. (1992) Effects of paniculate matter exposure
 41            and medication use on asthmatics. Arch. Environ.  Health 47:  51-56.
 42
 43     Skerry, B. S.; Alavi, A.; Lindgren, K. I. (1988) Environmental and electrochemical test methods for the
44            evaluation of protective organic coatings. J. Coat. Technol. 60: 97-106.
45
46     Stebbings, J.  H.,  Jr.; Hasselblad,  V.; Chapman, R. S.; McClain, K. E.  (1974) Ventilatory function in school
47            children Chattanooga 1971-1972. Research Triangle Park, NC: U.S. Environmental Protection Agency,
48            National Environmental Research Center,  Human Studies Laboratory; December 4.
49
50     Swift, D.  L.; Proctor, D. F. (1982) Human respiratory deposition of particles during oronasal breathing. Atmos.
51            Environ.  16: 2279-2282.
52
53
         April 1995                                     2-35       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Trijonis, J. C.; Pitchford, M.; McGown, M. (1987) Preliminary extinction budget results from the RESOLVE
 2             program. In: Bhardwaja, P. S., ed. Visibility protection: research and policy aspects, an APCA specialty
 3             conference; September  1986; Grand Teton National Park, WY. Pittsburgh, PA: Air Pollution Control
 4             Association; pp. 872-883. (APCA transactions no.  TR-10).
 5
 6      Trijonis, J.; McGown, M.;  Pitchford, M.; Blumenthal, D.; Roberts, P.; White, W.; Macias, E.; Weiss, R.;
 7             Waggoner, A.; Watson, J.; Chow, J.;  Flocchini, R. (1988) Visibility conditions and causes of visibility
 8             degradation in the Mojave Desert of California: executive summary,  RESOLVE project final  report.
 9             China Lake, CA: Department of the Navy, Naval Weapons Center; document no.  NWC TP 6869.
10             Available from: NTIS,  Springfield, VA; AD-A206 322.
11
12      Trijonis, J. C.; Malm, W. C.; Pitchford, M.; White, W. H. (1991) Visibility:  existing and historical
13             conditions—causes and effects. In: Irving, P. M., ed. Acidic deposition: state of science and  technology,
14             volume III, terrestrial, materials, health and  visibility effects. Washington, DC:  The U.S. National Acid
15             Precipitation Assessment Program.  (State of science and technology report no. 24).
16
17      U.S. Code. (1991) Clean Air Act, section 108, air quality criteria and control techniques,  § 109, national ambient
18             air quality standards. U.S. C. 42:  § 7408-7409.
19
20      U.S. Environmental Protection Agency. (1982) Air quality criteria for particulate matter and sulfur oxides.
21             Research Triangle Park, NC: Office of Health and Environmental Assessment, Environmental Criteria
22             and Assessment Office;  EPA report no. EPA-600/8-82-029aF-cF. 3v. Available from: NTIS, Springfield,
23             VA; PB84-156777.
24
25      U.S. Environmental Protection Agency. (1986) Second addendum to air quality criteria for particulate matter and
26             sulfur oxides (1982): assessment of newly available health effects information. Research Triangle  Park,
27             NC: Office of Health and Environmental Assessment, Environmental Criteria and Assessment Office;
28             EPA report no. EPA-600/8-86-020F. Available from: NTIS, Springfield, VA; PB87-176574.
29
30      U.S. Environmental Protection Agency. (1989) An acid aerosols issue paper: health effects and aerometrics.
31             Research Triangle Park, NC: Office of Health and Environmental Assessment, Environmental Criteria
32             and Assessment Office;  EPA report no. EPA-600/8-88-005F. Available from: NTIS,  Springfield,  VA;
33             PB91-125864.
34
35      Vedal, S.; Schenker, M. B.; Munoz, A.; Samet, J. M.; Batterman, S.; Speizer, F. E.  (1987) Daily air pollution
36             effects on children's respiratory symptoms and peak expiratory flow. Am. J. Public Health 77: 694-698.
37
38      Ware,  J. H.; Ferris, B. G., Jr.; Dockery,  D. W.; Spengler, J.  D.; Stram, D. O.; Speizer,  F.  E. (1986) Effects
39             of ambient sulfur oxides and suspended particles on respiratory health of preadolescent children. Am.
40             Rev.  Respir. Dis. 133: 834-842.
41
42      White, J.; Rothschild  (1987) Defects in the finish of motor cars. Met. Finish. 85:  15-18.
43
44      Xu, X.; Wang, L. (1993) Association of indoor and outdoor particulate level with chronic respiratory illness.
45             Am. Rev. Respir. Dis. 148: 1516-1522.
46
         April  1995                                     2-36       DRAFT-DO NOT QUOTE OR CITE

-------
  i                    3.  PHYSICS  AND CHEMISTRY OF
  2                           PARTICULATE MATTER
  3
  4
  5     3.1  INTRODUCTION
  6     3.1.1  Overview
  7          Atmospheric particles originate from a variety of sources and possess a range of
  8     morphological, chemical, physical, and thermodynamic properties.  Examples include
  9     combustion-generated particles such as  diesel soot or fly ash, photochemically produced
 10     particles such as those found in urban haze, salt particles formed from sea spray, and soil-
 11     like particles from resuspended dust. Some particles are liquid, some are solid; others
 12     contain a solid core surrounded by liquid.  Atmospheric particles contain inorganic ions and
 13     elements, elemental carbon, organics compounds, and crustal compounds.  Some atmospheric
 14     particles are hygroscopic and contain particle-bound water.  The organic fraction is especially
 15     complex.  Hundreds of organic compounds have been identified in atmospheric aerosols,
 16     including alkanes,  alkanoic and carboxcylic acids, poly cyclic aromatic hydrocarbons,  and
 17     nitrated organic compounds (Rogge et al.,  1993; Kaplan and Gordon, 1994; Mazurek et al.,
 18     1989; Standley and Simoneit,  1987; Ip  et al., 1984; Simoneit and  Mazurek, 1982; Schuetzle
 19     etal.,1975).
 20          Particle diameters span more than four orders of magnitude, from a few nanometers  to
 21     one hundred micrometers.  Combustion-generated particles, such as those from power
 22     generation, from automobiles, and in tobacco smoke, can be as small as 0.01 /^m and as
 23     large as  1  /xm. Particles produced in the atmosphere by photochemical processes range in
 24     diameter  from 0.05 to 2 pirn.   Fly ash produced by coal combustion ranges from 0.1 to
 25     50 ^m or more. Wind-blown dust, pollens, plant fragments, and cement dusts are generally
 26     above 2 /*m in diameter.  Particles as small as a few nanometers (Covert et al., 1992;
 27     Clarke, 1992) and  as large as  100 /*m have been measured in the atmosphere (Lin et  al.,
 28     1993).
29          Particles are ubiquitous in the atmosphere.  The lowest concentrations are found in
30     background marine environments, where particle number concentrations range from 100/cm3
31     to 400/cm3. In background continental  environments,  particle concentrations vary from
       April 1995                              3_!       DRAFT-DO NOT QUOTE OR CITE

-------
 1      100/cm3 to 5,000/cm3; while in urban areas of the United States concentrations may be as
 2      high as 4,000,000/cm3 (Willeke and Whitby, 1975; Whitby and Sverdrup, 1980). Particles
 3      account for a mass of a few /zg/m3  near the surface over dry continental areas to several
 4      hundred /ig/m3 in polluted urban areas.
 5          The composition and behavior of airborne particles are fundamentally linked with those
 6      of the  surrounding gas.  Although the term aerosol is often used to refer to suspended
 7      particles, aerosol is defined as a dilute suspension of solid or liquid particles in gas.
 8      Paniculate material can be primary  or secondary.
 9          Primary particles are those emitted in paniculate form and include wind-blown dust, sea
10      salt, road dust, mechanically generated particles  and combustion-generated particles such as
11      fly ash and soot.  The concentration of primary particles depends on their emission rate,
12      transport and dispersion, and removal rate from  the atmosphere.
13          Secondary particulate material may form from condensation of high temperature vapor
14      or from vapors generated at as a result of chemical reactions  involving gas-phase precursors.
15      Secondary formation processes can  result in either the formation of new particles
16      (Wiedensohler et al., 1994; Covert  et al., 1992;  Clarke  et al., 1991, 1993;  Frick and
17      Hoppel, 1993; Hoppel et al., 1994; Weber et al., 1994) or the addition of particulate
18      material to preexisting particles (Andreae et al.,  1986; Wall et al., 1988; Wu and Okada,
19      1994).  Most atmospheric sulfate is formed from atmospheric oxidation of sulfur dioxide.
20      Atmospheric nitrate is also essentially secondary, formed from reactions involving oxide of
21      nitrogen to form nitric acid.  A portion of the organic aerosol is also attributed to secondary
22      processes (Hildemann et al.,  1994;  Turpin and Huntzicker, 1991; Mylonas  et al., 1991;
23      Pickle et al., 1990; Gray et al., 1986).  Secondary aerosol formation can depend on
24      concentrations of other gaseous reactive species  such as ozone or hydrogen peroxide,
25      atmospheric conditions including solar radiation  and relative humidity,  and  the interactions of
26      precursors and preexisting particles with cloud or fog droplets (Meng and Seinfeld, 1993;
27      McMurry and Wilson, 1983; Hoppel and Frick,  1990).  As a result, it is considerably more
28      difficult to relate ambient concentrations of secondary species to sources of precursor
29      emissions than it is to identify the sources of primary particles.
30          Airborne particulate matter can be anthropogenic or biogenic in origin.  Both
31      anthropogenic and biogenic particulate material can occur from either primary or secondary

        April  1995                                 3-2       DRAFT-DO NOT  QUOTE OR CITE

-------
 1      processes.  Anthropogenic refers to participate matter which is directly emitted or formed
 2      from precursors which are emitted as a result of human activity.  Primary anthropogenic
 3      sources include fossil fuel combustion, fireplace emissions, and road dust.  Secondary
 4      anthropogenic paniculate material can be generated photochemically from anthropogenic
 5      SO2, NOX, or organic gases.  Primary biogenic sources include leaf waxes  and other plant
 6      fragments from plants (Simoneit and Mazurek,  1982).  In addition, plants emit gaseous
 7      species such as terpenes  (Lamb  et al., 1987).  Terpenes are photochemically reactive, and in
 8      the presence of nitrogen oxides  can form secondary organic particles (Kamens et al., 1981;
 9      Pondis et al., 1991, 1993).  Other types of primary particulate material such as sea salt and
10      wind-generated dust from soil undisturbed by man also are of non-anthropogenic origin.
11           In addition to secondary formation, volatilization and sorption processe's also  affect
12      concentrations  and compositions of airborne particles.  Some aerosol constituents are
13      semivolatile and exist in both gas and particle phases.  Their  gas-particle distribution depends
14      on  atmospheric conditions such as  temperature, the concentrations of other aerosol species
15      including water vapor, and the vapor pressure of the constituent.  Some inorganic compounds
16      such as ammonium nitrate (Stelson  and Seinfeld,  1982; Bassett and Seinfeld, 1983,  1984) and
17      organic compounds, including many polycyclic aromatic hydrocarbons (Yamasaki et al.,
18      1982; Ligocki  and Pankow, 1989; Pankow,  1987, 1994a,b) are semivolatile.  Diurnal
19      temperature fluctuations  can cause substantial changes in the particle-phase  concentrations of
20      semivolatile constituents as a result of gas-particle redistribution.  Evidence exists suggesting
21      that this volatilization-sorption cycle results  in the redistribution of semivolatile material
22      among particles of differing origins (Venkataraman and Hildemann, 1994).
23           Sampling semi-volatile species requires special techniques, such as the use of denuder
24      systems (Kautrukis et al., 1988). The processing of atmospheric particles also occurs in
25      clouds.  For example, in-cloud processes can lead to the combination of many small particles
26      (Andreae et al., 1986).
27           A complete description of the atmospheric aerosol would include an accounting of the
28      chemical composition, morphology,  and size of each particle  and the relative abundance of
29      each particle type  as a function  of particle size  (Friedlander,  1970).  However, most often
30      the physical and chemical characteristics of particles are measured  separately.  Number size
31      distributions are often determined by physical means,  such as electrical mobility or light-

        April 1995                                 3-3        DRAFT-DO NOT  QUOTE OR CITE

-------
 1     scattering.  Chemical composition is determined by analysis of collected samples.  The mass
 2     size distribution and the average chemical composition of the aerosol as a function of size
 3     can be determined by collection of size-segregated samples (Countess et al.,  1980; Hering
 4     and Friedlander, 1982; John et al., 1990; Sloane et al., 1991).  Recent developments in
 5     single particle analysis  techniques coupled with multivariate classification methods (Van
 6     Grieken and Xhoffer, 1992;  Germani and Buseck,  1991; Mansoori et al., 1994) are bringing
 7     the description envisioned by Friedlander closer to reality.  This introductory section
 8     describes some of the measurements that have been made on atmospheric particles, and the
 9     insights thus provided on the nature, origins, and atmospheric processes that affect particle
10     composition.
11
12     3.1.2  Major Chemical Constituents
13          The major constituents of atmospheric aerosol are sulfates, nitrates, carbonaceous
14     compounds, water,  ammonium ions and materials of crustal origin.  Inorganic ions,  including
15     sulfate and nitrate, are  typically analyzed by ion chromatography.  Crustal elements are
16     analyzed by x-ray fluorescence and/or proton-induced x-ray emission.  Average compositions
17     vary with particle size, by location and season.  The equilibrium models for  inorganic ions
18     predict that water is an important constituent of atmospheric particles, but measurements are
19     limited.  McMurry  and coworkers (McMurry and Stolzenburg,  1989; Zhang et al., 1993)
20     measured the sensitivity of particle size to relative humidity (RH) for Los Angeles and Grand
21     Canyon aerosols.   They found that atmospheric particles of a single size exhibited two
22     distinct hygroscopicities.   These were  described as  "more" and  "less" hygroscopic, as shown
23     in Figure 3-1.  For example, the diameters of more hygroscopic 0.2 /mi particles  humidified
24     to approximately  90%  RH increased by factors of 1.23 ± 0.08 and 1.49 ± 0.11 for Los
25     Angeles and Grand Canyon particles, respectively.  For relative humidities above  85 or 90%,
26     water was  the most abundant particulate species both in Los Angeles and at the Grand
27     Canyon.
28          Because of the multitude of carbonaceous compounds present in atmospheric aerosols,
29     carbonaceous material  is  often categorized as organic or elemental carbon (OC or EC).  Most
30     measurements of aerosol carbon are made using one of a variety of thermal techniques that
        April 1995                                3-4       DRAFT-DO NOT QUOTE OR CITE

-------


^
5
Q,
cf
^
2
O,
cf





<
s.
cf
*a
I


l.OU
1.55
1.50
1.45
1.40
1.35
1.30
1.25
1.20
1.15
1.10
1.05
1f\f\
.00
One
.yo
(
1Cf\
.60
1.55
1.45
1.40
1.35
1.30
1.25
1.20
1.15
1.10
1.05
1.00
A nc
i i i i i i i i i
o More Hygroscopic Particles
• Less Hygroscopic Particles
-
-
-
o o -
%
0°
o §S8
0 °
O ^ m
S& °& <$B^° i <*T i i 1*1

) 10 20 30 40 50 60 70 80 90 10
DMA2 Relative Humidity
1 I 1 1 1 1 1 1 Of
„ o More Hygroscopic Particles Oo -^ -
• Less Hygroscopic Particles 0 <§g3o
o
12N»
^K d> •
0 @% • ..
; Ose#& -
4, "WT -
° g ro0j*«1j'. -
» .^M«« %^° f*?1** ..:

1 1 f 1 ) |_ L 1 1
                             10   20   30    40   50   60   70
                                       DMA2 Relative Humidity
      Figure 3-1. Particle size related to RH.
               80   90  100
1     report paniculate organic and elemental carbon concentrations (Huntzicker et al., 1982;
2     Mueller, 1982; Turpin et al., 1990).  The split between organic and elemental carbon is
3     somewhat operationally defined, but the term elemental generally refers to the nonvolatile,
4     optically absorbing (black) portion of the carbon aerosol.  Elemental carbon is associated
5     with soot emissions from combustion.  The remaining, more volatile portion is termed
6     organic.  Various methods of further classifying the organic fraction include: selective
      April 1995
3-5
DRAFT-DO NOT QUOTE OR CITE

-------
 1     solvent extraction (Lioy and Daisey, 1986), functional group identification (Allen et al.,
 2     1994; Gordon et al., 1988), and division into neutral and acidic fractions (Hildemann et al.,
 3     1994a).  Radiocarbon dating techniques have been used to distinguish fossil and
 4     contemporary carbon (Currie et al., 1994; Kaplan and Gordon,  1994; Hildemann et al.,
 5     1994b).
 6
 7     3.1.3  Atmospheric Aerosol Size Distributions
 8          Size is one of the most important parameters in determining the atmospheric lifetime
 9     and deposition of particles.  As a result, the environmental, visual, and health effects of
10     atmospheric aerosols and the fate of the compounds that they contain are strongly dependent
11     on the particle size  distribution.  Particle size largely determines deposition patterns of
12     particle-phase compounds within the lung.  Light  scattering is also strongly dependent on
13     particle size, and thus particle size distributions have a strong influence on atmospheric
14     visibility and radiative  balance (i.e., climate).
15          Atmospheric size distributions for averaged continental background, urban-influenced
16     background,  averaged urban, and freeway-influenced urban aerosols are shown in
17     Figures 3-2 to 3-4 (Whitby and Sverdrup, 1980).  Figure 3-2 describes  the number of
18     particles as a function of particle diameter. For the same data,  the particle volume
19     distribution with respect to size is shown in Figures 3-3 and 3-4.  Number and volume
20     distributions are defined such that the number (or volume) of particles in a specified size
21     range is proportional to the corresponding area under the curve.  These distributions show
22     that most of the particles are quite small, below 0.1 /mi, while most of the particle volume
23     (and therefore most of the mass) is  found in particles > 0.1.
24          An important  feature of atmospheric aerosol size distributions is their multimodal
25     nature.  Volume distributions in ambient air are almost always bimodal, with a minimum
26     between 1 and 3 /on.  Particles in the larger mode are termed "coarse"  and those in the
27     smaller made, "fine".  Whitby and Sverdrup (1980) and Willeke and Whitby (1975)
28     identified three modes:  nuclei, accumulation, and coarse.  The three modes are most
29     apparent in the freeway-influenced size distribution of Figure 3-4.  The smallest mode,
30     corresponding to particles below about 0.08 /mi, is the nuclei mode.  The middle mode,
        April 1995                                3-6       DRAFT-DO NOT QUOTE OR CITE

-------
                   \
                    Q.
                    Q
                    J£

                    z
                    TJ
                       1,000,000--



                         10,000- -



                           100--
                          0.01 1
                         0.0001 - -
                       0.000001 - -
Average Background
Urban Influenced Background
Average Urban
Urban + Freeway
                                0.001
                                       0.01
                                              0.1
                                                    1
                                                          10
                                                                100
                                      Particle Diameter, Dp (urn)
Figure 3-2. Number of particles as a function of particle diameter.







f
TO
•^
"S
Q,
o>
o
TJ
1




70
65

60
55

50
45
40
35
30
25

20
15

10
5
0
0.(
I 	 \ 	 \ 	 7! 	 1 	 1 	
Background .' '.
	 Average •' '•
Background
	 Urban Influenced ,' '•
Background
	 South-Central • •
New .
1
-
-
/ A\
' : ! ',\
i . 1 t i
'• •' / Vi
1 '"''/ A
* i / '• \\
/ / \\\
:/'"^-/'J/""^^>^
X)1 0.01 0.1 1 10 1C
                                   Particle Diameter, Dp
Figure 3-3.  Particle volume distribution as a function of particle diameter.
April 1995
   3-7
DRAFT-DO NOT QUOTE OR CITE

-------
                              70
                              65
                              60--
                              55--
                           —, 50- •
                              40- •
                           § 30+
                           0) 25-
                           ^
                           "O
                              15-•
                              la-
                               s'-
Average Urban
Urban + Freeway
                              0.001     0.01     0.1      1      10
                                        Particle Diameter, Dp
                           100
       Figure 3-4. Particle volume distribution as a function of particle diameter in a
                   freeway-influenced area.
       Source: Whitby and Sverdrup (1980).
 1     from 0.08 to 1 or 2 /mi, is the accumulation mode. The largest particles (>  1 or 2 /mi)
 2     comprise  the coarse mode.  Formal delineation of these modes arises from the fitting of a
 3     trimodal,  lognormal distribution function to the data.
 4           Whitby and coworkers observed that continental background aerosols not influenced by
 5     sources have a small accumulation mode and no nuclei mode.  For urban aerosols, the
 6     accumulation and coarse particles modes are comparable in volume.  The nuclei mode is
 7     small in volume but dominates the number distributions of urban aerosols. More recent
 8     measurements of fine-particle size distributions (Eldering and Cass,  1994) and species-size
 9     distributions for sulfates, nitrates, and ammonium ion, as discussed below, indicate that the
10     accumulation mode can be further divided into a "condensation" and "droplet" mode (John
11     et al., 1990).  Measurements over remote areas (Hoppel et al.,  1986, 1990) indicate that the
12     nuclei mode can also be divided into two separate modes.
13           Many measurements indicate that the chemical compositions of coarse and fine particles
14     are distinct.  The processes that  affect the formation and removal of these two size fractions
       April 1995
           3-8
DRAFT-DO NOT QUOTE OR CITE

-------
  1      of atmospheric aerosols are also distinct.  Coarse particles are generated by mechanical
  2      processes and consist of soil dust, sea spray, plant fragments,  particles from tire wear, and
  3      emissions from rock-crushing operations.  These particles are  removed primarily by
  4      impaction and settling.  Nuclei and accumulation mode particles contain primary particles
  5      from combustion sources and secondary particles that result from condensation of low-
  6      volatility vapors formed from chemical reactions.  Particles  in the nuclei mode may be
  7      transferred into the accumulation mode by coagulation, but cloud coalescence and liquid
  8      phase cloud droplet transformations may be more important in atmosphere. In contrast,
  9      accumulation mode particles do not ordinarily grow into the coarse mode,  because number
 10      concentrations  are too low for coagulation to be effective.  Nuclei are readily removed by
 11      diffusion to surfaces.  However, accumulation mode particles  are not easily removed from
 12      the airstream.  They have long atmospheric lifetimes and are able to penetrate deep into  the
 13      lungs.  The nuclei and accumulation modes  are fairly independent of the coarse mode, both
 14      in formation and removal (Willeke and Whitby, 1975; Whitby and Sverdrup,  1980).
 15           Fine and coarse particles are best differentiated by their formation mechanism (Wilson,
 16      1995). Fine particles are formed by nucleation with gases while coarse particles are formed
 17      by mechanical processes from larger particles or bulk materials.  The most appropriate size
 18      cut for separating fine from coarse particles is still under consideration.
 19
 20      3.1.4   Chemical Composition  and Its Dependence  on Particle Size
 21           Since the work of Whitby, several studies have been conducted that provide chemical or
 22      elemental composition data on the coarse and fine fractions of the atmospheric aerosol.
 23      Generally this is done by separate collection of particles less than 2.5 ^m in diameter (fine or
 24      PM2 5) and particles less than 10  /^m  (PM10).  Coarse-particle  concentrations are obtained by
25      difference.  Alternatively, particles can be collected in two or  more size fractions, using
26      impactor methods.
27           Detailed size distributions of the inorganic ions in Los  Angeles are shown in Figure 3-5
28      (Wall et al.,  1988; John et al., 1990).  These data show two modes for sulfate and nitrate
29      aerosols between 0.1 and 1 /xm.  Similar results for sulfate aerosols were reported by Hering
30      and Freidlander (1982).  The smaller  mode, corresponding to particles near 0.2 pirn in
31      diameter, is attributed to gas-phase formation of condensible species and is referred to as the

        April 1995                                3.9       DRAFT-DO NOT QUOTE OR CITE

-------
                  500-
               o
               o>
               _g
               TJ
                  300-
               m
               £
               O
                  200 H
                  100-
                          Legend
                             Ammonium, NH
                             Nitrate, NO2"
                                       -2
Sulfate,SO4
Hydrogen Ion, H+
Sodium, Na+
Chloride, Cl"
                     0.01              0.1                1
                                       Aerodynamic Diameter, Dae (n,m)
       Figure 3-5.  In concentration as a function of particle size.
1      condensation mode.  The larger mode has a peak near 0.6 fim and is called the droplet
2      mode.  Its existence is attributed to secondary formation through heterogeneous, aqueous-
3      phase transformations.  McMurry and Wilson (1983) found 0.6 /^m sulfate particles in power
4      plant plumes and attributed their existence to formation by heterogeneous processes.  Further
5      analysis of the data by Meng and Seinfeld (1994) indicate that these aqueous reactions most
6      likely occur in cloud or fog droplets.
1           The data of Figure 3-5 in Los Angeles show that paniculate nitrate is found in both
2      coarse and fine particles.  Nitrate near the coast was predominantly in the coarse mode.
3      Coarse  mode nitrate was less prominent for inland sites.  Several investigators (Wall et al.,
4      1988; John et al., 1990; Andreae et al., 1986) proposed that the coarse particle nitrate results
5      from the heterogeneous reaction of nitric acid with sea salt.  On the basis of single particle
6      analysis by electron  microscopy-energy dispersive spectroscopy, Wu and Okada (1994)
7      concluded that coarse-particle nitrate in a coastal region of Japan formed on sea salt.  Course
8      nitrate collected at an inland site was associated with soil dust.  These data suggest that a
9      heterogeneous chemical reaction on the surface of a mechanically generated, primary particle
       April 1995                               3-10       DRAFT-DO NOT QUOTE OR CITE

-------
  1     may provide a mechanism for adding secondary material to the coarse particle mode.  They
  2     also show that secondary paniculate material can be formed by the interaction of a natural
  3     constituent (sea salt) with a species derived from anthropogenic emissions (nitric acid).
  4
  5     3.1.5    Particle-Vapor Partitioning
  6          Several atmospheric aerosol species, such as ammonium nitrate and certain organic
  7     compounds, are semi volatile and are found  in both gas and particle phases.  The gas-particle
  8     distribution of semivolatile organic compounds depends on compound vapor pressure,  total
  9     particulate surface area and composition, and atmospheric temperature (Pankow, 1987;
 10     Junge,  1977; Bidleman, 1988).  Junge (1977)  modeled this relationship using a linear  form of
 11     a Langmuir adsorption isotherm. Measurements of semivolatile organic compounds show
 12     that gas-particle distributions are highly correlated with total suspended particulate matter,
 13     temperature, and the sub-cooled liquid vapor pressure of the pure compound (Foreman and
 14     Bidleman, 1990; Ligocki and Pankow, 1989; Yamasaki et al., 1982).  Yamasaki et al. (1982)
 15     used this information to model an empirical relationship between the gas-particle distribution,
 16     total suspended particulate matter and temperature. Pankow showed that the expressions  of
 17     Junge (1977) and Yamasaki et al. (1982) are consistent and continued the theoretical
 18     development of equilibrium gas-particle partitioning (Pankow, 1987; 1991;  1994a,b).
 19          Although it is generally assumed that the gas-particle partitioning of semivolatile
 20     organics is  in equilibrium in the atmosphere, the kinetics of redistribution are not well
 21     understood.  Gerde and Scholander (1989) and Rounds and Pankow (1993) predicted that
 22     redistribution in the ambient air could take minutes to hours.  Since changes in atmospheric
 23     conditions (i.e., temperature) will drive redistribution, it is not clear whether equilibrium
 24     conditions are maintained. However, the gas and particle data agree reasonably well with
 25     equilibrium theories.  The development of an understanding of gas-particle partitioning of
 26     semivolatile organic compounds is hampered by the difficulty associated with measuring the
 27     multitude of compounds,  all present in small concentrations, for which diurnal temperature
28     fluctuations cause gas-particle partitioning to be dynamic on a time scale of a few hours.
29          Stelson and Seinfeld (1982) developed a thermodynamic model to predict the
30     temperature and relative humidity dependence of the ammonium nitrate equilibrium
31      dissociation constant, which has been supported by ambient data at inland sites in the Los

        April 1995                                3_H       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Angeles Basin (Hildemann et al., 1984; Doyle et al., 1979).  Bassett and Seinfeld extended
 2      the equilibrium model to include sulfates (1983) and the effect of particle size (1984).  With
 3      the inclusion of sodium chloride in the equilibrium model, Pilinis and Seinfeld (1987) were
 4      able to predict observations at coastal sites.  Atmospheric models based on equilibrium
 5      considerations have been successful in accounting for the gas-particle partitioning of
 6      inorganic species measured in Phoenix, Arizona (Watson et al., 1994b), and Uniontown,
 7      Pennsylvania (Saxena et al., 1994).  Wexler and Seinfeld (1992) found that under some
 8      atmospheric conditions the size distributions of ammonium ion and nitrate are not accurately
 9      predicted by equilibrium considerations alone,  and that transport kinetics can be important.
10
11      3.1.6   Single Particle Characteristics
12           The "mixing characteristics" of the aerosol describes the distribution of chemical
13      species among particles.  An aerosol in which  all particles contain the same homogeneous
14      blend of chemical species  is internally mixed.  In an externally mixed aerosol each chemical
15      species is found in a distinct set of particles. Experiments measuring atmospheric aerosol
16      properties for single-particle size ranges (Hering and McMurry, 1991; Covert et al.,  1990;
17      Zhang et al., 1993) and single-particle analyses (Bock et al., 1994; Sheridan et al., 1993;
18      Van Borm et al., 1989;  Anderson et al., 1988) indicate that atmospheric aerosols are to some
19      degree both internally and externally mixed. Single particle analyses provide descriptions of
20      individual particle compositions. These are then categorized either manually or through
21      multivariate methods such as cluster analysis (Kim and Hopke, 1988) to give an accounting
22      of the relative number of particles of each chemically defined particle type. Morphological
23      information can also be  included in particle type definitions.
24           Single-particle composition and morphology provide insights into the sources and
25      atmospheric processes affecting airborne particles.  For example, a priori one expects that
26      particles emitted from different sources would  in fact be distinct.  However, Andreae et al.
27      (1986) observed that over remote ocean areas between 80 and 90% of silicon-rich particles
28      (presumably originating from silicate mineral particles) were also rich in sodium, chlorine,
29      and variable amounts of potassium, magnesium, calcium, and sulfur (attributed to sea salt
30      particles).   The internal mixing of silicates with sea salt, particles originating from different
31      sources and externally mixed when emitted into the atmosphere, suggests the processing of

        April 1995                                 3-12       DRAFT-DO NOT QUOTE OR CITE

-------
  1     aerosol particles within clouds (see Section 3.2.1.4).  The hypothesis is that a single cloud
  2     droplet takes up two or more particles and that these particles remain together after droplet
  3     evaporation.  Other mechanisms of particle coalescence, such as differential settling,
  4     Brownian coagulation, and electrostatic attraction, are too slow to account for the large
  5     fraction of internal mixing observed.  Andreae et al. (1986) also found enrichment of sulfur
  6     (presumably sulfate) on sea salt particles.  This also was attributed to the interaction of
  7     clouds with particles.  Gas-to-particle conversion in cloud droplets or by condensation can
  8     also lead to mixtures of aerosol species.
  9          Particle morphology has many effects on atmospheric particle properties and processes.
 10     Chain agglomerates, for example,  have much  larger surface areas  on which adsorption and
 11     chemical reactions can take place than spherical particles of identical volumes.  In addition,
 12     the atmospheric lifetime is longer, and the optical absorption per unit mass is greater for
 13     chain agglomerates than for more compact particles. Combustion-generated soot particles are
 14     often chain agglomerates composed of a large  number of small primary spherules.
 15     Laboratory experiments conducted by Huang et al.  (1994) and Colbeck et al. (1990)
 16     demonstrated that condensation-evaporation processes can cause chain agglomerates to
 17     become more compact.  Colbeck et al. (1990) also  showed that the collapse of the soot
 18     aggregates resulting from  humidification results in a decrease in both the optical scattering
 19     and extinction of the particles.
 20
 21      3.1.7   Definitions
 22      3.1.7.1    Definitions of Particle Diameter
 23           The  diameter of a particle may be determined geometrically,  from optical or electron
 24      microscopy; by light scattering and Mie theory, or by its behavior, such as its electrical
 25      mobility, its settling velocity, or its aerodynamic behavior.  Although atmospheric particles
 26      are often not spherical, their diameters are described by an "equivalent" diameter, that of a
 27      sphere which would have the same physical behavior.  Two parameters that are often used to
 28      describe particle diameter  are the Stokes and aerodynamic diameters.  The Stokes diameter,
29      Dp, describes particle  size based on the aerodynamic drag force imparted on a particle when
30      its velocity differs from that of the surrounding fluid.  For a smooth,  spherically shaped
31      particle, Dp exactly equals the physical diameter of the particle.  For irregularly shaped

             1995                               3_13       DRAFT-DO  NOT QUOTE  OR CITE

-------
 1      particles, Dp is the diameter of an equivalent sphere that would have the same aerodynamic
 2      resistance.  Particles of equal Stokes diameters that carry the same electric charge will have
 3      the same electrical mobility.  Particles of equal density and equal Stokes diameter have the
 4      same settling velocity.
 5           Aerodynamic diameter, Da, depends on particle density and is defined as the diameter
 6      of a particle with equal settling velocity but a material density of 1 g/cm3. Particles with the
 7      same physical size and shape but different densities will have the same Stokes diameter but
 8      different aerodynamic diameters.  For particles greater than about 0.5 pm, the aerodynamic
 9      diameter is generally the quantity of interest because it is the parameter that is important to
10      particle transport,  collection, and respiratory tract deposition.  Respirable, thoracic, and
11      inhalable particle sampling are based on particle aerodynamic diameter.
12
13           Aerodynamic diameter, Da, is related to the Stokes diameter, Dp, by:
14
                                                 P C
                                                       1/2
(3-D
15
16      where p is the particle density, and C and Ca are the Cunningham slip factors evaluated for
17      the particle diameters Dp and Da respectively.  The slip factor is a function of the ratio
18      between particle diameter and mean free path of the suspending gas; it is given by the
19      expression (Hinds,  1982):
20

                             C  =  1  + — {2.514 + 0.800 exp(-0.55 _?)}                 (3-2)
                                      DP                            X
21
22      where X is the mean free path of the air.  C is an empirical factor that accounts for the
23      reduction in the drag force on particles due to the "slip" of the gas molecules at the particle
24      surface. It is important for particles less than 1  /nm in diameter, for which the surrounding
25      air cannot be modeled by a  continuous fluid.  At normal atmospheric conditions (temperature
26      = 20  °C, pressure =  1 atmosphere) X = 0.066 /xm.  For large particles (Dp > 5 pim)
27      C =  1; while for smaller particles C  >  1.
        April  1995                                3-14      DRAFT-DO NOT QUOTE OR CITE

-------
 1          For particles with diameters greater than the mean free path, the aerodynamic diameter
 2     given by equation (3-1) is approximated by:
                               Da=(p)1/2Dp                  (Dp>X)                    (3-3)
 3
 4     This expression, which shows that aerodynamic diameter is directly proportional to the
 5     square root of the particle density, is often used for particles as small as 0.5 /mi.  For
 6     particles with diameters much smaller than the mean free path, the slip factor must be taken
 7     into account.  In this case the aerodynamic diameter is directly proportional to the particle
 8     density [Da = (p) Dp for Dp«X].
 9
10     3.1.7.2   Definitions of Particle Size Fractions
11          In the preceding discussion several modes of the aerosol size distribution were
12     identified;  they are defined as follows:
13
14          Nuclei Mode:  that portion of the fine mode particles with diameters below about
15          0.08  /mi;
16
17          Accumulation Mode: particles formed from gases;
18
19          Condensation Mode:  that portion  of the accumulation mode with a volume (mass)
20          median diameter near 0.2 /mi;
21
22          Droplet Mode:  that portion of the accumulation mode with a volume (mass) median
23          diameter at 0.5 to 0.8 jim;
24
25          Fine  Particles: The combination of the modes listed above.
26
27          Coarse Mode or Coarse Particles:  mechanically generated particles.
28
29
30          There is some overlap between fine and coarse particles in the  1 to 3 jim region.  For
31     further discussion see Chapter 4. PM2 5 refers to particles less than  2.5 /xm diameter and is
32     frequently referred to as "fine" PM. A discussion of the best size to differentiate fine from
33     coarse particles  in given in Section 3.7.
34          Another set of definitions of particle size fractions arises from considerations of size-
35     selective sampling.  Size-selective sampling refers  to the collection of particles below a

       April 1995                               3-15       DRAFT-DO NOT QUOTE OR CITE

-------
1
2
3
4
5
6
7
8
9
specified aerodynamic size cut, and has arisen in an effort to measure the particle size
fractions of importance to human health.  The PM10 standard set by the U.S. Environmental
Protection Agency in 1987 is one example of size-selective sampling criteria, and it was
designed to match the penetration of particles into the thoracic region.  It is a fairly sharp
cutpoint with a 50% efficiency near 10 ^m in particle aerodynamic diameter.  The exact
definition is given by a table of efficiency  values (Federal Register, 1988) and is shown in
Figure 3-6.
                          100r
                                                                 A PM-10
                                                                 • IPM
                                                                 • TPM
                                                                 O RPM
                                                   10
                                       Aerodynamic Particle Diameter (urn)
       Figure 3-6. Efficiency values for size-selective sampling criteria.
                                                                  100
1
2
3
4
5
6
7
8
3.1.7.3  Other Terminology
     Other terminology that has been introduced in this section is summarized below:
     Primary Particles: those directly emitted to the atmosphere from either natural sources
           or sources derived from human activity;
     Secondary Particulate Material:  material formed in the atmosphere as the result of
           chemical and physical conversion of gaseous species;
       April 1995
                                          3-16
DRAFT-DO NOT QUOTE OR CITE

-------
 1           Internal Mixture:  an aerosol for which the chemical composition of each individual
 2                 particle is the same, that is, equal to the bulk composition;
 3
 4           External Mixture:   an aerosol for which different chemical species comprise separate
 5                 particles;
 6
 7           Anthropogenic:  derived from human activities;
 8
 9           Biogenic:  derived from plants;
10
11           Bioaerosols: airborne microorganisms and aeroallergens;
12
13           Fossil: derived from fossil fuel combustion; and
14
15           Contemporary carbon:  derived from non-fossil fuel sources such as plants, wood
16                 burning, and cooking oils.
17
18      3.1.8   Field Studies
19           Appropriate and reliable field measurements play a central role in shaping our under-
20      standing of atmospheric processes, in providing key model inputs, and in the evaluation of
21      models.  Real-world observations are all the more important in the case of atmospheric
22      aerosols, which, on the one hand, are the end product of many complex processes and, on
23      the other hand,  are key precursors of important microphysical cloud processes.  Field studies
24      include short-term, 3-D, high-resolution intensive research campaigns, as well as longer-term
25      surface and upper-air monitoring programs (in routine mode,  or in more comprehensive and
26      higher-resolution research mode). Research studies are generally mechanistic (targeted at
27      understanding of process rates and mechanisms), and/or diagnostic (aimed at development
28      and testing of individual process modules or subgrid-scale parameterizations  for use in
29      complex models). Routine monitoring studies are aimed more at operational evaluation of
30      overall model performance, or at generation of model input data including those (e.g.,
31      meteorological) which,  through dynamic assimilation into the computations,  can improve the
32      realism of the simulations. Since atmospheric fine particles (FP) are substantially of
33      secondary origin, measurements of their gaseous precursors and other reactants are also
34      important.   In North America, most of the anthropogenic emissions of FP and their
35      precursors are from large point sources  (power plants and smelters) and from urban-
        April 1995                                3-17      DRAFT-DO NOT QUOTE OR CITE

-------
 1      industrial complexes including vehicle emissions.  Consequently, special attention is given in
 2      this section to measurements in the plumes of such emissions.
 3
 4      3.1.9   Dry Deposition
 5           Dry deposition is the process whereby airborne gases and particles are transported
 6      down to the surface of the earth where they are removed.  Atmospheric turbulent mixing
 7      continually brings airborne  gases and particles into close proximity to the earth's surface,
 8      where they may diffuse across a thin layer of stagnant air to the surface itself.  Actual
 9      removal at the surface depends on the affinity between the airborne substance and the surface
10      element (ground, body of water, vegetation surface, etc.).  Dry deposition is a complex
11      process but it is represented as occurring in three steps: 1) transport down to the vicinity of
12      the earth by turbulent mixing processes; 2) diffusion across a thin quasi-laminar layer of air;
13      and 3) attachment to the surface itself.
14
15      3.1.10  Atmospheric Scavenging
16           Atmospheric gases are scavenged directly by absorption in droplets and by chemical
17      reactions in clouds.  The direct absorption of gases in falling droplets depends on the
18      solubility of the gas in water, and may be affected by the presence of other species in
19      solution (Seinfeld,  1986).  Particles are scavenged when they serve  as cloud condensation
20      nuclei (CCN) and when they are intercepted by  falling hydrometers. The wet removal  of
21      particles depends on the air trajectories through  clouds, the supersaturation to which the air
22      mass  is exposed, and the tune for which droplets are present before arriving at  the ground.
23
24
25      3.2   PHYSICAL PROPERTIES
26      3.2.1   Aerosol Size Distributions
27      3.2.1.1   Particle Size Distribution Functions
28           The distribution of particles  with respect to size is perhaps the most important physical
29      parameter governing their behavior.  The concentration of the number of particles as a
30      function of their diameter is given by a particle  number distribution.

        April 1995                               3_18       DRAFT-DO NOT QUOTE OR CITE

-------
  1           Because the sizes of atmospheric particles cover several orders of magnitude in particle
  2      size, and because atmospheric size distributions tend to be log-normally distributed (see
  3      Section 3.2.2), size distributions are often expressed in terms of the logarithm of the particle
  4      diameter, on the X-axis and differential concentration on the Y-axis:
                                                   dn
                                                d(logDp)
 6
 7      =  the number of particles per cm3 of air having diameters in the size range from log Dp to
 8      log(Dp + dDp).
 9
10      Formally, it is not proper to take the logarithm of a dimensional quantity, but one can think
11      of the distribution as a function of log(Dp/Dp0) where the reference diameter Dp0 = 1  jum is
12      not explicitly stated.
13           The number of particles is proportional to the area under the curve of n(logDp) versus
14      logDp. Similar considerations apply to distributions of surface, volume, and mass.
15
16      3.2.1.2   Log-Normal Size Distributions
17           As presented in Section 3.1, atmospheric aerosols tend to follow a sum of log-normal
18      distributions. A log-normal distribution is a specific functional form of the  size distribution
19      function for which the population of particles  follows a Gaussian distribution function with
20      respect to the logarithm of the particle diameter. The geometric standard deviation ag is the
21      standard deviation of the  quantity logDp  and defines the width of the distribution.  For a
22      monodisperse aerosol, that is, one for which all particles are the same diameter, oe =  1.
                                                                                    o
23      For polydisperse aerosols, ag >  1.  Typical values for one of the modes of the atmospheric
24      aerosol, such as the accumulation mode discussed above,  are 1.8  < ag <2.8.   For log-
25      normal distributions, 84.1% of the particles are below the size ffg-Dgn,  15.9% lie above the
26      size Dgn/ag, and 95% of the particles lie within two standard deviations of the mean, that is,
27      the range from Dgn/2ag to Dp-2ag.
        April 1995                                349       DRAFT-DO NOT QUOTE OR CITE

-------
 1           One of the properties of the log-normal distribution is that if the number distribution is
 2      log-normal, the surface and volume distributions are also log-normal, and their geometric
 3      standard deviation ag is the same as for the number distribution.
                          o
 4
 5      3.2.1.3   Ambient Aerosol Size Distributions
 6           Ambient aerosol size distributions are completely described by the geometric number
 7      mean diameter Dgn,  geometric standard deviation crg,  and number concentration N for each
 8      mode, as listed in Table 3-1. Also given are the parameters of the lognormal volume
 9      distributions, geometric mean diameters Dgv, and the corresponding total particle volume for
10      each mode V.  Because the distributions are lognormal, the geometric standard deviation is
11      the same for both number and volume distributions. The results  from more recent
12      measurements in  a nonurban area of New Mexico,  for which the distribution is described by
13      the sum  of two lognormal distributions, were obtained with laser light scattering
14      instrumentation, and were fitted to a bimodal lognormal form,  corresponding to the
15      accumulation and coarse particle modes.  The lognormal fit parameters characterizing these
16      distributions are also listed in Table 3-1.  Note that the volume geometric mean diameters for
17      the accumulation mode  vary from 0.2 /xm to 0.4 mm,  those for the coarse mode from 5 to
18      12 urn.  The standard deviations for the coarse particle mode tend to be larger than for the
19      accumulation mode.
20
21      3.2.1.4   Coagulation of Spherical Particles
22           Many processes affect the size distribution of an aerosol,  including addition of volume
                                                                                      s
23      by gas-to-particle conversion, and losses by deposition.  Even without these processes, under
24      conditions in which the total volume of an aerosol is conserved the number of particles will
25      decrease by coagulation while the average volume per particle  increases.  The coalescence of
26      two  particles always reduces the total surface area and therefore is  favored
27      thermodynamically.  Thus, in this sense, aerosols are  inherently unstable. In some cases
28      coagulation leads to  the formation of chain agglomerates, such  as for soot and some metals.
        April 1995                                3_20      DRAFT-DO NOT QUOTE OR CITE

-------
                    TABLE 3-1. LOGNORMAL PARAMETERS FOR AMBIENT AEROSOLS
K A. Parameters of
i§






Site of Measurement
Clean continental background
Average continental background
Urban influenced background
Urban average
Urban and freeway
South central NM - February 1989
South central NM - July 1989


Nuclei
Num.
(cm3)
N: 1,000
N
N
N
N
N
N
: 6,400
: 6,600
: 106,000
: 2,120,000


Model
Dgn
0.016 1.6
0.015 1.7
0.014 1.6
0.014 1.8
0.013 1.74


the Number
Accumulation Mode
Num.
(cm3)
800.00
2,300.00
9,600.00
32,000.00
37,000.00
706.00
253.00
Dgn
(/mi)
0.067
0.076
0.120
0.054
0.032
0.13
0.13
B. Parameters of
u>
to


DRAFT-DO 1
3
O
a
o
Site of Measurement
Clean continental background
Average continental background
Urban influenced background
Urban average
Urban and freeway
Sources: (1) Whitby and Sverdrup




V
V
V
V
V
(1980);


Nuclei
Volume
(/*m3 cm3)
: 0.01
: 0.04
: 0.03
: 0.63
: 9.20
(2) Kim et al.


Model
Dgv
0.030 1.6
0.034 1.7
0.028 1.6
0.038 1.8
0.032 1.74
(1993).


a
2.1
2.0
1.84

1.98
1.72
1.71
the Volume
Accumulation Mode
Volume
(/Ltm3 cm3)
1.50
4.45
44.00
38.40
37.50



Dgv
0.35
0.32
0.36
0.32
0.25



a
2.1
2.0
1.84
2.16
1.98



Distribution



Coarse Mode
Num.
(cm3)
0.72
3.2
7.2
5.4
4.9
0.42
0.72
Distribution
Dgn
(/im)
0.93
1.02
0.83
0.86
1.08
2.45
1.59

Og
2.2
2.16
2.12
2.25
2.13
1.91
2.27

Reference
(1)
(1)
(1)
(1)
(1)
(2)
(2)

Coarse Mode
Volume
(/mi3 cm3)
5.0
25.9
27.4
30.8
42.7



Dgv
6.0
6.04
4.51
5.7
6.0



ffg
2.0
2.16
2.12
2.25
2.13



Reference
(1)
(1)
(1)
(1)
(1)




n

-------
 1      3.2.2   Particle Formation and Growth
 2           A significant portion of the fine atmospheric aerosol is secondary, it is material added
 3      to the particle phase as the result of gas-to-particle conversion processes.  For example, fine-
 4      particle sulfate and nitrate particles are mostly formed by secondary processes.  One
 5      mechanism of gas-to-particle conversion is homogeneous gas-phase chemical reactions to
 6      form a condensible species, such as the oxidation of sulfur dioxide to form sulfuric acid.
 7      Condensible species can either nucleate to form a new particle (nucleation), or can condense
 8      onto the surface of an existing particle  (condensation). Another important class of gas-to-
 9      particle conversion mechanisms is  heterogeneous chemical reactions, which are chemical
10      reactions involving both gas-phase and particle-phase constituents.  Transformation on the
11      surface of particles, such as the uptake of nitric acid on the surface of sodium chloride (sea
12      salt) particles to produce nitrate is  one  type of heterogeneous reaction. Chemical reactions
13      within aerosol and cloud droplets,  such as when sulfur dioxide dissolves within an aqueous
14      droplet and is subsequently oxidized to sulfate, are another important heterogeneous gas-to-
15      particle mechanism.  Heterogeneous reactions lead to addition of aerosol material to existing
16      particles.  Nucleation results in an increase in particle number as well as an increase  in
17      particle mass. In this section we consider the physical aspects of these gas-to-particle
18      conversion mechanisms, and their  effects on the particle  size  distribution.
19
20      3.2.2.1   Equilibrium Vapor Pressures
21           An important parameter in particle nucleation and in particle growth by condensation is
22      the  saturation ratio S, defined as the ratio of the partial pressure of a species, p,  to its%
23      equilibrium vapor pressure above a flat surface, p0:  S = p/p0.  For either condensation or
24      nucleation to occur, the species vapor pressure must exceed its equilibrium vapor pressure.
25      For particles, the equilibrium vapor pressure  is not the same  as p0. Two effects are
26      important:  (1) the Kelvin effect, which is an increase in the equilibrium vapor pressure
27      above the  surface due to its curvature;  thus very small particles have  higher vapor pressures
28      and will not be stable to evaporation until they attain a critical size and (2) the solute effect,
29      which is a decrease in the equilibrium vapor pressure due to the presence of other
30      compounds.
        April 1995                                3-22       DRAFT-DO NOT QUOTE OR CITE

-------
  1           For an aqueous solution of a nonvolatile salt, the presence of the salt decreases the
  2      equilibrium vapor pressure of the drop. This effect is in the opposite direction as the Kelvin
  3      effect,  which increases the equilibrium vapor pressure above a droplet because of its
  4      curvature.
  5
  6      3.2.2.2  New Particle Formation
  7           When the vapor concentration of a species exceeds its  equilibrium concentration
  8      (expressed as its equilibrium vapor pressure), it is considered condensible. Condensible
  9      species can either condense on the surface of existing particles or can form new particles.
10      The relative importance of nucleation versus condensation depends on the rate of formation
11      of the condensible  species and on the surface area of existing particles.  An analytical
12      relation for the relative importance of each pathway is dependent on the ratio of the square of
13      the available surface area to the rate of formation (McMurry and Friedlander, 1979).  In
14      urban environments, it was found that new particle formation is found only near sources of
15      nuclei such as freeways because  the available surface  area is sufficient to rapidly scavenge
16      the newly formed condensible species. Wilson et al.  (1977) report observations of nuclei
17      mode in traffic.  New particle formation can also be observed in cleaner, remote regions.
18      Bursts of new particle formation in the atmosphere under clean conditions correspond to low
19      aerosol surface area concentrations (Covert et al., 1992).  The highest concentration of
20      volatile ultrafine particles occur in regions corresponding to the lowest particle mass
21      concentrations, indicating that new particle formation  is inversely related to the available
22      aerosol surface area Clarke (1992). In contrast to continental aerosols where sulfate
23      formation is the result of conversion of sulfur dioxide, the sulfur particles over the oceans
24      are formed from the oxidation of dimethylsulfide emitted by phytoplankton (Charlson et al.,
25      1987).
26
27      3.2.2.3  Particle  Growth
28           When material is added to the particle phase by  condensation or by heterogeneous
29      chemical reactions, particles of different sizes may grow at different rates, depending  on the
30      mechanism involved.  Condensational growth can have a different effect on the size
31      distribution of the aerosol than the effect of heterogeneous conversion through chemical

        April 1995                                3_23       DRAFT-DO NOT QUOTE OR CITE

-------
 1      reactions within a droplet.  The relative rates at which the size of particles change depends
 2      on whether the rate-limiting step in the growth process is transport to the particle, chemical
 3      reactions at the surface of the particle, or chemical reactions within the particle.  These are
 4      referred to as transport-limited, surface-reaction rate-limited or volume-reaction rate-limited.
 5      These different physical mechanisms give rise to a different form of the growth law for the
 6      particle.  Growth laws are the expressions for dv/dt or dD /dt as a function of particle size
 7      (where v is single particle volume and Dp is particle diameter).
 8           For condensational growth, the rate-limiting step relevant to the rate at which particles
 9      of different size grow is transport of condensible species to the particle surface.  For
10      particles much smaller than the mean free path of air, transport is governed by single
11      molecular bombardment of the surface, and the volume (or mass) of these particles grows in
12      proportion to their surface area.  For particles  larger than the mean free path, transport is
13      governed by diffusion.  In this regime the loss  of diffusing species at the  surface of the
14      particle causes a  gradient in the concentration of the diffusing species near the surface of the
15      particle such that the volume of the particle grows in proportion to particle diameter rather
16      than surface area.
17
18      3.2.2.4  Resuspension
19           The resuspension of deposited material as well as the suspension of material which has
20      not been previously  airborne can be an important source of particulate contamination.  This
21      discussion will use "resuspension" to include both resuspension and suspension.  Surface
22      contamination may result from the atmospheric deposition of a number of materials; for some
23      of these (e.g., plutonium), resuspension has been considered to  be the most important
24      exposure pathway.  Likewise, resuspended soil particles  have the greatest atmospheric mass
25      over continents of any single particle type (Peterson and Junge, 1971).  Despite this
26      importance, the literature shows relatively few experimental or theoretical studies for the
27      resuspension mechanism compared to other aerosol generation mechanisms.  The following
28      summarizes work on the physics of resuspension, physical/chemical properties of
29      resuspension generated  particles, and levels of  production and transport of resuspended
30      particles.
31

        April  1995                               3_24      DRAFT-DO NOT QUOTE OR CITE

-------
  1      Resuspension Mechanics
  2           Resuspension studies may be divided into applied research and detailed studies of
  3      mechanisms.  Applied studies are usually motivated by atmospheric deposition of hazardous
  4      substance (i.e., radionuclides from the Chernobyl 1982 accident [Cambray, 1989]) and the
  5      need to predict the spreading of contamination and the lifetime of hazardous air
  6      concentrations.  Resuspension experiments have been conducted over a wide range of surface
  7      types.  Many experiments have been conducted in dry or arid regions, simply because many
  8      contamination events have occurred  in such locations (i.e., the Nevada Test Site).  Of the
  9      experiments conducted over vegetation, most have been related to short grass.  Alternately,
10      applied studies may be motivated by mitigation efforts for soil erosion by wind or by need
11      for measurement of high atmospheric paniculate  concentrations caused by  resuspension,  so-
12      called "fugitive dust". Experiments concerning wind erosion have largely  occurred in
13      locations where wind erosion is prevalent, i.e., in the "Dust Bowl" area of the central United
14      States).
15
16           Applied Studies
17           Resuspension can occur due to the action of wind or by mechanical stresses.  Applied
18      research considers resuspension factors, K (air concentration divided by  surface
19      concentration) (units of length"1) and resuspension rates (flux of contaminant divided by
20      surface concentration) (unit of time"1).  Mechanical stresses, such as disturbances by traffic
21      or agricultural operations,  might result  in large amounts of resuspension over short intervals
22      in specific localities. For  example,  Sehmel (1984) quotes K values of 4 X 10 m"1 (for
23      beryllium particles by vigorous  sweeping in an unventilated room) to 7 x  10"3 m"1 for
24      plutonium particles in extensive traffic at the Nevada Test Site to 3 x 10"7 m"1 for gamma-
25      radioactive-fallout by walking on the deposit in an Australian desert.
26           Wind  generated resuspension is considered to be of major importance because it can be
27      relatively continuous and can occur  over large regions. Resuspension has been found to
28      increase as  a power of wind speed (with the resuspension rate being related to the second or
29      third power of wind speed). Examples of resuspension factors from wind  stresses quoted by
30      (Sehmel, 1984) range from 3 x  10"4 m"1 for uranium at Maralinga trials to 9 x 10"11  m"1 for
31      yttrium chloride on a cleared, sandy soil. Part of the range of K's quoted  above might be

        April 1995                                3_25       DRAFT-DO NOT  QUOTE OR CITE

-------
 1      caused by the aging of deposits, although a lack of understanding of the mechanisms
 2      dominant in the resuspension process has precluded identifying any reasons for the wide
 3      range  of results.
 4          Nicholson's (1993) data verify previous work, giving an approximate I/time decrease of
 5      the resuspension rate.  Makhon'ko's (1986) data for resuspension from grass suggest a
 6      relationship between relative resuspension rate K' versus phytomass m in grams per square
 7      meter,
 8
 9                                   K'  = 2.9 x 1(T8 m-1-4 [sec'1].                         (3-4)
10
11          Aerodynamic Resuspension
12          Aerodynamic Models include (1) balance of forces models and (2) statistical
13      mechanisms.  Balance of forces models account for forces holding the particles to the
14      surfaces versus  those forces acting to remove the particles from the surfaces.  Experimental
15      studies of particle motions show that particles being entrained into a turbulent fluid tend to
16      move  vertically into the stream with unsteady motions (Sutherland, 1967). Braaten et al.
17      (1990) and Braaten and Paw U (1992) stressed the  importance of bursts of a sweeping eddy
18      having the characteristics of large shear stress near the wall where particles are sparsely
19      deposited, breaking up the viscous sublayer and transporting fluid forces to the particles.
20      This mechanism removes particles  from a surface in short bursts followed by periods of little
21      resuspension activity.  Observations of Lycopodium spores placed on the  flat floor of a wind
22      tunnel were used to verify the model.
23          Reeks et al.  (1988) proposed a different aerodynamic mechanism that would account for
24      sudden random  injections of particles into the air, the injections taking place more randomly
25      in time than in the above force balance model.   Their mechanism calls for the individual
26      particles to  accumulate energy from the turbulent stream (most efficiently at  a resonant
27      frequency for the particle).  Accumulation of energy takes place because energy dissipation is
28      limited by the local fluid and substrate.  Once sufficient energy  has accumulated  to overcome
29      the potential energy well holding it in place, the particle is resuspended.  Slow motion
30      movies of saltating sand surfaces showed such a vibrating motion of a particle before it
31      becomes  airborne (Willetts, 1992).

        April  1995                                3_26      DRAFT-DO NOT QUOTE OR CITE

-------
  1           Mechanical Resuspension
  2           The importance of mechanical disturbance is seen in the differences of resuspension
  3      factors given by Sehmel (1984) for mechanical activities over contaminated soil versus those
  4      for wind. Another example of the comparison of resuspension by mechanical disturbance
  5      with resuspension by the wind was given by Garland (1979) as a two-order-of-magnitude
  6      increase of the resuspension factor for the mechanical disturbance of a full 5-liter bottle
  7      dragged along the grass 20 times in 5 minutes in wind compared to the 10 m/s wind alone.
  8      Sehmel (1984) conducted experiments to determine the fraction of tracer particles
  9      resuspended by driving cars and trucks through the deposited tracer or near the deposited
 10      tracer. The fraction increases with speed and size of vehicle.  The fraction resuspended per
 11      vehicle pass increased as the first power of vehicle speed for the truck driven through the
 12      tracer, the fourth power of vehicle speed for a car driven through the tracer, and the third
 13      power of the vehicle speed for a car driven near the tracer.
 14           The emission of PM-10 particles in wind erosion is driven by the mechanical process of
 15      sandblasting, although Shinn et al.  (1983) have pointed out the importance of direct
 16      aerodynamic emission for low emission  rates below erosion threshold.  Threshold velocities
 17      for particles smaller than 10 micrometer diameter are several times greater than that for  100
 18      micrometer particles (Bagnold,  1941).  Nonetheless, one observes submicrometer to 10-
 19      micrometer particles in wind erosion events for winds very much below the threshold
 20      velocity for  the above mentioned particles.  Gillette and Walker (1977) interpreted  this to be
 21      caused by the mechanical suspension (sandblasting) of fine particles by  more-easily-eroded
 22      sand particles.  Shao et al.  (1993) showed that sand-grain bombardment (saltation)  is the
 23      overwhelmingly dominant mechanism in maintaining  fine particle emissions from the surface.
 24      To derive an expression for the emission of dust, Shao et al. (1993) assumed that the number
25      of dust particles dislodged from a surface per sand grain impact was proportional to the ratio
26      for the kinetic energy loss of the impacting sand grain to the binding potential energy holding
27      a dust particle to the surface. This assumption led to the prediction that the dust flux is
28      proportional to the sand grain mass flux, which was in turn proportional to the friction
29      velocity cubed.  Dust emission is highly sporadic. After the wind stress threshold is
30      surpassed, the vertical flux increases with the third power of friction velocity.
31

        April 1995                                3_27       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Physical and Chemical Properties of Resuspended Particles
 2           The physical and chemical properties of resuspended particles depend partly on the
 3      properties of the particles that were deposited on the surface in the initial stage of
 4      resuspension.  But,
 5           "the deposited particles probably lose their individual identity by becoming
 6           attached to host (soil) particles.  When the pollutant particle is transported
 7           downwind, it is usually attached (aggregated) to this  host particle" (Sehmel,
 8           1978).
 9      Furthermore, the host particle is most likely an aggregate itself.  Studies of the cross section
10      of particles,  mineralogy, and scanning electron microscope analysis of dust samples show
11      that particles suspended from the soil are aggregated.  For these  reasons,  this section
12      describes physical properties of the aggregated (host plus pollutant) particles.
13           The size distribution of resuspended soil particles may be described  as lognormal
14      bimodal with one mode at 2 to 5 micrometers and another mode  at 30 to  60 micrometers
15      (Sviridenkov et al.,  1993;  Patterson and Gillette, 1977; Gillette and Nagamoto,  1993;
16      Gillette, 1974).  Because the mass mode of the distribution for particles smaller than 10
17      micrometers is roughly at  2.5 micrometers, a rough approximation is that half the PM10
18      mass is smaller than 2.5 micrometers and half is larger. The parameter
19
                                               ^  <0.1                                   (3-5)
                                               u*
20
21      defines the upper size of suspended dust, where vsed is the sedimentation velocity of the
22      upper size limit, and u*  is friction velocity. Data from Pinnick (1985) shows that very
23      similar size distributions result from resuspension by traffic.
24           Mineralogically (chemically) the dust consists of (in order of the most abundant) for
25      particles 1 to 10 micrometers:  quartz, mica, kaolinite, mixed layer phyllosilicates and
26      feldspars.  For particles smaller than 1 micrometers:  mica,  kaolinite, quartz, and mixed layer
27      phyllosilicates (Gillette et al., 1978).   Studies of elemental composition show that
28      composition of the resuspended material compared to that of the total  sediment is enriched in
        April 1995                                3-28      DRAFT-DO NOT QUOTE OR CITE

-------
 1      elements associated with the smallest particles (i.e., titanium) and impoverished in elements
 2      associated with the coarsest materials (i.e., silicon).
 3
 4      Levels of Production and Transport of Resuspended Aerosols
 5           Airborne dust measurements in the southern and central Great Plains states of the
 6      United States were made  in the early 1970's. The total mass of dust produced by individual
 7      dust storms was 300,000  to 500,000 million Tg (Gillette et al.,  1978).  Individual dust storm
 8      production rates may be compared to the global production rate estimated by d'Alameida
 9      (1989) of  1,800 to 2,000  Tg per year.  The Great Plains study, part of a severe storm study,
10      showed that  the dust storms were typically associated with vigorous frontal activity,  and that
11      the dust travels great distances (many 100's of km) as tracked by jet aircraft.  Estimates of
12      transport  distance for dust of well over 1,000 km (from West Texas dust storms to deposition
13      sites  in northern Minnesota) were supported by isentropic trajectories, positions of rainclouds
14      and elevated concentrations of calcium in collections of rainwater in the National Acid
15      Deposition Program/National Trends Network.  Even greater transport distances of
16      resuspended  dust are  shown by oxygen isotopic 18 to 16 ratios (618) in quartz (parts per
17      thousand). By matching the S18 value for deposited quartz and source areas for the quartz
18      (wind credible soils) the following long-range transport paths were found:  Asian deserts to
19      Hawaii; Sahara desert to  the Caribbean, South America, and Florida; and U.S.  sources to
20      Greenland and northern Europe (Jackson et al., 1973).
21           A model developed  for national acid rain and decertification/paleoclimate studies
22      (Gillette and Passi, 1988) expressed the emission of dust for a given study area  as an integral
23      over  friction velocity (expressing the forcing function), and the threshold friction velocity
24      (expressing the resistance of the soil and environment to ablation). Results from the model
25      for the contiguous United States (Figure 3-7) show a strong agreement  of the model dust
26      emissions  with known dusty areas (Gillette and Hanson,  1989).  Predicted alkaline emissions
27      also agree in many respects with observed wet deposition patterns of alkaline elements
28      (Gillette et al., 1993).  A considerable fraction of wind emitted dust is  from dust devils
29      (Gillette and Sinclair,  1989).
        April 1995                                3-29      DRAFT-DO NOT QUOTE OR CITE

-------
                               10   15   20   25   30   35   40   45   50   55   60
                                                                               45
                            >10'3g/cm2
                            10-4-1Q-3g/cm2
                      0    5   10   15    20   25   30   35   40   45   50   55   60
       Figure 3-7. Model dust emissions for the United States.

       Source: Gillette and Manson, 1989.
 1     3.2.3   Particle Removal Mechanisms and Deposition
 2          Particles in the air are in constant motion.  They are subject to Brownian motion, which
 3     is the constant random movement along an irregular path caused by the bombardment by
 4     surrounding air molecules.  This process is most important for small particles, and is related
 5     to the particle diffusion coefficient. Particles are also subject to the earth's gravitational
 6     force, as characterized by a sedimentation velocity. Gravitational settling is most important
 7     for larger particles.  Both of these processes involve the motion of the particle relative to its
 8     surrounding air medium.
 9          Brownian diffusion is important for small particles, whereas gravitational settling is
10     important for large ones.  During a time period of 1 s a 0.1 pun particle will travel a distance
11     of about 40 pirn from Brownian motion, while it will fall about 1 pun due to gravity.   In the
12     same 1  s time period a 1 jum particle  will diffuse about 8 jum and will fall 35 pirn.  Note that
13     the diffusion constant is directly proportional to the particle mobility B, while the settling
        April 1995
3-30
DRAFT-DO NOT QUOTE OR CITE

-------
1
2
3
4
velocity depends on the product of particle mass and mobility, mB.  Diffusion constants and
settling velocities are plotted in Figure 3-8.
                            0.01
       Figure 3-8.  Diffusion constants and settling velocities for particles.
1           The deposition of particles in the atmosphere is not easily modeled, and is characterized
2      by a deposition velocity, which is defined as the ratio of the flux of particles to the surface to
3      the ambient concentration.  Results from wind tunnel studies, shown in Figure 3-9, show
4      characteristic minima.  Small particles are collected by diffusion, larger particles are
5      collected by  impaction and sedimentation.  Deposition models which account for these
6      mechanisms  are given by Sehmel (1982), Fernandez de la Mora and Friedlander (1982) and
7      Fernandez de la Mora (1986).  Atmospheric data from Lin et al.  (1994), shown in
8      Figure 3-10 show that inertial mechanisms,  as well as sedimentation,  are important for the
9      deposition of large particles.
       April 1995
                                          3-31
DRAFT-DO NOT QUOTE OR CITE

-------
                       10
1
*
                       10"
                          10
                                   Grass about 10cm high
                                   (Chamberlain, Clough, Little)
                                   Filter paper (Clough)
                                   Smooth surface (Sehmel)
                10~1          1           10

                    Particle Diameter (urn)
Figure 3-9.  Particle deposition from wind tunnel studies.
         10'
 •a


 lio-*
 i


 2 10"
 x
 £
   10-
                        10
                               o Flux      • Mass
                               7 Calculated deposition velocity
                                                                   101
                                                                   10'
                                                                   10'3
                          0.1
                  1           10
                 Particle diameter, urn
100
                                                                       I
                                                                       EE
                                                                       8
o>
O
Figure 3-10. Sedimentation and inertia effects on large particle deposition.
April 1995
                        3-32        DRAFT-DO NOT QUOTE OR CITE

-------
 1          This section examines the present state of knowledge regarding the loading, size
 2     distribution, and chemical composition of the fine particle component of tropospheric
 3     aerosols and the processes that govern these properties.  Reasons why it is necessary to know
 4     the physical properties and composition of these aerosols include the following:
 5          1.    Identification of the processes  and sources responsible for the aerosol.
 6          2.    Development of a data base of measurements to be used in setting standards
 7          3.    Relating aerosol loadings and composition to putative deleterious consequences,
 8                e.g. epidemiological studies.
 9          4.    Development of models whose objective is to relate aerosol loading and chemical
10                and physical properties to sources of the aerosols and their precursors and
11                evaluation the performance of such models.
12
13
14     3.3 CHEMISTRY AND CHEMICAL COMPOSITION
15     3.3.1  Fine Particle Chemistry
16     3.3.1.1 Acid Aerosols and Paniculate Sulfates
17          Sulfuric acid and its  neutralization products with ammonia constitute a major
18     anthropogenic contribution to fine particle  aerosol.  This section reviews recent advances in
19     understanding of the sources, removal processes, loadings and properties of tropospheric
20     sulfate aerosols. Emphasis is given to properties and processes pertinent to these aerosols in
21     regions influenced by anthropogenic emissions as distinguished from remote locations
22     influenced primarily by natural sources.
23
24     Sources
25          Aerosol sulfate in the troposphere consists of particles  emitted directly from sources
26     (primary sulfate) and of sulfate formed by  atmospheric oxidation of gaseous sulfur
27     compounds, mainly SO2.  Knowledge of the sources of this  particulate material is important
28     to understanding the processes responsible  for observed loading, composition, and size
29     distribution of the  material and to developing effective methods to control its concentration.
30     Principal sources of ambient sulfate may be distinguished into primary emissions (that is
31     material emitted into the atmosphere as particulate sulfate or as gas-phase SO3 and/or H2SO4,

       April 1995                               3.33       DRAFT-DO NOT QUOTE OR CITE

-------
 1     both of which readily form condensed-phase species) and gas-to-particle conversion in the
 2     atmosphere that produces SO3 and/or H2SO4.
 3
 4     Atmospheric oxidation of SO2
 5          Atmospheric oxidation of SO2 takes place both by gas-phase reaction and by aqueous-
 6     phase reaction. The principal gas-phase mechanism is thought to be the OH-initiated
 7     reaction.  The principal aqueous-phase reactions are thought to be oxidation by H2O2 and O3.
 8     Aqueous-phase reactions followed by cloud evaporation can result in formation of clear-air
 9     aerosol.  Evaporation can be a major production route for atmospheric sulfate aerosols.  The
10     relative proportion of sulfate aerosol produced by the aqueous and gas-phase routes is not
11     well established.
12
13          Gas-phase oxidation of SO2. Gas phase oxidation of SO2 is thought to occur largely,  if
14     not entirely, by a sequence of reactions initiated by the reaction of OH  with SO2.
15
                                      S02 + OH + M -»  HS05 + M                          (3-6)
16
                                       HSO 5 + O2 -* SO3 + HO2                            (3-7)
17
                                                                                         (3-8)
18
19     The gaseous H2SO4 subsequently adds to existing particles or may nucleate to form new
20     particles.
21          Until recently the evidence for the occurrence of this reaction in the atmosphere has
22     relied on modeled OH concentrations and on laboratory-determined reaction rate coefficient
23     (Gleason et al., 1987) for the OH + SO2 reaction.  However, recent measurements of OH
24     and H2SO4 in the atmosphere provide empirical evidence for this mechanism (Eisele and
25     Bradshaw, 1993;  Eisele and Tanner, 1993).  Simultaneous measurements of OH and SO2
26     allow the  gas-phase reaction production rate of H2SO4 to be calculated at the time and
27     location of the measurement. Likewise, measurements of particle size distribution allow the
28     effective first-order rate coefficient for diffusive uptake of H2SO4 monomer by aerosol

       April 1995                               3.34      DRAFT-DO NOT QUOTE OR CITE

-------
  1      particles to be calculated, and measurement of the concentration of H2SO4 monomer allows
  2      the loss rate by this mechanism to be calculated. Comparison of the calculated production
  3      and loss rates of H2SO4 monomer show them to be equal, consistent with the observed steady
  4      state concentration of this species. This study lends substantial confidence to the applicability
  5      of the laboratory mechanism and rate to evaluation of the rate of sulfuric acid formation in
  6      the ambient atmosphere.   At night, however,  the calculated  loss rate substantially exceeded
  7      the apparent production rate.  This suggests an additional source of H2SO4 monomer, either
  8      from some hitherto unrecognized reaction, or from release of H2SO4 from the particles back
  9      to the gas  phase.  The investigators suggest that the explanation is the latter, in view of
10      correlation of particle concentration and H2SO4 monomer concentration.
11           The substantial progress made in the past few years in measurement of OH leads to the
12      expectation of increased confidence in models that calculate the concentration of this species
13      from local photolysis rate constants and abundances.  This may be expected to lead in turn to
14      enhanced confidence in OH concentrations and sulfuric acid production rates calculated by
15      regional scale transport models.
16           The reaction of SO3 has recently been reexamined by Kolb et al. (1994), who find that
17      the reaction is second order in water vapor and propose that the reaction takes place by
18      interaction of SO3 with water vapor dimer:
19
                                     S03 + (H2O)2 - H2SO4 + H2O                         (3-9)
20
21           The investigators note that it is probable that sufficient water dimer exists in the
22      atmosphere to allow the reaction to efficiently form sulfuric acid vapor.  Other processes
23      may involve H2SO3  + H2O.  The complex H2O x SO3 also may be involved in sulfuric
24      formation (Leopold et al., 1985).
25
26      3.3.2  Formation of  Sulfates in Clouds
27      3.3.2.1  Particle Formation in Clouds
28      3.3.2.LI  Introduction
29          The atmospheric aqueous phase (clouds, fogs) can be viewed as  a processor of the
30      aerosol size/composition  distribution (Pandis et al., 1990a,b).  Precipitating clouds are well

        April 1995                               3.35       DRAFT-DO NOT QUOTE OR CITE

-------
 1     known to be the major removal mechanism of aerosol particles from the atmosphere.  At the
 2     same time, the liquid droplets provide the reacting medium for aqueous-phase reactions
 3     (Graedel and Weschler, 1981;  Chameides and Davis, 1982;  Graedel and Goldberg, 1983;
 4     Jacob and Hoffmann, 1983; Munger and Hoffman, 1983; Chameides, 1984; Seigneur and
 5     Saxena, 1984; Hoffman and Jacob,  1984; Fuzzi et al., 1984; Hong and Carmichael,  1986;
 6     Hill et al., 1986; Jacob, 1986; Jacob et al., 1986; Johnson et al., 1987; Fuzzi et al.,  1988;
 7     Dlugi, 1989; Pandis and Seinfeld, 1989; Munger et al., 1990; Forkel et al., 1990; Bott,
 8     1991; Joos and Baltensperger,  1991; Earth, 1994; De Valk,  1994). Several gaseous  species
 9     dissolve in cloudwater  and  react giving products that remain in the aerosol phase after the
10     cloud dissipates; for example, the dissolution of SO2, its ionization, and subsequent oxidation
11     to sulfate.  These species can attract additional gaseous species, such as ammonia and water
12     into the aerosol phase and thereby increase  further the aerosol mass.  Therefore, aerosol
13     processing by nonprecipitating clouds represents a mechanism by which atmospheric  particles
14     can grow during their residence time in the atmosphere.  A  detailed review of the state of
15     science in 1990 has been presented by United States National Acid Precipitation Assessment
16     Program (U.S. NAPAP) (1991).
17          A cyclical relationship between the occurrence of smog and fog in polluted areas has
18     been proposed by Munger et al. (1983) and was termed the  smog-fog-smog cycle.  In a
19     polluted atmosphere with high aerosol concentration, the formation of late night and early
20     morning  fogs is augmented enhancing smog production, visibility reduction, and aerosol
21     sulfate the next day (Cass,  1979; Cass and Shair, 1984; Pandis et al., 1990). Processing of
22     aerosol by clouds can result in similar cyclical relationships  and enhanced contribution of the
23     aerosol produced in clouds  to ground-level paniculate concentrations (Altshuller, 1987).
24     This processing cycle accelerates the production of atmospheric acidity through aqueous-
25     phase reactions (Schwartz,  1989).
26
27     3.3.2.1.2  Cloud Effects On Particle Number Concentration
28          There has been a series of observations of enhanced aerosol number concentrations  in
29     the vicinity of clouds (Saxena and Hendler, 1983; Hegg et al., 1990; Radke and Hobbs,
30     1991; Hegg et al.,  1991).  Saxena and Hendler (1983) suggested  that the observed high
31     aerosol number concentrations near  clouds could be due to the shattering of rapidly

       April 1995                               3_36      DRAFT-DO NOT QUOTE OR CITE

-------
 1      evaporating droplets.  Hegg et al. (1991) proposed that the high actinic radiation fluxes near
 2      cloud tops could lead to high OH concentrations and nucleation of new H2SO4/H2O particles.
 3      The high relative humidity areas around clouds  often have total particle number
 4      concentrations about twice those in the air at the same level but well removed from the cloud
 5      boundaries (Radke and Hobbs, 1991).  Kerminen and Wexler (1994) have demonstrated that
 6      there is high nucleation probability associated with these high relative humidity areas,
 7      especially near relatively clean clouds.  All these speculated mechanisms for production of
 8      new particles produce negligible new aerosol mass, but may influence the shape of the
 9      aerosol distribution, especially in remote regions. Aqueous-phase reactions producing sulfate
10      and nitrate increase the aerosol mass, but do not influence directly the aerosol number
11      concentration. The removal of gas-phase SO2,  H2SO4, and NH3,  due to their transfer to
12      aqueous-phase, indirectly slows down the production of new particles in the vicinity the
13      cloud.
14
15      3.3.2.1.3 Cloud Effects On Aerosol Mass Concentration
16           Significant  production of sulfate has been  detected in clouds  and fogs in different
17      environments (Hegg and Hobbs, 1987, 1988; Pandis and  Seinfeld, 1989b;  Husain et al.,
18      1991; Swozdziak and Swozdziak, 1992; Pandis  et al., 1992; De Valk, 1994; Liu et al.,
19      1994). The detection of sulfate-producing reactions is often hindered by the variability of
20      cloud liquid water content and the temporal instability and spatial  variability  in concentrations
21      of reagents and product species (Kelly et al., 1989).  The production of sulfate has also been
22      detected and investigated in laboratory clouds (Hansen et  al.,  1991).  Aqueous-phase
23      oxidation of HSO3"  by H2O2 is particularly fast, as illustrated  by the mutual exclusivity of
24      SO2 and H2O2 observed in clouds (Daum et al., 1984;  1987).  Other reactions, including
25      oxidation of dissolved SO2 by ozone and oxidation by O2 catalyzed by Fe3+  and  Mn2+ may
26      also contribute, significantly in some cases, to sulfate production (Pandis et al.,  1989; Earth
27      et al., 1992; Earth,  1994).  During aqueous-phase sulfate production the reactants including
28      SO2, H2O2, O3,  and OH are transferred from the gas-phase to the cloud droplets.  This
29      transport includes a series of steps (gas-phase diffusion, transport  across the gas-liquid
30      interface, dissociation and aqueous-phase diffusion) that ultimately couple the gas and
        April 1995                                3.37       DRAFT-DO NOT QUOTE OR CITE

-------
 1      aqueous phases and in some cases control the overall sulfate production rate (Schwartz,
 2      1988).
 3           Hydrogen peroxide is the most important oxidant for the conversion of SO2 in cloud
 4      water at pH 4.5 or lower (Calvert et al., 1985) and dominates the aqueous sulfate formation
 5      pathways (McHenry and Dennis, 1994) in the northeastern United States.  The measured
 6      H2O2 gas-phase mixing  ratio over the northeastern and central United States has been
 7      reported to vary from 0.2 to 6.7 ppb (Sakugawa et al.,  1990) with the highest values during
 8      the summer and the lowest during the winter months.  The H2O2 concentrations usually
 9      increase with decreasing latitude and increasing altitude (Sakugawa et al., 1990). The
10      availability of hydrogen peroxide is often limiting the sulfate formation in clouds.  This
11      limitation is more pronounced near SO2 sources and during the winter months.  The seasonal
12      contribution of clouds to sulfate levels depends on both the availability of oxidants and on the
13      cloud cover.  In cases where the sulfate cloud production is oxidant limited, changes in
14      aerosol sulfate levels will be less than proportional to SO2 emission changes, with the
15      relationship being more nonlinear in winter than in spring or summer (U.S. NAPAP,  1991).
16          Evaluations of the rate of the SO2-H2O2 reaction in cloudwater indicate that the
17     characteristic time for this reaction is a few minutes to an hour, depending on conditions
18     (Schwartz, 1984; Meagher et al., 1990). Since such a reaction time is shorter than the
19     lifetime of stratiform clouds in the troposphere it  is anticipated that the reaction of SO2 and
20     H2O2 will proceed to completion in liquid water stratiform clouds. Evidence of this
21      occurring would be that only one or the other of these species would be present in such
22     clouds, but not both at the same time.  This expectation has been borne out in field
23     measurements supporting the inference of rapid reaction given by the model estimates.
24     Daum and colleagues (Daum et al,  1984; Daum,  1988) have presented results of
25     simultaneous aircraft measurements of H202 in collected cloudwater samples and SO2 in air
26     (filter pack measurements) in nonprecipitating stratiform clouds indicating that in almost all
27     instances either one or the other species was at very low concentrations, and by inference
28     that the reaction has proceeded essentially to completion in the clouds. A rather different set
29     of results was reported  by Husain et al. (1991) who conducted measurements of gas-phase
30     SO2 and H2O2 during cloud events  at Whiteface Mountain, NY.  Although a general  negative
        April 1995                               3-38       DRAFT-DO NOT QUOTE OR CITE

-------
  1     correlation between the two species concentrations was exhibited, the data indicated
  2     substantial periods of apparent coexistence of these species.
  3          There is the possibility of spatial inhomogeneities in the clouds that are not resolved in
  4     the sampling period (typically 30 min in the Daum studies; an hour or more for the Husain
  5     studies), in which one region was H2O2 rich and another SO2 rich.  In such instances a lack
  6     of coexistence of the two species would be masked by the extended duration of sampling.
  7     Such spatial inhomogeneities might also account for the  few instances reported by Daum  in
  8     which SO2 and H2O2  apparently coexisted in clouds. Additionally,  local patches of
  9     subsaturated air in the clouds during the sampling period might also account for these
 10     observations, although Daum took efforts to exclude such instances  from their data base.
 11     Yet another possible explanation of the Husain results is that the cloud  was relatively newly
 12     formed, and the material had not had time to react. An obvious improvement in this
 13     approach is to measure the species, as well as cloud liquid water content, with greater time
 14     resolution. Burkhard  et al.  (1994) have present aircraft  measurements of gas-phase SO2 and
 15     H2O2 during  in-cloud  flights; traces of liquid water content are also shown.  These data
 16     support  a strong anticorrelation of SO2 and H2O2 in clouds  on various time (location) scales,
 17     with numerous instances of peaks of S02 coincident with valleys of H2O2 and vice versa.
 18          A  quantitative estimate of the amount of cloudwater sulfate that is formed by in-cloud
 19     reaction can be gained by inferring the amount of cloudwater sulfate that derives from
 20     preexisting sulfate aerosol.   Husain et al. (1991) has used selenium as a tracer to allow such
 21     inferences to  be drawn. By measuring the sulfate to selenium ratio  in clear air aerosol that is
 22     representative of the aerosol that is the pre-cloud aerosol of the clouds under investigation,
 23     and by assuming that the fractional incorporation of the sulfate and selenium into cloudwater
 24     is identical (and/or by measuring this ratio),  it is possible to infer the amount of cloudwater
 25     sulfate derived from preexisting sulfate aerosol and by difference, the amount formed by in-
 26     cloud reaction.  A series of  such studies carried out at Whiteface Mountain, NY, indicates
27     that assumption of identical scavenging of sulfate and selenium is valid  (1.04 + 0.29; 1.04
28     ± 0.19 in two separate cloud systems).  Evidence of enhanced sulfate in cloudwater,
29     attributed to sulfate formed by in-cloud reaction, was found in five of six cloud systems
30     studied;  amounts formed were consistent with ambient SO2 concentrations.  Examination of
31      the pH dependence of the concentration of in-cloud produced sulfate inferred by this

        April 1995                                3.39       DRAFT-DO NOT QUOTE OR CITE

-------
 1      technique indicated that sulfate was produced by in-cloud reaction only at pH values below
 2      4.0, consistent with oxidation by H2O2, but not with oxidation by O3.
 3           Recently  Snider and Vali (1994) reported studies of oxidation of SO2 in winter
 4      orographic clouds in which SO2 was released and the extent of increased concentrations of
 5      sulfate in cloudwater (relative to the unperturbed cloud) were compared to decreased
 6      concentrations  of H2O2 (sum of gaseous plus aqueous, inferred from aqueous concentrations).
 7      Despite considerable scatter, the data fall  fairly close  to the one-to-one line, indicative of the
 8      expected stoichiometry of reaction, Figure 3-11.  The investigators also modeled the reaction
 9      kinetics.  The rate of reaction is sensitive to the liquid water content (LWC) of the cloud
10      during the time between the point of cloud condensation to the point of sampling.  Since this
11      profile was not known the investigators assumed a linear profile for LWC versus time.  The
12      resulting model predictions agreed closely with the extent of reaction inferred from changes
13      in H2O2 and sulfate concentrations, supporting the applicability of the  model.
14           In contrast to the H2O2 reaction, oxidation of SO2 by O3 exhibits a strong pH
15      dependence. The reaction is quite rapid at high pH (~6) but  is expected to greatly slow
16      down as strong acid is produced in the course of the reaction.   However, if concentrations of
17      NH3 or other basic materials are sufficiently high  to maintain a pH above 5, the reaction can
18      continue to proceed at a high rate.
19           Walcek et al. (1990) calculated that, during the  passage of a midlatitude storm system,
20      over 65 % of the sulfate in the troposphere over the northeastern United States was formed in
21      cloud droplets  via aqueous-phase chemical reactions.  The same authors indicated that,
22      during a 3-day springtime period, chemical reactions  in clouds occupying 1 to 2% of the
23      tropospheric volume were  responsible for sulfate production comparable to the gas-phase
24      reactions throughout the entire tropospheric volume under consideration.  McHenry and
25      Dennis (1994)  estimated that annually more than 60% of the ambient sulfate in Central and
26      Eastern United States is produced in clouds with the non-precipitating cloud production
27      dominating over precipitating  cloud production. Spatial variability of emissions and ambient
28      H2O2 concentrations induces spatial variability in the  contribution of in-cloud sulfate
29      production, making it highest in the south.  These conclusions are in quantitative agreement
30      with similar calculations of Dennis et al.  (1994) and Karamachdani and Venkatram (1992).
        April 1995                               3-40      DRAFT-DO NOT QUOTE OR CITE

-------
                        0.25 -
                        0.20 -
                        0.15 -
                     m
                     O
                     >
                     f  0.10
                        0.05 -
-



0.00
L 1 | 1 1 1 1 | 1 1 1
0.10
0.05

-<

0.00
2 D94
F+
D66
_

-
.•
o
-
8F

— <
c

0.00
0.05 0.10 0.15
% . Q94 -
8B 8J
o I
84
0.05 0.10
-
•
•
™
0.20 0.25
DH,O, (PPbv), OBSERVATION
                        0.00 -    _ _
      Figure 3-11.  Comparison of observed H2O2 depletions (D^O^ abscissa) and observed
                    sulfate yields (YSO4, ordinate). Errors associated with experiments 84, 8B,
                    8F,  and 8J are indicated and data values from these experiments are
                    labeled in the inset figure. Data values corresponding to experiments 94
                    and 66 are also labeled.  The slope of the best fit line, forced through the
                    origin, and calculated using only those data values indicated by circles is
                    1.21 (±0.13).
      Source: Snider and Vali (1994).
1     Aqueous-phase oxidation in clouds is also the most important pathway for the conversion of
2     SO2 to sulfate on the global scale (Hegg, 1985; Langner and Rodhe, 1991).
3          Clouds could under some conditions also be a significant source of aerosol nitrate
4     during the night.  Choularton et al. (1992) and Colvile et al. (1994) observed production of
5     around 0.5 mg m~3 of nitrate during the processing of an air parcel by a cloud.  They
6     speculated that the sources of this nitrate were gaseous N2O5 and NO3.
7          Chemical heterogeneities in the droplet population affect significantly the overall  sulfate
8     production rate and the produced sulfate size distribution (Seidl, 1989; Twohy et al., 1989;
9     Lin and Chameides, 1990; Pandis et al., 1990a,b;  Ayers and Larson, 1990; Hegg and
      April 1995
3-41
DRAFT-DO NOT QUOTE OR CITE

-------
 1      Larson,  1990; Bower et al., 1991; Ogren and Charlson, 1992; Roelofs, 1992a,b; 1993;
 2      Carter and Borys, 1993; Bott and Carmichael, 1993; Collett et al., 1993b).  Neglecting these
 3      chemical concentration differences could result in significant underestimations of the sulfate
 4      production rates in some cases (Hegg and Larson, 1990; Roelofs,  1993).  Ice-related
 5      microphysical processes can also have a significant impact on cloud chemistry (Taylor, 1989;
 6      Wang and Chang, 1993; Collett et al.,  1993a).
 7           Fogs in polluted environments have the potential to increase aerosol sulfate
 8      concentrations but at the same time to cause reductions in the aerosol concentrations of
 9      nitrate,  chloride, ammonium and sodium as well as in the total aerosol mass concentration
10      (Pandis et al., 1990a). Pandis et al.  (1992) calculated that more than half of the sulfate in a
11      typical Los Angeles air pollution episode was produced inside a fog layer the previous night.
12      This heterogeneously produced sulfate represented 5 to 8% of the measured PM10 mass.
13
14      3.3.2.1.4 Cloud Effects On Aerosol Size/Composition Distribution
15           Several measurements of the  aerosol mass distributions in urban areas have shown that
16      two distinct modes can exist in the 0.1  to 1 /^m diameter range (Hering and Friendlander,
17      1982; McMurry and Wilson,  1983; Wall et al., 1988; John et al.,  1990).  These are referred
18      to as  the condensation mode (approximate aerodynamic diameter 0.2 jum) and the droplet
19      mode (aerodynamic diameter around  0.7 /mi).  Hering and Friedlander (1982) and John et al.
20      (1990) postulated that the larger mode could result from aqueous-phase chemistry.   Meng and
21      Seinfeld (1994) proposed that growth of condensation mode particles by accretion of water
22      vapor or by gas-phase or aerosol-phase sulfate production cannot explain the existence of the
23      droplet mode. Activation of condensation mode particles, formation of cloud/fog drops
24      followed by aqueous-phase chemistry, and aqueous droplet evaporation was shown by these
25      authors to be a plausible mechanism  for formation of the urban and regional aerosol droplet
26      mode. The sulfate formed during fog/cloud processing of an air mass favors the aerosol
27      particles that had access to most of the fog/cloud liquid water content, which are usually the
28      particles with dry diameters around 1 /mi (Pandis et al., 1990b).  These two submicron
29      mass-distribution modes have been also observed in  non-urban continental locations
30      (McMurry and Wilson, 1983; Hobbs et al., 1985; Radke et al., 1989), but the frequency of
31      their  co-existence remains unknown.  Thus, cloud processing of an air parcel can clearly

        April 1995                               3-42      DRAFT-DO NOT QUOTE OR CITE

-------
  1      impact the scattering efficiency and in general the radiative properties of the corresponding
  2      aerosol (Hegg et al., 1992; Bower and Choularton, 1993).
  3           The aerosol distribution is also modified during in-cloud processing by collision-
  4      coalescence of droplets and impaction scavenging of aerosols (Pruppacher and Klett, 1980).
  5      The aerosol scavenging by droplets is a relatively slow process, and collision coalescence
  6      among droplets of different sizes causes a redistribution of aerosol mass in such a manner
  7      that the main aerosol mass is associated with the main water mass (Flossmann et al.,  1985).
  8      The processing of the remote aerosol distribution by clouds has been clearly demonstrated  in
  9      a series of field studies (Frick and Hoppel,  1993).  This multiple processing of remote
 10      aerosol by nonprecipitating clouds results in an extra mode in the aerosol number distribution
 11      (Hoppel et al., 1986; Frick and Hoppel, 1993).
 12           Clouds and fogs can influence the atmospheric aerosol number and mass concentration
 13      and chemical composition, the shape of the aerosol size distribution, the aerosol acididity and
 14      radiative properties.  These effects can be important in all environments (urban, rural and
 15      remote) and all seasons.  Our qualitative understanding of the aerosol-cloud interactions has
 16      improved significantly, but, with few exceptions, the quantification of these effects remains
 17      uncertain (Altshuller, 1987; Kelly et al., 1989; Pandis et al., 1992).
 18
 19      3.3.3  Aqueous-Phase Oxidation Of SO2 In Clear-Air Aerosols
 20           Until recently it was thought that the low amount of liquid water associated with clear-
 21      air aerosols (volume fraction on the order of 1 x 10-10 , compared to clouds, for which the
 22      volume fraction is the order of 1 x 10-7) precluded significant aqueous-phase conversion of
 23      SO2 in such droplets.  However Sievering and colleagues  have called attention to the
 24      possibility of rapid rate of oxidation of SO2 by O3 in aqueous sea-salt aerosols, which are
 25      buffered by the alkalinity of sea salt particles,.  Indeed it  appears that such a rate may
26      initially be quite rapid, 1  /jM s-1 corresponding to 8% hr-1, in the example given by
27      Sievering et al. (1991) for liquid water content 50 /xg m-3 and SO2 concentration 2 n mol m-
28      3  (mixing ratio 0.05 ppb). Despite this rapid initial rate, it would appear that the extent of
29      such oxidation may be quite limited.  For the example given by Sievering et al. (1991), the
30      sea-salt sodium concentration is given as 100 n mol m-3.  Based on the concentrations of
31      (HCO3~ + CO32") and Na+ in seawater (2.25 and 454 m  mol kg-1, respectively), the

        April 1995                                3.43      DRAFT-DO NOT QUOTE OR CITE

-------
  1      alkalinity of the sea salt aerosol is expected to be 0.5 n mol m-3.  Consequently, after only
  2      0.25 n mol m-3 of SO2 is taken up in solution and oxidized (i.e., 12% of the initial SO2), the
  3      initial alkalinity would be exhausted, and the reaction rapidly quenched.
  4           Sievering et al.  1994 have presented field measurements over Lake Michigan of coarse-
  5      mode sulfate (diameter 5-20 /xm), which they ascribe at least in part to oxidation of SO2 in
  6      such particles derived from wind driven spray  of lake water,  in which the pH is maintained
  7      high by alkalinity present in the lake water.  Calculations were carried out for liquid water
  8      volume fraction of 13 x 10-12 (13 /^g m-3).  The alkalinity was inferred from the measured
  9      cation minus anion difference (cations NH4 + , Mg2 + , Ca2 + ; anions SO42-, NO3-) in the
10      coarse mode,  which averaged 26 neq m-3, corresponding to an aqueous alkalinity of 2 x 10-5
11      M.  In the absence of mass transport limitation the rate of the aqueous-phase O3-SO2 reaction
12      was calculated to be 7 ± 3 x 10-4 M  s-1; however, mass transport limitation reduced this
13      rate by a factor  of 20 to 40 at pH 7.   The conversion rate referred to  gas-phase SC^ was
14      calculated as 0.5 to 1.7 % hr-1.  The  investigators concluded that this mechanism is a
15      significant contributor to the SO2 oxidation under these conditions. Again, however, concern
16      may be raised with that conclusion, namely that the indicated oxidation rate, 2 x 10 -5 M s-1
17      after taking mass transport limitation into account, would quickly produce an acidity equal to
18      the initial alkalinity, thereby quenching the reaction.
19
20      3.3.3   Physical and Chemical Considerations  in Particulate  Sampling and
21              Analysis
22      3.3.3.1 Semi-Volatile Organic Compounds (SOCs)
23      General
24       SOCs are defined here to be organic compounds with intermediate pure compound, sub-
25      cooled liquid vapor pressures (p^). Definition ranges vary somewhat, but SOCs can be
26      thought to include compounds with/j£ values in the range 10"2 down to 10"9  torr.  For SOCs
27      and for semi-volatile inorganic materials,  there are health and sampling reasons for under-
28      standing the factors controlling the relative amounts that are  in the gaseous (G) and aerosol
29      particulate  (P) phases.  G/P partitioning of SOCs has often been considered to involve mainly
30      simple physical adsorption to particle surfaces (e.g., Junge, 1977; Yamasaki et al., 1982;
31      Pankow, 1987).   However, absorptive phase partitioning must also often play some role

        April 1995                               3.44       DRAFT-DO NOT QUOTE OR CITE

-------
  1      (Pankow, 1994) as into aerosol particles containing plant wax as well as organic carbon from
  2      primary emissions (Turpin et al.,  1991) and from secondary organic aerosol formation
  3      (Pandis et al., 1992).
  4
  5      Theory
  6           A useful parameterization of G/P partitioning is (Yamasaki et al., 1982; Pankow, 1991)

                                           *,  -                                         (3-9)
 8      where:  K (m3 /zg"1) = partitioning constant; TSP (/jg m~3) = concentration of total
 9      suspended paniculate matter; and F (ng m"3) and A (ng m"3) = the P-associated and G
10      concentrations of the compound of interest, respectively. The symbols F and A originate in
11      the common usage of a filter followed by an adsorbent to collect the P and G portions,
12      respectively. With urban particulate matter (UPM), a given SOC at a given temperature T
13      tends to exhibit similar K values from sampling event to sampling event.  The fraction 0 of
14      the total compound that is on/ in the P phase is given by
                                          A + F      Kp TSP  + 1
15
16      Though not yet used in practice, it may also prove useful to define Kp 10 = (F10 / PM10 ) / A
17      where PM10 (pg m"3)  = concentration of particles with aerodynamic diameters smaller than
18      10 pirn, and F10  (ng m"3) = PM70-associated concentration of the compound of interest.
19           Theory (Pankow, 1994) predicts that the values of Kp for a given compound class will
20      be given by a relation of the form Kp  = [C1 + C2] I p£ , where C;/p£ and C2/p£ represent
21      the adsorptive and absorptive contributions to Kp, respectively.  Log Kp values measured
22      under given conditions (e.g. , I) for a compound class such as the polycyclic aromatic
23      hydrocarbons  (PAHs) will thus tend to be linearly correlated with the corresponding log /?£
24      values according to log Kp  = mr log p£ + br  For PAHs sorbing to UPM in Osaka, Japan,
25      mr « -1.028 and br ~ -8.11 (Pankow and Bidleman, 1992). (Table 3-2 gives /?£ values for
26      several PAHs at 20 °C.) This correlation allows Kp to be predicted for a compound that is
27      within the compound class of interest, but was not examined in a given study.  K for a

        April 1995                               3.45      DRAFT-DO NOT QUOTE OR CITE

-------
                TABLE 3-2. VALUES OF LOG P£ FOR VARIOUS PAHS AT 20 °C
                                  Compound                                log /?£ (torr)
         Fluorene                                                             -2.72
         Phenanthrene                                                         -3.50
         Anthracene                                                           -3.53
         Fluoranthene                                                         -4.54
         Pyrene                                                              -4.73
         Benzo [a] fluorene                                                      -5.24
         Benzo[b]fluorene                                                      -5.22
         Benz[a]anthracene                                                     -6.02
         Chrysene                                                             -6.06
         Triphenylene                                                         -6.06
         Benzo [b] fluoranthene                                                  -7.12
         Benzo [kjfluoranthene                                                  -7.13
         Benzo[a]pyrene                                                       -7.33
         Benzo[e]pyrene                                                       -7.37
 1     given compound depends on T (Kelvin) according to log Kp = mpIT + bp where mp depends
 2     on the enthalpy of desorption; values of the intercept bp will be similar within a given
 3     compound class (Table 3-3). Increasing the relative humidity from 40 to 90% appears to
 4     cause K values to decrease by a factor of about two for PAHs sorbing to UPM (Pankow et
 5     al., 1993).
 6          For constant K , then  will increase as TSP increases. For constant TSP and T, as
 7     volatility increases (i.e., as /?£  increases), then Kp and <£ will decrease. When <£ « 0, one
 8     can sample just the G phase when determining the atmospheric concentration of an SOC;
 9     when   ~ I, one can sample just the P phase;  when  is between 0 and 1, one must sample
10     both phases.
       April 1995                               3-46      DRAFT-DO NOT QUOTE OR CITE

-------
          TABLE 3-3. mp Values for PAHs Sorbing to UPM in Osaka, Japan.  Obtained by
                            Fitting to a Common j-Intercept br of -18.48.
                                   Compound                                     mp
         Phenanthrene & Anthracene                                              4,124
         Methylphenanthrene & Methylanthracene                                  4,240
         Fluoranthene                                                           4,412
         Pyrene                                                                 4,451
         Benzo[a]fluorene & Benzo[b]fluorene                                     4,549
         Benz[a]anthracene, Chrysene, & Triphenylene                             4,836
         Benzo[b]fluoranthene & Benzo[k]fluoranthene                             5,180
         Benzo[a]pyrene & Benzo[e]pyrene                                        5,301
       Source: Pankow, 1991.
 1     Sampling Methods and Associated Sampling Artifacts
 2          Atmospheric SOCs can be determined using a filter followed by an adsorbent.  These
 3     collect the P and G portions, respectively.  Filter types include glass fiber filters (GFFs),
 4     quartz fiber filters (QFFs), and teflon membrane filters (TMFs).  Adsorbent types  include
 5     polyurethane foam (PUF), Tenax, and XAD resins.  Safe sampling volumes for G-phase
 6     SOCs on Tenax and PUF can be predicted based on studies of retention volumes on these
 7     adsorbents (Pankow, 1988 and 1989).  Volatilization losses from particles (i.e., "blow-off")
 8     can occur from a filter/adsorbent when T increases during sampling, when the general level
 9     of air contamination decreases during sampling, and/or when a large pressure drop develops
10     across the filter (Zhang and McMurry, 1991). In the first case, K for a given compound
11     and the already-filtered particles will decrease, leading to desorption from the sampled P-
12     phase.  In the  second case, even with T constant,  if A in the air being  sampled decreases,
13     then desorption losses from the collected particles can occur.  Volatilization is of particular
14     concern with long sampling times since large overnight T cycles and/or large changes in the
15     level of contamination are then more likely. Material volatilized from the filter will be
16     collected on the adsorbent following the filter. Adsorption gains to particles from the gas
17     phase due to decreases  in T and/or increases in A during sampling is a second possible

       April 1995                               3.47       DRAFT-DO NOT QUOTE OR CITE

-------
 1     artifact type with filter/adsorbent samplers. Adsorption to the filter from the G phase is a
 2     third artifact type.  In this last case, a portion of the value of A for an SOC of interest sorbs
 3     directly to the filter and so incorrectly contributes to the measured value of F for the
 4     compound.  It is difficult to generalize regarding the magnitudes of the first two artifact
 5     types.  One can attempt to correct for the third artifact type through the use of a backup-
 6     filter (Hart et al., 1994).  For sampling of UPM in Portland, Oregon,  Hart et al. (1994)
 7     estimated that failure to correct for gas adsorption to the filter caused F values for PAHs to
 8     be overestimated by a factor of —1.4.  Correction of the G-adsorption effect through the use
 9     of a backup filter is subject to two possible confounding effects:  1) the atmospheres to which
10     the front and back filters are exposed may differ, making for different G-adsorption to the
11     two filters; 2) organic compounds sorbed to a backup filter could have in part volatilized
12     from the front filter. Table 3-4 summarizes how the three artifact types act to cause
13     measured values of F, A, and  to deviate from the true, volume-averaged values.
14
15
           TABLE 3-4.  Effects of Three Types of Artifacts on Volume-Aver aged Values of 
                             Measured Using a Filter/Adsorbent Sampler
Artifact
Volatilization from collected particles
Adsorption to collected particles
Gas adsorption to filter itself
Artifact Effect
On A On F and 0
Too large Too small
Too small Too large
Too small Too large
 1          A sampler employing a diffusion denuder may avoid some of the artifact problems of
 2     filter/adsorbent samplers.  Air drawn into a diffusion denuder can be stripped of G-phase
 3     SOCs by a sorbent that coats the walls of the denuder:  G-phase SOCs diffuse from the core
 4     of the air flow toward the walls.  Sorbent coatings that have been used include silicones,  gas
 5     chromatographic stationary phases (Krieger and Kites, 1992 and 1994), and finely divided
 6     XAD resin (Gundel et al., 1994; Kamens et al., 1995).  The majority of the P-phase SOCs
 7     do not deposit on the walls of  the denuder because  aerosol particles have much smaller dif-

       April 1995                                3-48       DRAFT-DO NOT  QUOTE OR CITE

-------
  1      fusion coefficients than do gaseous molecules.  The particles exiting the denuder are collected
  2      on a filter.  Because the air stream flowing onto the filter has been largely stripped of G-
  3      phase SOCs, some desorption of the filtered P-phase SOCs can occur, and so an adsorbent is
  4      often placed after the filter to collect any such desorbed SOCs.  F for a given compound is
  5      taken as the sum of the amounts on the filter and the subsequent adsorbent.  Analysis of the
  6      denuder sorbent provides A.  When the denuder sorbent cannot be analyzed (as with silicone
  7      rubber), A can be determined by difference using a second, total (A +  F) determination for
  8      SOCs (Lane et al., 1988; Coutant et al., 1988 and 1992; and Eatough et al., 1989 and  1993).
  9      Although sampling artifacts are not often discussed for denuder-based samplers, artifacts
 10      cannot be assumed to be absent.  In the denuder section, less than 100% efficiency for G-
 11      phase collection will tend to  make measured A values too small and F and 0 values too large;
 12      greater than 0% efficiency for P-phase collection will tend to make measured A values  too
 13      large and F and <£ values too small. Turpin et al. (1993) have presented a new denuder
 14      design which does not use a  sorbent-coated wall.  Rather, a laminar flow separator is used to
 15      separate a portion of the G phase from a mixed G+P flow; collection of the G-phase
 16      compounds on a sorbent like PUF allows determination of the G-phase concentrations. P-
 17      phase concentrations are determined by difference.  Other sampling and analysis  issues are
 18      discussed in Chapter 4 of this document.
 19
20      3.3.4  Particulate Nitrates
21      3.3.4.1  Sources
22           By analogy to the sulfur system sources of aerosol nitrates might be distinguished into
23      primary, gas-phase, and aqueous-phase.  However, as primary nitric acid emissions are
24      considered to be small, the present discussion focuses on in-situ production mechanisms in
25      the atmosphere.
26
27      3.3.4.2  Gas-phase
28           The principal mechanism for gas-phase production of nitrates is reaction of OH with
29      NO2 to  form HNO3.
       April 1995                                3.49      DRAFT-DO NOT QUOTE OR CITE

-------
                                       OH+NO2 + M -> HNO5                          (3-10)

 1
 2     Here, as with SO2, the mechanism and rate of the gas-phase reaction is well established from
 3     laboratory studies (see NAPAP SOST, 1990), and the principal source of uncertainty in
 4     describing the reaction rate is the concentrations of the reagent species, mainly OH.  As
 5     noted above substantial progress has been made in the past few years  in measurement of
 6     OH.  It may thus be expected that improved knowledge of the concentration of this species
 7     will allow more confident evaluation of the rate of this reaction in specific situations and
 8     ultimately in regional-scale models.
 9          A second  key pathway for formation of nitric acid is the reaction sequence:
10
                                                        + O2                           (3-11)

11
                                                    +*  N205                           (3-12)

12
                                     N2O5+H2O(1) -* 2HNO3(aq)                        (3-13)

13
14
15     In addition, in daytime, photolysis of NO3 must be  considered:
16

                                          NOJ^NO + 02                             (3-14)

17
18          Other reactions of NO3 and/or N2O5, for example N2O5 with aromatics (Pitts et al.,
19     1985) must also be considered.  Reaction of N2O5 with liquid water appears to be rapid and
20     irreversible.  Studies of the uptake of N2O5 on aqueous sulfuric droplets give mass
21     accommodation coefficients of about 0.1 (Mozurkewich and Calvert, 1988; Van Doren et al.,
22     1990; Fried et al., 1994) .  Thus the overall rate and yield of this reaction can be evaluated

       April 1995                               3-50       DRAFT-DO NOT QUOTE OR CITE

-------
  1      from the pertinent gas-phase rate constants and the mass transfer rate constant for uptake of
  2      N2O5 by aqueous aerosol or cloud droplets.
  3           Recently a study has been reported which claims to provide indication of uptake of
  4      nitrogen oxides to nitrate during passage of an airstream through a hill cap cloud (Colvile et
  5      al. 1994).  However, as noted above, claims of this sort must be viewed with caution.  In
  6      this particular study it was necessary to invoke corrections for entrainment and dry deposition
  7      of magnitude comparable to the measured differences.
  8
  9      3.3.5  Water Content  and  Aerosol Equilibria
 10      3.3.5.1   Water Content of Atmospheric Aerosols, and Its Dependence on Ambient
 11               Humidity
 12           Water is an important ingredient of atmospheric aerosols (AA).  The water content of
 13      AA and the behavior of A A with respect to changes in ambient humidity are  of great
 14      importance in the global water cycle, the global energy budget, and also in atmospheric
 15      chemistry and optics. Understanding the relationship between AA and water  has proven to
 16      be a difficult problem. Most of the water associated with AA is "unbound" (Pilinis et al.,
 17      1989) i.e., it can increase or decrease with ambient humidity in a non-linear manner.  This
 18      non-linear relationship depends on  particle size and composition, indeed on size-dependent
 19      composition.  More recent  studies  have included monitoring of particle size distributions
 20      (either directly, or indirectly through light scattering and use of Mie theory) and size-
 21      dependent chemical composition under controlled RH (e.g., Covert and Heintzenberg,  1984;
 22      Rood et al., 1985).  Such studies have presented increasing evidence in favor  of external
 23      mixtures in particles.  Covert and Heintzenberg  (1984) found that size spectra of sulfur-
 24      bearing species were sensitive to RH while those of EC were not, and concluded that sulfur
 25      and  EC are, to some extent, externally mixed.  Harrison (1985) segregated the particles into
26      CCN (cloud condensation nuclei) and non-CCN  fractions and measured their chemical
27      compositions. Both fractions contained sulfate, nitrate and soot,  but sulfate was 15% of the
28      CCN mass and only 5.8% of the non-CCN mass.  Again, this was taken as evidence of
29      external mixture to some extent.  The differential mobility analyzer (DMA) has been a useful
30      tool  permitting study of particle properties for monodispersed size classes. Using the DMA,
31      Covert et al.  (1990) and Hering and McMurry (1991) showed that monodispersed particles

        April 1995                               3.51       DRAFT-DO NOT QUOTE OR CITE

-------
 1      scatter varying amounts of light in a single particle optical counter, indicating different
 2      refractive indices, and hence, different chemical composition.  Using a Tandem
 3      DMA (TDMA), McMurry and Stolzenberg (1989) showed that hygroscopic and hydrophobic
 4      particles of the same size co-exist frequently in Los Angeles, again an indication of external
 5      mixing.
 6
 7      3.3.5.2   Equilibria with water vapor
 8           The principal equilibrium of concern pertinent to ambient aerosols is that with water
 9      vapor.  This equilibrium is important as it  influences the size of the particles and in turn
10      their aerodynamic properties (important for deposition to the surface, to airways, following
11      inhalation, and to sampling instrumentation) and their light scattering properties.  This
12      section reviews recent work describing this equilibrium as it pertains to ambient aerosols.
13           Briefly the interaction of particles with water vapor may be described as follows. As
14      relative humidity increases, crystalline  soluble salts in aerosol  particles undergo a phase
15      transition to become aqueous solution aerosols.  According to  the phase rule, for particles
16      consisting of a single component, this phase transition is abrupt, taking place at a relative
17      humidity that corresponds to the vapor pressure  of water above the saturated solution (the
18      deliquescence point).  With further increase in relative humidity the particle growth is such
19      that the vapor pressure of the solution  (concentration  of which decreases as additional water
20      is accreted) is maintained equal to that of the surrounding relative  humidity; the particle thus
21      tends to  follow the equilibrium growth curve.  As relative humidity decreases,  the particle
22      follows the equilibrium curve to the deliquescence point. However, rather than crystallizing
23      at the deliquescence relative humidity,  the particle remains a solution (supersaturated
24      solution) to considerably  lower relative humidities.  Ultimately the particle abruptly loses its
25      water vapor (efflorescence), returning typically to the initial, stable crystalline form.  This
26      behavior has been amply demonstrated in numerous laboratory studies (Tang and
27      Munkelwitz,  1977; Tang, 1980).  Recently Tang and Munkelwitz  (1994) have presented data
28      for water activity (equilibrium relative humidity) as a function of composition for several
29      sulfate salts.
30           For particles consisting of more than one component, the solid to liquid transition will
31      take place over a  range of relative humidities, with an abrupt  onset at the lowest

        April 1995                                3-52      DRAFT-DO NOT QUOTE OR CITE

-------
 1      deliquescence point of the several components, and with subsequent growth as crystalline
 2      material in the particle dissolves according to the phase diagram for the particular
 3      multicomponent system.   Under such circumstances a single particle may undergo several
 4      more or less abrupt phase transitions until the soluble material is fully dissolved.  At
 5      decreasing relative humidity such particles tend to remain in solution  to relative humidities
 6      well below the several deliquescence points;  such behavior has been amply demonstrated.  In
 7      the case of the sulfuric acid-ammonium sulfate-water system the phase diagram is fairly
 8      completely worked out, but in the case of the mixed anion system with nitrate, there are
 9      remaining uncertainties (Tang et al, 1978,  1981;  Spann and Richardson, 1985).  Spann and
10      Richardson also give the compositional dependence of the relative humidity of efflorescence.
11      For particles of composition intermediate between NH4HSO4 and (NH4)2SO4 this transition
12      occurs in the range from  40% to below 10%, indicating that for certain compositions the
13      solution cannot be dried in the atmosphere.  Particles of this  composition would likely be
14      present at low relative humidities in the atmosphere as supersaturated salts and exhibit
15      apparent hygroscopic rather than deliquescent behavior.
16           Evidence of the interaction of ambient aerosol particles with water vapor has been
17      obtained by several investigators.  Koutrakis et al. (1989) found systematically  increasing
18      aerosol mean diameter with increasing relative humidity, which they attributed  to water
19      accretion on sulfates.  Rood et al. (1989) examined the response of light scattering coefficient
20      of ambient aerosols to increase in temperature  (effectively, reduced relative humidity) and
21      established that metastable supersaturated aerosols were essentially "ubiquitous".  More
22      detailed information regarding the size dependence of hygroscopic properties has been
23      obtained examining the change in particle size  of a monodisperse size cut selected with a
24      mobility analyzer, subjecting that aerosol to an increase or decrease in relative  humidity,  and
25      reanalyzing the size at the new humidity.  Studies of this phenomenon in the Los Angeles
26      area indicate this phenomenon, but also frequently indicate the presence of externally mixed
27      aerosol, in which some of the aerosol exhibits  the growth expected of soluble salts, where
28      another, apparently hydrophobic, fraction does not exhibit such growth (McMurry and
29      Stolzenburg, 1989).  Such bimodal RH growth is exhibited for particles present at Hopi Point
30      Arizona (Pitchford and McMurry  (1994).  In the latter study the relative humidity
31      dependence of the size of the  more hygroscopic fraction was found to be consistent with that

        April 1995                                 3.53      DRAFT-DO  NOT QUOTE OR CITE

-------
 1      expected for sulfate salts.  Such external mixtures have also been commonly observed in
 2      European aerosols (Hansson and Svenningsson, 1994).
 3           The time constant that characterizes the rate of exchange of water vapor between the
 4      gas phase and a solution droplet is of interest relative to the rate of response of particles to
 5      changes in relative humidity in the ambient environment, especially in the vicinity of
 6      surfaces, and relative to changes experienced by particles following inhalation or during
 7      sampling.  It is generally assumed that the rate of this water exchange is rapid.  The
 8      characteristic time for diffusional growth in response to a change in relative humidity was
 9      calculated by Pilinis et al. (1989) to be about 1  x 10'7  s.  However Klystov et al. (1993)
10      noted that this estimate was erroneously low by several orders of magnitude.  The latter
11      investigators examined the characteristic time for establishment of phase equilibrium in
12      response to a change in relative humidity for (NH4)2SO4 aerosol particles  (dry radius 0.5
13      /xm). The characteristic time increases  from ca 1 ms at 8% RH to 1.6 s at 99% RH.  Above
14      99%  RH the characteristic time can become much longer because  of the large change in
15      droplet radius at  such relative humidities.  These calculations  indicate that the water
16      equilibrium can be expected to be rapidly achieved in the ambient environment. A possible
17      but important exception is near 100% RH, pertinent to dry deposition  of particles to
18      vegetation or to water, where the equilibrium size might not be reached in the time required
19      for the  particle to traverse the diffusive layer adjacent  to the surface.
20           The lability of water associated with ambient aerosol has been evidenced in
21      comparisons by Malm et  al.  (1994) of measured paniculate light scattering coefficient
22      obtained with an integrating nephelometer with values  reconstructed from aerosol
23      composition,  taking into account the relative humidity  dependence of light scattering
24      coefficients of the aerosol components.  The reconstructed values were found to
25      systematically exceed the measured value.  However when in  the reconstruction the relative
26      humidity was taken as that in the nephelometer chamber (invariably lower than ambient
27      because of heating in the chamber) the reconstruction was markedly improved.
28
29      3.3.5.3   Ammonium Nitrate Vaporization Equilibria
30           In the sulfate system the vapor pressure of H2SO4 is negligible,  so that all sulfate may
31      be considered present  in the particles.  Also, at least for acidic sulfates (that is, not fully

        April 1995                                3.54      DRAFT-DO NOT QUOTE OR CITE

-------
 1     neutralized) the vapor pressure of NH3 is likewise negligible.  Even for fully neutral
 2     (NH4)2SO4 any hydrolysis of NH4+ to form NH3 that might escape to the vapor phase is
 3     suppressed by the resultant acidity.  In contrast, nitrates in aerosols are distinguished from
 4     sulfates because of the volatility of NO3- (as HNO3) and of NH4NO3 (as NH3 -I- HNO3).
 5     The equilibrium
 6
                               NH^N05 (s) or (aq)  ?* NH5(g) + HNO/g)

 7
 8     is such that at ambient conditions the partial pressures of NH3 and/or HNO3 are appreciable
 9     above crystalline NH4NO3 and likewise above solutions containing NH4 +  and NO3- ions (of
10     not necessarily equal concentrations).  It is thus necessary to consider these equilibria not just
11     for the crystalline material but also for  solutions, in the latter case as a function of
12     concentration or, equivalently, water activity.  Such a treatment has been given in detail by
13     Stelson and Seinfeld (1982a,b), and that study  is the basis for much subsequent interpretation
14     of field measurements.
15           As an example of such a study, Harrison and Msibi (1994) show comparison of
16     measured concentration product of HNO3 and NH3 versus the equilibrium constant for the
17     reaction.  Agreement is found roughly within a factor of 2 or so based on assumption of
18     equilibrium with pure NH4NO3 (crystal or solution).   However  when the observations were
19     stratified by RH, no strong trend of measured concentration product with RH was evidenced.
20           As noted above, the time scale of reaching this  equilibrium is of interest, for example
21     as it may influence dry deposition or accommodation to changing gaseous environments, as
22     in human airways.  Wexler and Seinfeld (1990) modeled the time dependence  of achieving
23     this equilibrium and concluded that equilibrium is generally reached within seconds to
24     minutes for typical  aerosol loadings.  However they caution that at low temperatures and low
25     aerosol loadings the time constant for achieving this equilibrium can be a day or more.
26           An important  implication of the high vapor pressure of ammonium nitrate (as NH3 +
27     HNO3) is that ammonia will distill from any ammonium nitrate if there is an acidic site
28     present, for example acidic sulfate that is less than  fully neutralized by ammonia.  As a
29     consequence ammonium nitrate aerosol is not expected to coexist with acidic aerosol.  As

       April 1995                                3.55       DRAFT-DO NOT  QUOTE OR CITE

-------
 1     noted above this situation appears generally to obtain, for example in the work of Gebhart
 2     et al. (1994).
 3          A further consequence of this equilibrium is the influence it may exert on dry
 4     deposition.  Sievering et al.  1994 report high rates of deposition of paniculate nitrate 2 to 9
 5     cm s-1, comparable to that of HNO3, to forest canopies, inferred from steep gradients of
 6     NH4NO3 concentration with height above the canopy.  They attribute this to the  large
 7     particle size of the nitrate, 2 to 2.5 pim mean diameter,  citing calculation of Peters and Eiden
 8     (1992).  An alternative explanation of the observations,  which does not appear to be ruled
 9     out, is that the deposition is actually of HNO3; that deposition of HNO3 perturbs the
10     equilibrium of NH4NO3 with NH3 + HNO3, leading to decrease of NH4NO3 in the vicinity
11     of the  surface and apparent deposition of this species.
12
13     3.3.6  Carbon-containing Particulate Matter
14     3.3.6.1  Introduction
15          The carbonaceous fraction of ambient paniculate matter consists of both elemental (EC)
16     and organic carbon  (OC).  Elemental carbon, also called black carbon or graphitic carbon,
17     has a chemical structure similar to impure  graphite and is emitted directly into the
18     atmosphere predominantly during combustion.  Organic carbon is either emitted  directly by
19     sources (primary OC) or can be formed in situ by condensation of low volatility  products of
20     the photooxidation of hydrocarbons (secondary OC).  The primary carbonaceous aerosol
21     (sum of primary EC and OC) is traditionally called soot.  Small additional quantities of
22     aerosol carbon may exist either as carbonates (e.g., CaCO3) or CO2 adsorbed onto
23     paniculate  matter such as soot (Appel et al., 1989; Clarke and Karani, 1992).
24
25     3.3.6.2  Elemental Carbon
26          Elemental carbon is a strong absorber of visible radiation and is the major  species
27     responsible for light absorption by atmospheric particles  (Novakov, 1984; Goldberg, 1985;
28     Fmlayson-Pitts and  Pitts, 1986; Japar et al., 1986; Sloane et al., 1991; Hamilton and
29     Mansfield, 1991). Elemental carbon found in atmospheric particles is a complex three
30     dimensional array of carbon with small amounts of other elements such as oxygen, nitrogen,
31     and hydrogen incorporated in its graphitic  hexagonal structure (Chang et al.,  1982).

       April 1995                                3.56      DRAFT-DO NOT QUOTE OR CITE

-------
 1           Wood-burning fireplaces and diesels are major sources of EC (Mulhbaier and Williams,
 2      1982; Dasch and Cadle, 1989; Brown et al., 1989; Dod et al., 1989; Hansen and Rosen,
 3      1990; Burtscher, 1992). In areas where wood burning is significant, more particulate
 4      graphitic carbon is expected in winter than in summer.  Tracer techniques have been
 5      developed  for the  calculation of the source contribution to the EC concentrations, including
 6      use of K as a woodsmoke tracer (Currie et al.,  1994)  and use of the carbon isotopic tracers
 7      14C and 12C (Lewis et al.,  1988; Klouda et al., 1988; Currie et al., 1989).  Around 47% of
 8      the EC in Detroit, 93% in  Los Angeles and 30 to 60% in a rural area in Pennsylvania has
 9      been attributed to  motor vehicle sources (Wolff and Korsog, 1985; Pratsinis et al.,  1988;
10      Keeler et al., 1990),  The corresponding contribution  of diesel emissions to EC
11      concentrations in Western Europe is estimated to be 70 to 90% (Hamilton and  Mansfield,
12      1991). Elemental carbon was also a major constituent of the Kuwait oil fires,  with
13      concentrations as high as 178 mg m"3 inside the plume (Cofer et al.,  1992; Daum et al.,
14      1993; and  references therein).  Global emissions of EC were estimated by Penner et al.
15      (1993) to be 12.6  to 24 Tg C yr'1, while the EC emission for the US was 0.4 to  1.1 Tg yr1
16      and for the rest of North America 0.2 Tg yr"1.
17           Elemental carbon also scatters light (Appel et al., 1985) although its light scattering
18      efficiency is smaller than the efficiencies of the other  aerosol principal components  (Sloane
19      et al., 1991). Because EC both absorbs and scatters light, its contribution to total light
20      extinction exceeds its contribution to fine particle mass.  For example,  in Los Angeles, EC
21      was found to represent 8.5 to 10% of the fine particulate mass, but to account  for 14 to 21%
22      of  the total light extinction (Pratsinis et al., 1984). A significant fraction of the dark colored
23      fine EC particles is able to penetrate the indoor atmosphere of buildings and may constitute a
24      soiling hazard of objects like works of art (Ligocki et al., 1993).
25           The concentration of EC varies significantly depending on location and season.
26      Elemental carbon concentrations in rural and remote areas usually vary from 0.2 to 2.0 mg
27      m'3 (Wolff, 1981;  Clarke et al., 1984;  Goldberg, 1985; Cadle  and  Dash, 1988; Japar et al.,
28      1986; Shah et al., 1986; Pinnick  et al.,  1993) and from  1.5 to 20  mg m'3 in urban areas
29      (Wolff, 1981; Delumyea and Kalivretenos, 1987; Pratsinis et al., 1984,  1986; Grosjean,
30      1984; Heitzenberg and Winkler, 1984; Goldberg, 1985; Shah et  al., 1986; Rau, 1989). The
31      concentration of EC over the remote oceans is approximately 5 to 20 ng m"3 (Clarke, 1989).

        April  1995                               3.57      DRAFT-DO NOT QUOTE OR CITE

-------
 1     Average EC concentration values are around 1.3 and 3.8 mg m"3 for U.S. rural and urban
 2     sites respectively (Shah et al., 1986).  Average PM10EC values exceeding 10 mg m~3 are
 3     common for some urban locations (Chow et al., 1994). The ratio of EC to total carbon has
 4     been observed to vary from 0.15 to 0.20 in rural areas, to 0.2 to 0.6 in urban areas (Wolff et
 5     al., 1982; Gray et al.,  1984;  Grosjean, 1984;  Pratsinis et al., 1984; Chow et al.,  1993).  The
 6     annual mean of this ratio was approximately 0.4 for the Los Angeles basin in 1982 (Gray et
 7     al., 1986), while this ratio in the same area decreases to 0.2 during summer midday periods
 8     (Larson et al., 1989; Wolff et al., 1991).  Aging of an air mass results in lowering of the EC
 9     fraction of the aerosol due to its mixing  with non-combustion particles or by condensation of
10     material  from the gas phase (Burtscher et al.,  1993).
11          The distribution of EC emitted by automobiles is unimodal with over 85 % of the mass
12     in particles smaller that 0.12  mm aerodynamic diameter (Venkataraman et al. 1994).  The
13     ambient distribution of EC is bimodal with peaks in the 0.05 to 0.12 mm (mode I) and 0.5 to
14     1.0 mm (mode II) size ranges (Nunes and  Pio, 1993; Venkataraman and Friedlander, 1994).
15     The creation of mode II is mainly the result of accumulation of secondary aerosol products
16     on primary aerosol particles.
17          The degree of mixing of EC particles with the rest of the aerosol components remains a
18     controversial issue. Particles emitted from spark-ignition engines have been found to consist
19     of a core of EC  covered with a layer of PAHs and an  outermost shell of volatile compounds
20     (Steiner et al., 1992).  Ambient carbonaceous aerosol in urban areas has been found to
21     consist of aggregated spherules, with a range of carbon structures from  amorphous (OC) to
22     graphitic (EC) within aggregates (Katrinak et al., 1992). These aggregates are often
23     (especially during summer months) coated with sulfates and nitrates (Katrinak et al.,  1992,
24     1993).  However, often sulfate and  EC are externally mixed  (Covert and Heintzenberg,
25     1984).  Coating  of EC with organic compounds may alter its hygroscopicity and its lifetime
26     in the atmosphere (Andrews and Larson, 1993).  Noone et al. (1992) reported that the
27     interstitial aerosol inside urban fogs is enriched  in EC, something that would tend to  increase
28     its lifetime in the atmosphere with respect to other  species like sulfate or OC (Nunes and
29     Pio, 1993). However, the degree of incorporation of EC in droplets is highly variable (0 to
30     80%) and its behavior appears to vary from hygroscopic to hydrophobic (Hansen  and
        April 1995                               3-58       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Novakov, 1988).  Our luck of understanding of these processes makes a quantitative estimate
 2      of the atmospheric lifetime of EC problematic.
 3           The participation of EC and soot in atmospheric chemical reactions with SO2,  O3 and
 4      NO2 has been the subject of a series of studies (Baldwin, 1982; Dlugi and Giinsten, 1983;
 5      Akhter et al., 1985; 1985; Jassim et al., 1989; Sergides et al., 1987; Gundel et al., 1989;
 6      Chughtai et al., 1991).  The strong dependence of the often conflicting results of these
 7      studies on the nature of the samples is inhibiting the extrapolation of their  conclusions to  the
 8      atmosphere.  Chughtai et al. (1991) reported that oxidation and hydrolysis of accessible
 9      reactive sites on the soot surface result in particle solubilization and accelerated particle
10      removal from the atmosphere.  DeSantis and Allegrini (1992) suggested that NO2 reactions in
11      the presence of SO2 on carbon-containing particles could be a source of HNO2 in the urban
12      environment.  The reaction of soot with ozone is faster than its reaction with NO2 that is  in
13      turn faster than the reaction with SO2 (Smith et al., 1989).
14
15      3.3.6.3 Organic  Carbon
16           The organic component of ambient aerosol both in polluted and remote areas is a
17      complex mixture of hundreds of organic compounds (Cass et al.,  1982; Seinfeld,  1986;
18      Rogge, 1993; Hahn, 1980; Simoneit and Mazurek,  1982; Zafiriou et al.,  1985; Graedel,
19      1986).  Compounds identified in the ambient aerosol include n-alkanes, n-alkanoic acids,  n-
20      alkanals, aliphatic dicarboxylic acids, diterpenoid acids and retene, aromatic polycarboxylic
21      acids, polycyclic aromatic hydrocarbons, polycyclic aromatic ketones and quinones, steroids,
22      N-containing compounds, regular steranes, pentacyclic triterpanes, iso- and anteiso-alkanes,
23      etc.   (Graedel, 1986; Mazurek et al., 1989; Hildemann et al.,  1993; Rogge, 1993).  OC does
24      not strongly absorb light, but its light scattering efficiency in urban hazes is similar to that of
25      nitrate and sulfate.
26           Aerosol OC measurements are often subject to sampling artifacts due to adsorption of
27      organic vapors on the filters used or evaporation of the collected mass. These sampling
28      problems are discussed  in Section 3.3.3.1. Wolff et al. (1991) found that  this sampling error
29      represented roughly  20%  of the measured OC under urban polluted conditions.  McMurry
30      and  Zhang (1989) observed in ambient and smog chamber measurements that a consistently
31      large fraction of the OC (40 to 70%) was collected on the quartz filters following their

        April 1995                                3.59      DRAFT-DO NOT QUOTE OR CITE

-------
 1     impactors.  The strong possibility of sampling artifacts in the laboratory and field
 2     measurements presented below, increases the uncertainty of our current knowledge about
 3     aerosol OC.  Most of the investigators report the OC concentration as concentration of
 4     carbon.  These values neglect the contribution to the aerosol mass of the  other elements
 5     (namely oxygen, hydrogen and nitrogen) of the organic aerosol compounds. Wolff et al.
 6     (1991) suggested that the measured OC values should be multiplied by a  factor of 1.5 for the
 7     calculation of the total organic mass associated with the OC.
 8          The concentration of OC is around 3.5 mg C m~3 in rural locations  (Stevens et al.,
 9     1984) and 5 to 20 mg C m"3 in polluted atmospheres (Grosjean, 1984a; Wolff et al.,  1991).
10     Wolff et al. (1991) and Chow et al. (1994) summarizing measurements during the summer
11     and fall of 1987 in the Los Angeles basin, reported that OC represented on average 10 to
12     18% of the PM10 mass and 11 to 24% of the PM2 5 mass during the summer and 15  to 25%
13     of the PM10 and 16 to 25% of the PM2 5 during the fall.  Wolff et al. (1991) suggested that
14     these values should be reduced by roughly 20% to correct for the sampling bias and then
15     multiplied by 1.5 to account for the non-carbon mass of the organic aerosol compounds (an
16     overall increase by roughly a factor of 1.3).  In rural areas of the western U.S. paniculate
17     OC concentrations are comparable to  sulfate (White and Macias, 1989).  In other areas OC
18     contributes roughly  10 to 15% of the  PM2 5 and PM10 mass (Stevens et al., 1984).  Organic
19     compounds accumulate mainly in the  submicrometer aerosol size range (Finlayson-Pitts and
20     Pitts, 1986; McMurry and Zhang, 1989) and their mass distribution is typically bimodal with
21     the first peak around diameter of 0.2  jum and the second around 1 jiim (Pickle et al.,  1990;
22     Mylonas et al., 1991).
23          The contribution of the primary  and secondary components of aerosol OC have been
24     difficult to quantify.  The lack of a direct chemical analysis method for the identification of
25     either of these OC components has led researchers to the  employment of several indirect
26     methods.  These methods include the  use of tracer compounds for either  the primary or the
27     secondary OC (Larson et al., 1989; Turpin and Huntzicker, 1991; Turpin et al., 1991), the
28     use of models describing the emission and dispersion of primary OC (Gray, 1986;  Gray et
29     al., 1986; Larson et al., 1989; Hildemann, 1990) and  the use of models describing the
30     formation of secondary OC (Pilinis and Seinfeld,  1988; Pandis et al., 1992; Pandis et al.,
31     1993).  The above studies concluded  that the secondary OC contribution  is maximized in the

       April 1995                               3-60      DRAFT-DO NOT QUOTE OR CITE

-------
  1      early afternoon of summer days, varying from 30 to 60% of the total OC depending on
  2      location.  The yearly averaged contribution of secondary OC is smaller,  accounting for 10 to
  3      40% of the OC.
  4           The interactions of the OC compounds with each other and the inorganic aerosol
  5      species are poorly understood. The compounds have the potential to form organic films
  6      around the inorganic and EC core of the aerosol (Gill et al., 1983).  Goschnich et al. (1990)
  7      provided evidence for such formation by reporting  that carbon compounds and organic
  8      hydrogen were enriched within the particles' outer layer, while inorganics like NH4NO3 were
  9      enriched inside the particles.   The presence of such films can inhibit the  transport of water
10      and other inorganic components between the gas and aerosol phases  (Otani and Wang, 1984;
11      Rubel and Gentry,  1984).
12
13      3.3.6.4 Primary Organic Carbon
14           Primary carbonaceous particles (OC) are produced by combustion (pyrogenic), chemical
15      (commercial products),  geological (fossil fuels), and natural (biogenic) sources.  The
16      complexity of the mixture molecular composition of OC is such that tracer compounds are
17      still necessary to decouple the contributions of the  various  sources.  Rogge et al.  (1991)
18      suggested that fine aerosol  cholesterol could be used as a tracer for meat smoke.  An
19      alternative proposed meat smoke tracer set consists of myristic acid (n-tetradecanoic acid),
20      palmitic acid (n-hexadecanoic  acid), stearic acid (n-octadecanoic acid), oleic acid  (cis-9-
21      octadecenoic acid), nonanal and 2-decanone (Rogge et al.,  1991).  Benzothiazole  has been
22      used as a tracer for tire wear contributions to ambient aerosol (Kim et al., 1990;  Rogge
23      et al., 1993b).  Steranes and pentacyclic triterpanes (hopanes) can be used as tracer
24      compounds for the vehicular sources  (Rogge et al., 1993a). The  odd carbon number n-
25      alkanes ranging from C27 to C33 can serve as a molecular tracer assemblage for biogenic
26      primary OC (green,  dead, and degraded plant wax material directly emitted or resuspended
27      from soil and road dust) (Mazurek and Simoneit, 1984; Simoneit,  1989;  Rogge et al.,
28      1993c).  The iso- and anteiso- alkanes can be used to trace the cigarette  smoke contribution
29      to the outdoor atmosphere (Rogge et al., 1994),
30           Primary biogenic organic matter consists predominantly of lipids, humic and fulvic
31      acids, and often represents  a major fraction of the carbonaceous aerosol mass (Duce et al.,

        April 1995                               3.51       DRAFT-DO NOT QUOTE OR CITE

-------
  1      1983; Gagosian et al., 1987; Mazurek et al., 1989, 1991; Simoneit, 1984, 1986, 1989).
  2      Mamane et al. (1990) reported that most coarse OC in the Great Lakes region is of biologic
  3      origin while most fine OC is anthropogenic.
  4
  5           Secondary Organic Carbon
  6           Secondary organic  aerosol material is formed in the atmosphere by the condensation on
  7      already existing particles of low vapor pressure products of the oxidation of organic gases.
  8      As the hydrocarbons are oxidized in the gas-phase by species such as the hydroxyl radical
  9      (OH), ozone (O3) and the nitrate radical (N03), their oxidation products accumulate in the
10      gas-phase.  If the concentration of such a product is smaller than its saturation concentration,
11      the species remains mainly in the gas-phase.  Small amounts of the species can be adsorbed
12      on aerosol surfaces or dissolved in the aerosol phase at this stage (Yamasaki et al., 1982;
13      Pankow, 1987; Ligocki and Pankow, 1989; Pankow and Bidleman, 1991; Pankow, 1994a, b;
14      Pandis et al., 1992).  If the gas-phase concentration of a species exceeds its saturation
15      concentration, the species condenses on the available aerosol surface so that at equilibrium its
16      gas-phase concentration equals  its saturation concentration. If this gas-phase concentration is
17      reduced  to less than the saturation value as a result of dilution, deposition or chemical
18      reaction, the aerosol species evaporates in an effort to maintain thermodynamic equilibrium
19      (Pilinis and Seinfeld, 1988).  Many volatile organic compounds (VOC) do not form aerosol
20      under atmospheric conditions due to the high vapor pressure  of their products (Grosjean and
21      Seinfeld, 1989).  These VOC include all alkanes with up to six carbon atoms (from methane
22      to hexane isomers), all alkenes  with up to six carbon atoms (from ethylene to hexene
23      isomers), benzene and many low-molecular-weight carbonyls, chlorinated compounds and
24      oxygenated solvents.
25           Organic aerosols formed by gas-phase photochemical reactions of hydrocarbons, ozone
26      and nitrogen oxides have been identified in both urban and rural atmospheres (Grosjean,
27      1977). Most of these species are di- or poly-functionally substituted alkane  derivatives.
28      These compounds include aliphatic organic nitrates (Grosjean and Friedlander, 1975),
29      dicarboxylic  acids (adipic and glutaric acids) (O'Brien et al., 1975), carboxylic acids derived
30      from aromatic hydrocarbons (benzoic and phenylacetic acids), polysubstituted phenols and
31      nitroaromatics from aromatic hydrocarbons (Kawamura et al.,  1985; Satsumakayashi et al.,

        April 1995                                3_62        DRAFT-DO NOT QUOTE OR CITE

-------
  1      1989, 1990).  Some species that have been identified in ambient aerosol and are believed to
  2      be secondary in nature are depicted in Table 3-5. Despite the above studies, the available
  3      information about the molecular composition of atmospheric secondary  OC and about the
  4      composition of the OC produced during the oxidation of specific hydrocarbons remains
  5      incomplete.  The reaction mechanisms leading to the observed products are  to a great extent
  6      speculative at present (Finlayson-Pitts and Pitts,  1986). Natural hydrocarbons like the
  7      monoterpenes (C10H16) and isoprene (C5H8) are  emitted by various types of trees and plants.
  8      In the United States  the biogenic hydrocarbon  sources are estimated to produce 30 to 60 Mt
  9      of carbon per year (isoprene and monoterpenes combined) whereas anthropogenic
10      hydrocarbon sources have been estimated to account for 27 Mt of carbon per year (Lamb et
11      al., 1987;  Zimmerman, 1979; Altshuller, 1983). Laboratory investigations  have indicated
12      that biogenic hydrocarbons are very reactive under typical atmospheric  conditions (Arnts and
13      Gay, 1979).  The aerosol forming potential of biogenic hydrocarbons has been investigated in
14      a series of smog chamber studies (Kamens et al., 1981, 1982; Hatakeyama et  al., 1989;
15      1991; Pandis et al.,  1991; Zhang et al., 1992) and it has been suggested that isoprene
16      photooxidation does  not contribute to the production of secondary  aerosol under ambient
17      conditions.  On the contrary, pinenes and other monoterpenes form secondary aerosol in their
18      reactions with O3 and OH and have the potential to contribute significantly to  aerosol in
19      areas with high  vegetation coverage. Monoterpenes were estimated to contribute around
20      15% of the secondary organic aerosol (SOA) in urban areas with low vegetation like Los
21      Angeles, while they  are expected to dominate the SOA in areas with high vegetation
22      coverage like Atlanta (Pandis et  al., 1991, 1992).  The chemical composition of the majority
23      of the aerosol products of the monoterpene photooxidation remains unknown or is speculative
24      (Paulson et al.,  1990; Palen et al., 1992).   The  few products that have been identified
25      include nopinone, pinanediol, pinonic acid and 5-(l-hydroxy-l-methylethyl)-2-methyl-2-
26      cyclohexen-1-one. Several investigators have studied the SOA formation from selected
27      anthropogenic hydrocarbons. The literature data up to  1976 have  been  reviewed by Grosjean
28      (1977). Other studies focused on toluene and other aromatic hydrocarbons (Leone et al.,
29      1985; Stern et al., 1987; Gery et al., 1985, 1987; Izumi and Fukuyama, 1990),  styrenes
30      (Izumi and Fukuyama, 1990), cyclic olefins (Hatakeyma et al., 1985, 1987; Izumi et al.,
31      1988), cresols and nitrocresols (Grosjean, 1985)  and alkenes with  more than six carbon

        April 1995                               3_63      DRAFT-DO NOT QUOTE OR CITE

-------
          TABLE 3-5.  SOME SECONDARY ORGANIC COMPOUNDS IDENTIFIED IN
       	AMBIENT PARTICLES IN URBAN AIR	
       Compound                                                            n
       HOOC(CH2)nCOOH                                                   1-8
       HOOC(CH2)nCHO                                                    3-5
       HOOC(CH2)nCH2OH                                                  3-5
       HOOC(CH2)nCH2ONO or CHO(CH2)nCH2ONO2                           3-5
       CHO(CH2)nCHO                                                      3-5
       CHO(CH2)nCHO                                                      3-5
       HOOC(CH2)nCOONO or HOOC(CH2)nCOONO2                           3-5
       CHO(CH2)nCOONO                                                   3-4
       HOOC(CH2)nCOONO                                                  3-4
       HOOC(CH2)nCOONO2                                                 4-5
       HOOC(CH2)nCH2ON02                                                3-4
       (C6H6)-(CH2)nCOOH                                                  1-3
       HOOC-(C6H6)-(CH2)n                                                  1-3
 1     atoms (Grosjean,  1984b; McMurry and Grosjean, 1985; Wang et al., 1992).  Measured and
 2     estimated aerosol  yields from a variety of SOA precursors have been tabulated by Grosjean
 3     and Seinfeld (1989) and Pandis et al. (1992).
 4         The calculated contribution of the main individual secondary organic aerosol precursors
 5     to the secondary organic aerosol concentration in Los Angeles on August 28,  1987 is
 6     presented in Table 3-6 (Grosjean and Seinfeld, 1989; Pandis et al., 1992).  Toluene, the
 7     nonmethane hydrocarbon with the highest emission rate in the Los Angeles area (165 t d"1 )
 8     was predicted to contribute 28% of the secondary organic aerosols.  Differences were
 9     attributed to sampling artifacts and calibration uncertainties during the interpretation of the
10     ambient data.
11         Grosjean (1992) calculated the daily production rates of various chemical functionalities
12     of the secondary organic aerosol formed in situ during a smog episode in Los Angeles using
13     the precursor hydrocarbon emission inventory and the results of smog chamber studies. His
       April 1995                             3.54      DRAFT-DO NOT QUOTE OR CITE

-------
    TABLE 3-6. PREDICTED PER CENT CONTRIBUTION TO SECONDARY
        ORGANIC AEROSOL CONCENTRATIONS AT LOS ANGELES

Species
Contribution

Grosjean and Seinfeld (1989) Pandis et al. (1992)



1
2
3
4
Aromatics
Biogenic Hydrocarbons
Alkanes
Olefins
58 65
10 16
21 15
11 4




estimates are presented in Table 3-7. These predictions were compared with the available
measurements of ambient


OC functional group relative abundances (Grosjean,


TABLE 3-7. AMOUNT OF SECONDARY AEROSOL PRODUCED IN
LOS ANGELES SMOG EPISODE ACCORDING TO FUNCTIONAL
(GROSJEAN, 1992)





1
2
3
Precursor
Alkenes
Cyclic olefins
Terpenes
Alkanes
Cycloalkanes
Aromatics
TOTAL
Pickle et al. (1990)
Aerosol produced (kg day"1)
1992).


A TYPICAL
GROUPS

Carbonyls Aliphatic Acids Nitrophenols Aliphatic Nitrate
608
62 131
295 623
243
72
3118
672 1362 3118
and Mylonas et al. (1991) argued that the SO A mass
-
9
41
121
72
-
243
size
distribution in urban areas like Los Angeles is typically bimodal with maxima in the 0.1 and
1.0 /j,m size ranges. Our
understanding of the mechanisms of creation of these
two modes
April 1995
3-65
DRAFT-DO NOT QUOTE OR CITE

-------
 1     remains tentative (Pandis et al., 1993).  The effect of relative humidity in the SO A
 2     partitioning between gas and aerosol phases is generally not understood.  Thibodeaux et al.
 3     (1991) developed a theoretical model based on classical adsorption theory and predicted that
 4     as air relative humidity increases (remaining less than 60%) the equilibrium secondary
 5     organic carbon content on the aerosol particles decreases due to competition for adsorption
 6     sites with water molecules.  This theoretical result seems to be supported by the
 7     littleavailable experimental information, but the necessary experimental data for the
 8     incorporation of these relative humidity effects on SOA partitioning in aerosol model do not
 9     exist.  Knowledge of the saturation concentrations of the organic condensable species remains
10     incomplete.  These concentrations  are expected  to vary  significantly with temperature.  The
11     few available relevant measurements include the saturation vapor concentrations of
12     monocarboxylic and dicarboxylic acids (Tao and McMurry, 1989) and the b-pinene aerosol
13     products (Pandis et al., 1991).  The saturation vapor concentrations of condensable products
14     from the oxidation of some aromatic hydrocarbons (toluene, m-xylene, and 1,3,5-
15     trimethylbenzene)  were estimated to lie in the range 3 to 30 ppt (Seinfeld et al., 1987).
16     McMurry and Grosjean (1985) estimated saturation vapor concentrations for condensable
17     products from the oxidation of 1-heptene (0.14 to 1.28  ppb), o-cresol (0.06 to 1.6 ppb) and
18     nitrocresol (1.7 to 2.2 ppb).
19
20          Polycyclic Aromatic Hydrocarbons (PAH)
21          Polycyclic aromatic hydrocarbons are formed during the incomplete combustion of
22     organic matter, for example,  coal,  oil, wood and gasoline fuel (National Academy of
23     Sciences, 1983; Bjorseth,  1983).  Stationary sources  (residential heating, industrial processes,
24     open burning, power generation) are estimated to  account for roughly 80% of the  annual
25     total PAH emissions in the US with the remainder produced by mobile sources (Peters et al.,
26     1981; Ramdahl et al.,  1983).  Mobile sources are however the major contributors  in urban
27     areas (National Academy of Sciences,  1983; Freeman and Cattell, 1990). More than one
28     hundred PAH compounds have been identified in urban air.  The PAH observed in the
29     atmosphere range  from bicyclic species such as naphthalene, present mainly in  the gas phase,
30     to PAH containing seven or more  fused rings, such as coronene, which are present
31     exclusively in the  aerosol phase (Finlayson-Pitts and Pitts,  1986).  Intermediate PAH such as

       April 1995                                3-66      DRAFT-DO NOT QUOTE OR CITE

-------
  1      pyrene and athracene are distributed in both the gas and aerosol phases (see also
  2      Section 3.3.3.1).
  3           Measurements of the size distribution of PAH indicate that while they are found
  4      exclusively in the 0.01 to 0.5 jum diameter mode of fresh combustion emissions
  5      (Venkataraman et al.,  1994) they exhibit a bi-modal distribution in ambient urban aerosol,
  6      with an additional mode in the 0.5 to 1.0 /im diameter range (Venkataraman and Friedlander,
  7      1994). The growth of nuclei-mode particles by condensation of secondary aerosol species
  8      like nitrate, sulfate and secondary organic aerosol has been proposed as an explanation of this
  9      distribution.
 10           Polycyclic aromatic hydrocarbons adsorbed on the surfaces of combustion generated
 11      particles are released into an atmosphere containing gaseous co-pollutants including O3, NO2,
 12      SO2, HNO3, PAN, radicals and are exposed to sunlight.  Under these conditions PAH
 13      undergo chemical transformations that might lead to significant degradation and formation of
 14      products more polar than the parent PAH (National Academy of Sciences, 1983).  Several
 15      studies have focused on the reaction rates and products of reactions of PAH adsorbed on
 16      specific substrates and exposed in the dark or in the light to other pollutants.  However, the
 17      extrapolation of these laboratory results to real atmospheric conditions remains difficult.
 18           Benzo(a)pyrene (BaP) and other PAH on a variety of aerosol substrates react with
 19      gaseous NO2 and HNO3 to form mono- and dinitro-PAH (Finlayson-Pitts and Pitts, 1986).
20      The presence of HNO3 along with NO2 is necessary for PAH nitrification. The reaction rate
21      depends strongly on the nature of the  aerosol  substrate (Ramdahl et al., 1984), but the
22      qualitative composition of the products does not.  The aerosol water is also a favorable
23      medium for heterogeneous PAH nitration reactions (Nielsen et al., 1983).  Nielsen (1984)
24      proposed a reactivity classification of PAH based on chemical and spectroscopic parameters
25      (Table 3-8).  The PAH nitration rate under typical urban conditions remains poorly
26      understood. Bjorseth et al.  (1979) observed a lack of significant PAH reactions during their
27      transport from central  to northern Europe and suggested that these reactions are slow in most
28      environments.  However, this may not be the case in heavily polluted areas with high NO2
29      and HNO3 concentrations and acidic particles (Finlayson-Pitts and Pitts, 1986).  Reactions of
30      fluoranthene and pyrene with NO2 in the  gas phase and condensation the 2-nitro-PAH
       April 1995                                3_67      DRAFT-DO NOT QUOTE OR CITE

-------
         TABLE 3-8.  REACTIVITY SCALE FOR THE ELECTROPHILIC REACTIONS OF
        	PAH (REACTIVITY DECREASES IN THE ORDER I TO VI)	
         I    Benzo(a)tetracene, dibenzo(a,h)pyrene, pentacene, tetracene
         II   Anthanthrene, anthracene, benzo(a)pyrene, cyclopenta(cd)pyrene, dibenzo(a,l)pyrene,
             dibenzo(a,i)pyrene, dibenzo(a,c)tetracene, perylene
         III  Benz(a)anthracene, benzo(g)chrysene, benzo(ghi)perylene, dibenzo(a,e)pyrene,
             picene, pyrene
         IV  Benzo(c)chrysene, benzo(c)phenanthrene, benzo(e)pyrene, chrysene, coronene,
             dibenzanthracene, dibenzo(e, l)pyrene
         V   Acenaphthylene, benzofluoranthenes, fluranthene, indeno(l,2,3-cd)fluoranthene,
             indeno(l,2,3-cd)pyrene, naphtalene, phenanthrene, triphenylene
         VI  Biphenyl
 1     derivatives on the aerosol surface have been proposed as an alternative reaction pathway for
 2     the production of the observed aerosol nitro-PAH (Pitts et al., 1985a).
 3          Nitrogen oxide (N2O5) has been proposed as an additional nitrating agent for certain
 4     PAH during nighttime (Kamens et al., 1990).  Pitts et al. (1985b) exposed six PAH  to N2O5
 5     and proposed the following reactivity  order: pyrene > fluoranthene > BaP  >
 6     benz(a)anthracene >  perylene > chrysene.  Nitro-PAH photodecompose into quinones and
 7     possibly phenolic derivatives.  For example 6-NO2-BaP on silica gel photolyses to the 1,6-,
 8     3,6-, and 6,12- isomers of BaP quinones and a host of other oxy-PAH (Finlayson-Pitts and
 9     Pitts, 1986).  These reactions are expected to depend strongly on the chemical composition
10     and structure of the aerosol substrate and are not well understood for ambient particles.
11          Aerosol PAH react with O3 to produce oxidized PAH.  Pyrene, BaP and athracenes
12     react rapidly and the benzofluoranthenes slowly (Finlayson-Pitts and Pitts, 1986; Alebic-
13     Juretic  et al., 1990).  Reaction rates of 15 to 30% hr"1 were observed for the most reactive
14     PAH adsorbed on filters during exposure to 200 ppb of O3 (Pitts et al., 1986). However,
15     other researchers (Atkinson and Aschmann,  1986; Coutant  et al.,  1988; De Raat et al., 1990)
16     have suggested  that the PAH-O3 reaction is of negligible importance for typical atmospheric
17     conditions. Relatively little is known  about the full ranges  of products and the mechanisms
18     of their formation.  Polycyclic aromatic hydrocarbons exposed to sunlight have been found to

       April 1995                               3-68       DRAFT-DO NOT QUOTE OR CITE

-------
 1      photodegrade in a series of laboratory studies (Valerio and Lazzarotto,  1985; Behymer and
 2      Kites, 1988). The photodegradation rates depend strongly on the chemical composition and
 3      the pH of the aerosol substrate (Dlugi and Giisten, 1983; Valerio and Lazzarotto, 1985;
 4      Behymer and Hites,  1988). Polycyclic aromatic hydrocarbons appear to be more stable when
 5      adsorbed on ambient aerosol than when present in pure form or in solution or on artificial
 6      surfaces (Baek et al., 1991). The occurrence of PAH-SOX reactions remains uncertain
 7      (Baek etal., 1991).
 8
 9      3.3.7 Metals and Other Trace Elements
10           The major components of fine particles are sulfate,  nitrate, organic and elemental
11      carbon, ammonium ions and a variety of trace elements (Godish, 1985; Pitts and Pitts,
12      1986).  Trace elements that are found predominantly in the fine particle size distribution are
13      Na, Cs, Cl,  Br, Cu, Zn, As, Ag, Cd, In, Sn,  Sb, W, and Pb, and greater than 75% of their
14      mass is associated with particles of diameter less than 2 um.  Metals which are found in both
15      fine and coarse modes are  V, Cr, Mn, Fe, Co, and Se, while elements found primarily
16      within large  particle distributions are Ca, Al, Ti, Sc, and La (Klee, 1984).  The
17      concentrations and the relative proportions of these species in the various particle size ranges
18      depend on a number of factors such as the nature of the emissions, the  photochemical activity
19      and the meteorology (Pitts  and Pitts, 1986).  The concentration ranges of various elements
20      associated with particulate  matter in the atmosphere are shown in Table 3-6. For most
21      elements the range in concentrations is greater than three orders of magnitude.  This reflects
22      the different sources and the different pollution control strategies that exist in each area.
23      This information was compiled by Schroeder et al. (1987), and includes a large number of
24      studies from the United States, and abroad, which indicates the need to complete site specific
25      evaluations for high  end concentrations (references can be found in the  original paper by
26      Schroeder et al., 1987).
27           In general, remote areas recorded measurable concentrations of some elements
28      associated with crustal origin, as well as some elements indicative of anthropogenic sources.
29      This supports hypotheses which suggest that long range transport occurs in these remote
30      areas (Schroeder et al., 1987).  The urban data (Table 3-9) reflect elemental concentrations
31      in different parts of the world.  Elements such as lead, iron, and copper are measured in

        April 1995                                3_69      DRAFT-DO NOT QUOTE OR CITE

-------
       TABLE 3-9 CONCENTRATION RANGES OF VARIOUS ELEMENTS ASSOCIATED
                WITH PARTICIPATE MATTER IN THE ATMOSPHERE (ng/m3)
Elements
As
Cd
Ni
Pb
V
Zn
Co
Cr
Cu
Fe
Hg
Mn
Se
Sb
Remote
0.007
0.003
0.01 -
0.007
0.001
- 1.9
- 1.1
60.0
-64
- 14
0.03 - 460
0.001
0.005
0.029
0.62-
0.005
0.01 -
0.0056
0.0008
-0.9
- 11.2
- 12
4,160
- 1.3
16.7
-0.19
- 1.19
Rural
1.0
0.4-
0.6
2- 1
2.7
11 -
0.08-
1.1
3-
-28
1,000
-78
,700
-97
403
- 10.1
-44
280
55 - 14,530
0.05
3.7
0.01
0.6-
- 160
-99
-3.0
-7.0
Urban (USA)
2 - 2,320
0.2 - 7,000
1 -328
30 - 96,270
0.4- 1,460
15 - 8,328
0.2 - 83
2.2 - 124
3 - 5,140
130 - 13,800
0.58 -458
4-488
0.2 - 30
0.5 - 171
       Source: Schroeder et al., 1987
 1      greatest abundance in paniculate matter from all locations, while elements such as cobalt,
 2      mercury and antimony are found in the smallest quantities (Schroeder et al., 1987).
 3          Potential sources of trace metals found in fine airborne particles are primarily
 4      anthropogenic and include combustion of coal and oil, wood burning, waste incineration, and
 5      metal smelting operations. Biomass burning which includes residential wood combustion and
 6      forest fires, is another source for the release of trace  elements in the atmosphere.  In a
 7      profile of biomass burning, zinc was the characteristic trace element present in the fine
 8      particles in concentration (0.0866 ± 0.0355 %) of primary mass emitted.  Other trace
 9      elements present were Cl (1.9083 + 0.6396 %), K (3.9926 ± 1.2397 %) and S (0.5211 +
10      0.1761  %) (Chow et al., 1992).
11          The chemical composition of paniculate matter  analyzed in New Jersey as part of the
12      Airborne Toxic  Element and Organic Substances project (ATEOS), identified the trace
      April 1995
3-70
DRAFT-DO NOT QUOTE OR CITE

-------
 l     elements Pb, Fe, Zn, V and As (Daisey, 1987; Morandi et al., 1991).  The main source for
 2     atmospheric lead concentration is the combustion of leaded gasoline in motor vehicles.
 3     However with increased use of unleaded gasoline, levels of atmospheric lead have been
 4     reduced, and other sources of lead tend now to be more significant components of the
 5     residual lead.  Morandi (1985) has reported evidence of contributions to airborne lead from
 6     resuspended soil, oil burning and small scale smelting,  which taken together accounted for
 7     more than half of the airborne lead at a New Jersey site.  Vanadium levels were derived
 8     from oil burning for space heating and power production, while Zn is attributed to a zinc
 9     smelter in the area (Daisey, 1987).
10           Road dust  aerosols are analyzed for trace elements in a variety of studies (Barnard
11     et al., 1987; Barnard et al, 1988; Warren et al., 1987).  Recent source  apportionment studies
12     in California's South Coast Air Basin, provide additional information on trace element
13     concentrations in roadside dusts as well as in motor vehicle exhaust for particle sizes
14     < 2.5 urn  (Watson et al., 1994). In addition to elemental carbon,  Al, Si, K, Ca, Ti and
15     Fe were present in paved road dust in abundances which exceeded  1%. Elevated
16     concentrations of Pb and Br were detected, which illustrated the deposition from the tailpipe
17     exhaust from vehicles that burned leaded fuels  (Watson et al.  1994; Chow et al., 1992).
18     Significant amounts of SO=4, Br, Cl", and Pb were detected in the motor vehicle exhaust
19     profile, though Pb levels were much lower than those reported in earlier tests (Watson et al.,
20     1994; Pierson and Brachaczek, 1983).
21           Ambient measurements of the mass and chemical composition of PM10 and PM2 5, and
22     associated source profiles have been taken through the years.  Data base summaries identify
23     locations, sampling times and chemical species of data available since 1988, complementing
24     previous existing databases (Watson and Chow, 1992; Lioy et al., 1980).  Size  specific
25     measurements show that over 90% of the mass from geological material is in the coarse
26     particle size fraction, while the combustion related source categories contained  —90% of
27     their mass concentrations in the PM2-5 size fraction (Chow et al., 1992). In a municipal
28     incinerator  profile, elements in the fine particle fraction include Cu, Zn, As, Cd, Sb, Pb and
29     Ba, while trace elements in the coarse particle fraction include Ca, Cr, Mn,  and Ni (Olmez
30     et al., 1988).  In an oil-fired power plant, trace elements such as V, Ni, Co, Ba and Cu are
31      present in both fine and coarse particles (Olmez et al., 1988).

       April 1995                                3.7 j       DRAFT-DO NOT QUOTE OR  CITE

-------
 1          Although a knowledge of the elemental and ionic composition of ambient particles is
 2     necessary in order to understand their sources and chemistry, the chemical forms in which
 3     important species exist are not known.  For example, sulfates, nitrates and ammonium ions,
 4     which are the main constituents of fine particles, may exist in forms other than simple
 5     ammonium salts (Pitts and Pitts, 1986). Table 3-10 lists some compounds identified in
 6     aerosols by a roadway at Argonne National Laboratory, and Table 3-11 lists compounds
 7     observed in aerosols in a forested area,  at State College, Pennsylvania (Tani et al., 1983).
 8     However, there are uncertainties associated with the compounds shown in Tables 3-10 and
 9     3-11.  Tani et al. pointed out that both physical and chemical changes may occur during or
10     following impaction of aerosol particles on a collector, which would lead to the formation of
11     compounds not  initially present in the ambient aerosols (Tani et al.,  1983).
12
13
         TABLE 3-10. COMPOUNDS OBSERVED IN AEROSOLS BY A ROADWAY AT
                           ARGONNE NATIONAL LABORATORY
                                            K2Sn(S04)2
       CaCO3                                (NH4)2Co(SO4)2 . 6H2O
       CaMg(CO3)2                           (NH4)3H(SO4)2 (letovicite)
       CaS04.2H20                           3(NH4N03).(NH4)2S04
       (NH4)2Pb(S04)2                        2(NH4N03) . (NH4)2SO4
       (NH4)2Ca(SO4)2.H2O                   NH4MgCl3.6H2O
       (NH4)HSO4                           NaCl
       (NH4)2S04 _ (NH4)2Ni(S04)2 . 6H2O
      Source: Tani et al., 1983.
        TABLE 3-11.  COMPOUNDS OBSERVED IN AEROSOLS IN A FORESTED AREA,
      	STATE COLLEGE, PA.	
                                         (NH4)2S04
                                  (NH4)3H(SO4)2 (letovicite)
                                         (NH4)HSO4
                                    2(NH4N03).(NH4)2SO4
                                       (NH4)2Pb(S04)2
      Source: Tani et al., 1983

      April 1995                             3_72      DRAFT-DO NOT QUOTE OR CITE

-------
  1            Metals such as Al, Ca, Fe, Mg and Pb known to be present in atmospheric aerosols,
  2      also exist in uncertain chemical forms (Pitts and Pitts, 1986).  This  is partially due to the use
  3      of analytical techniques that normally provide information on total metal content (Schroeder
  4      et al., 1987).  It is generally assumed that many of the elements, especially from combustion
  5      sources, are present  in the form of oxides (Olmez et al. 1988), while trace elements in
  6      incinerator emissions may be in the form of chlorides  (Schroeder, 1987).  Data from Los
  7      Angeles indicate that arsenic may be present in two chemical forms in atmospheric aerosols,
  8      as arsenite and arsenate. Both forms were identified in both the fine and coarse particle
  9      fractions (Rabano  et al., 1989).  Fe2O3, Fe3O4, A12O3, and A1P04 have been identified in
 10      roadside particulate matter (Biggins and Harrison, 1980).  Ca and Mg may exist in the form
 11      of oxides (i.e., CaO, MgO), although in the presence of water, Stelson and Seinfeld (1981)
 12      suggest that, on equilibrium considerations, CaO  and  MgO should react to form their
 13      hydroxides, Ca(OH)2 and Mg(OH)2, respectively. Similarly the oxides Na2O and K2O
 14      should form NaOH and KOH when water is present.   Lead has been observed in roadside
 15      particulate matter  in  a wide  variety of forms, such as  PbSO4, Pb3O4, PbSO4.(NH4)2SO4,
 16      PbO.PbS04, 2PbC03.Pb(OH)2, 2PbBrCl.NH4Cl, PbBrCl, (PbO)2PbBrCl,
 17      3Pb3(PO4)2.PbBrCl,  and elemental lead  (Biggins and Harrison,  1980; Post and Buseck,
 18      1985).
 19            Heterogeneous oxidation of sulphur dioxide in air can be catalyzed by species such as
 20      iron, manganese (Barrie and Georgii, 1976) and cadmium,  while vanadium is suspected to
 21      catalyze the formation of sulfuric acid during oil combustion.  Oxides of iron, manganese
 22      and  lead are reported to absorb SO2 (Schroeder, 1987).
 23            It has been  suggested  that the elements arsenic,  cadmium, manganese, nickel, lead,
 24      antimony, selenium,  vanadium and zinc  volatilize at high temperatures during fossil fuel
 25      combustion and condense uniformly on surfaces of entrained flyash particles  as  the
 26      temperature falls beyond the combustion zone (Linton et al., 1976;). Accumulation of trace
 27      metals in the fine fraction of airborne dust sampled in  iron  foundries showed Pb and Zn
28      localized on the surface of the fine particles (Michaud  et al., 1993).  From the viewpoint of
29      toxicity, such emissions are more important than natural sources where trace elements  are
30      usually bound within the matrix of natural aerosols and thus less mobile and bioavailable
31      (Schroeder,  1987).

        April 1995                               3.73       DRAFT-DO NOT QUOTE OR CITE

-------
 1            Trace metal compounds found in road dust can accumulate from anthropogenic or
 2      natural sources.  Subsequently these can become re-entrained in the atmosphere. In such
 3      samples lead and zinc were found to be strongly associated with carbonate and iron-
 4      manganese oxide phases, with small amounts being associated with an organic phase.  Half
 5      of cadmium was associated with carbonate and iron-manganese oxide phases, while copper
 6      was mainly associated with the organic phase.  These associations  influence  the relative
 7      mobility and bioavailability of trace metals in the environment (Harrison et al., 1981).
 8            Resuspension of particles from contaminated surfaces may also contribute to an
 9      increase in the toxic trace elements in airborne particles (Kitsa et al., 1992,  Kitsa and Lioy,
10      1992; Pastuszka and Kwapulinski, 1988; Falerios et al., 1992). Kitsa et al.  (1992) measured
11      elemental concentrations in particles resuspended from a waste site in New Jersey.  Close to
12      the resuspension source, coarse particles were dominant, but farther  downwind from the site,
13      fine particles were prevailing.  The fine particles were enriched in chromium and lead,
14      indicating the potential for elevated human exposure through inhalation.  Chromium may
15      exist in different valence states, but the most stable and abundant are the trivalent and
16      hexavalent states. Hexavalent chromium is classified as a known respiratory carcinogen in
17      humans.
18            Oxidation of the species present in aerosols  results from interaction with various
19      atmospheric oxidants,  such as molecular  oxygen, ozone or hydrogen peroxide.  The presence
20      of oxides of As,  Cd, Co, Cr, Cu, Hg, Mn, Ni, Pb, Sb, Se, V and Zn has been measured in
21      emissions of cement plants, blast furnace and sintering operations, secondary iron foundries,
22      non-ferrous smelting of arsenic-bearing ores, zinc and lead smelters  and many other sources
23      (Schroeder et al., 1987).
24            Sulphation, and possibly nitration, of metallic oxides  can be surmised to be  an
25      important transformation as particles age.  A statistical assessment of multielemental
26      measurements in a study in the rural and urban atmospheres of Arizona showed strong
27      correlations of lead, copper, cadmium and zinc with sulfates in the rural atmosphere and
28      moderate correlation of lead and copper with sulfates and nitrates in urban atmosphere
29      (Moyers et al., 1977).  Nickel has also great affinity for sulfur which may lead to the
30      emission of nickel sulfate containing particulates from combustion sources.  In the absence of
        April 1995                               3-74       DRAFT-DO NOT QUOTE OR CITE

-------
  1      sulfur, nickel oxides or complex metal oxides containing nickel may form (U.S.
  2      Environmental Protection Agency, 1986a).
  3            Lead is emitted in the air from automobiles as lead halides and as double salts with
  4      ammonium halides (e.g. PbBrC1.2NH4Cl).  From mines and smelters, the dominant species
  5      are PbSO4, PbO.PbSO4, and PbS. In the atmosphere lead is present as sulfate with minor
  6      amounts of halides. Lead sulfide is the main constituent of samples associated with ore
  7      handling and fugitive dust from open mounds of ore concentrate. The major constituents
  8      from sintering and blast furnace operations appeared to  be PbSO4 and PbO.PbSO4
  9      respectively (U.S. Environmental Protection Agency, 1986b).
10
11      3.3.8  Removal Processes
12            Removal of accumulation mode aerosol particles  from the atmosphere occurs largely
13      by the precipitation process (e.g., Slinn, 1983).  These  particles are the dominant particles on
14      which cloud droplets form (cloud condensation nuclei, CCN); once  a cloud droplet (of
15      diameter of a few up to about 20 micrometers) is formed, it is much more susceptible to
16      scavenging and  removal in precipitation than is the original submicrometer particle. The
17      fraction of aerosol particles incorporated in cloud droplets on cloud formation is the subject
18      of active current research.   Earlier work yielded a fairly wide spread in this fractional
19      incorporation, based in part on limitations of then existing techniques and in part on
20      definitions of incorporation efficiencies (based on number, mass, light scattering efficiency,
21      or amount of specific compounds; ten Brink et al., 1987).  More recent work indicates a
22      high fractional incorporation at low concentrations of aerosol particles decreasing as the
23      aerosol particle  loading increases (Leaitch et al., 1992;  Gillani et al., 1992).  Model
24      calculations of the efficiency of incorporation of accumulation-mode aerosol particles into
25      cloud droplets and precipitation are highly sensitive to assumptions and approach (Jensen and
26      Charlson,  1984; Flossmann et al., 1985; Hanel, 1987; Ahr et al., 1989; Albeit et al.,  1990).
27            The dominance of precipitation removal processes for accumulation mode particles
28      results in high variabity in temporal patterns of aerosol  loadings, that may be attributed to the
29      episodicity of precipitation events  and synoptic scale meteorology that delivers air of
30      differing origins to a given location (e.g., Waldman et al,,  1990).  This  variability leads to
31      difficulties in attempts to estimate mean residence times based on budget considerations

        April 1995                                3.75       DRAFT-DO NOT QUOTE OR CITE

-------
 1     (Junge, 1963; Schwartz, 1979). A unique approach to estimation of the mean residence time
 2     of accumulation mode aerosol particles was presented by measurement of the decay of
 3     atmospheric concentrations of Cei37 at several mid-latitude surface stations in Europe and
 4     Asia in the weeks following the Chernobyl accident (Cambray et al., 1987); the Ce137 was
 5     present largely  in this size range.  This study led to an estimate for the mean residence time
 6     of 7 days , consistent with other estimates. It may be noted, however, that this residence
 7     time is applicable to particles in the free troposphere,  where the Ce137 was mainly present
 8     during the several week period. The mean residence time of accumulation mode particles in
 9     the  boundary layer is expected to be somewhat less (Benkovitz et al., 1994).
10
11
12     3.4 TRANSPORT AND TRANSFORMATIONS TO SECONDARY
13          PARTICIPATE MATTER
14     3.4.1  Aqueous-Phase Chemical Equilibria and Chemical Kinetics of
15            Transformations to Secondary Particulate Matter
16     3.4.1.1  Aqueous-Phase Equilibria
17          The liquid water content of the atmosphere, WL,  is usually expressed either in g of
18     water per m3 of air or as a dimensionless volume fraction L (e.g., m3 of liquid water per
19     m3  of air).  Typical liquid water content values are 0.1 to 1 g m"3  (L= 10"7- 10"6) for
20     clouds, 0.05 to 0.5 g nV3 (L= 5 x 10'7 - 5 x  10"6) for fogs, and only 10'5 to 10'4g in3
21     (L= 10-n-10-10) for aerosols.
22          For dilute solutions the equilibrium distribution of a reagent gas A between the gas and
23     aqueous phases is given by Henry's law
24                                        [A]  = H^A                               (3-15)
25
26     where pA is the partial pressure of A in the gas-phase, [A] is the equilibrium aqueous-phase
27     concentration of A and HA is the Henry's  law coefficient for species A. The customary units
28     of //A are mole I"1 atm"1. HA can be viewed  as the equilibrium constant of the reaction
29
30                                       A(g)  £> A(aq)                             (3-16)
31

       April 1995                              3-76       DRAFT-DO NOT QUOTE OR CITE

-------
1 Table 3-12 gives the Henry's law coefficients of some atmospheric gases in liquid water
2 at 298 K. The values given reflect only the
3 subsequent fate of the dissolved species A.
physical solubility of the gas regardless of the
Some of the species included in Table 3-12
4 dissociate after dissolution or react with water. Henry's law constants do not account for
5 these processes, and the modifications necessary will be discussed in the next paragraph.
6 Henry's law coefficients generally decrease
7 solubilities at higher temperatures (Seinfeld,
8
9
for increasing temperatures, resulting in lower
1986).


TABLE 3-12. HENRY'S LAW COEFFICIENTS OF SOME ATMOSPHERIC GASES
DISSOLVING
Species
02
NO
C2H4
NO2
03
N2O
CO2
H2S
S02
CH3ONO2
CH3C(O)O2NO2
CH302
OH
HNO2
NH3
CH3OH
CH3OOH
CH3C(O)OOH
HC1
HO2
HCOOH
HCHO
CH3COOH
H202
HN03
NO3
IN LIQUID WATER
H (M/atm) (298 K)
1.3xlQ-3
1.9xlO-3
4.8xlO-3
l.OxlQ-2
1.13xl(T2
2.5xlO-2
3.4xlQ-2
0.12
1.23
2.6
2.9
6.0
25.
49.
75.
220.
227.
473.
727.
2.0xl03
3.5xl03
6.3xl03
8.7xl03
7.45xl04
2.1xl05
2.1xl05
April 1995
3-77
DRAFT-DO NOT QUOTE OR CITE

-------
1
2
3
     Several gases, after dissolving in the aqueous-phase, ionize and establish an
aqueous-phase chemical equilibrium system. For example for SO2,

                                 S02(g) * S02- H20
                                                                                    (3-17)
                                    SO2- H2O ^ HSO3 + H +
                                                                              (3-18)
                                      HSO   *» SO
                                                   2"
                                                                                    (3-19)
5
1
2
3
4
5
6
7
8
      with
                     H
                       so.
[SO2- H2O]
   Pso,
                                         [SO2- H2O]
                                                                   [HSO~]
                                                                              (3-20)
ATsl and Ks2 are the first and second dissociation constants for SO2. It is convenient to
consider the total dissolved sulfur in oxidation state IV as a single entity and refer to it as
                              = [SO2- H2O]  +  [HSO3]  =
                                                                  2-,
                                                                                    (3-21)
     The three sulfur species are in rapid equilibrium and therefore [S(IV)] changes only
when SO2 is transferred between the gas and aqueous phases. The total dissolved sulfur,
S(IV), can be expressed as a function of only the pH and the partial pressure of SO2 over the
solution by:
                                                                                    (3-22)
                                                             [H + ]
      April 1995
                                        3.78       DRAFT-DO NOT QUOTE OR CITE

-------
 1      The above equation can be expressed in a form similar to Henry's law as

                                                 =  Hs*IV)Ps02                           (3-23)
 2
 3      where Hf(IV)  is the effective (or modified) Henry's law coefficient given for S(IV) by
                                                                                       (3-24)
 4
 5      The modified Henry's law coefficient relates the total dissolved S(IV) (and not only with the
 6      SO2 vapor pressure over the solution. The effective Henry's law coefficient always exceeds
 7      the Henry's law coefficient, indicating that the dissociation of a species enhances its
 8      solubility in the aqueous phase.
 9           Several of the species that are in rapid equilibrium can be also considered as single
10      entities:

                              [S(IV)]  = [H2S04(aq)] + [HSO4] + [SO42~]

11
                                    [N(V)] = [HN03(aq)] + [NO~]

12
                                    [H02T]  = [HN02(aq)]  + [O~]

                                 [HCHOT] = [HCHO]  + [H2C(OH)2]
13
14      Equations relating the total  concentrations of these aqueous-phase species with the
15      corresponding equilibrium concentrations of the gas-phase species can be derived similarly to
16      those for S(IV).

        April 1995                               3.79       DRAFT-DO NOT QUOTE OR CITE

-------
 1     3.4.1.2 Aqueous-Phase Transformation of SO2 to Sulfate
 2          The aqueous-phase conversion of dissolved SO2 to sulfate is thought to be the most
 3     important chemical transformation in cloudwater.  Dissolution of SO2 in water results in the
 4     formation of three chemical species: hydrated SO2 (SO2 • H2O), the bisulfite ion (HSO^) and
 5     the sulfite ion (S0>3~).  At the pH range of atmospheric interest (pH =2-7)  most of the  S(IV)
 6     is in the form of HSOJ, whereas at low pH (pH <2), all of the S(IV) occurs as SO2 • H2O.
 7     At higher pH values (pH  >7), (SO^~) is the preferred S(IV)  state (Seinfeld, 1986). The
 8     individual dissociations are fast, occurring on timescales of milliseconds or less (Martin,
 9     1984; Schwartz and Freiberg, 1981; Seinfeld, 1986).  Therefore, during a reaction
10     consuming  one of the three species, SO2  • H2O, HSO^, or SO^", the corresponding
11     aqueous-phase equilibria are re-established instantaneously.  The dissociation of dissolved
12     SO2 enhances its aqueous solubility and the total amount of dissolved S(IV)  always exceeds
13     that predicted by Henry's law for SO2 alone and is quite pH dependent.  The Henry's law
14     coefficient for SO2 alone, Hso , is 1.23 M atm"1 at 298 K, while  for the same temperature,
15     the effective Henry's law coefficient for S(IV),  HS(IV> is 16.4 M atm"1 for pH=3,
16     152 M atm"1 for pH=4 and 1,524 M atm"1  for pH=5. Equilibrium S(IV) concentrations for
17     SO2 gas-phase concentrations of 0.2-200 ppb, and over a pH range 1-6 vary approximately
18     from 0.001 to 1000 mM.
19          Several pathways for S(IV) transformation to S(VI) have been identified involving
20     reactions of S(IV) with O3, H2O2, O2 (catalyzed by Mn + and Fe3+), OH,  SO5~, HSO5~
21     SO^, PAN, CH3OOH, CH3C(O)OOH, HO2,  NO3, NO2, N(III), HCHO and Cl^  (Pandis
22     and Seinfeld, 1989a).
23          Although ozone reacts very slowly with  SO2 in the gas phase, the aqueous-phase
24     reaction is rapid. The possible importance of O3 as an aqueous-phase oxidant for S(IV)  was
25     first suggested by Penkett (1972) and the kinetics of
26
                                     S(IV) + O3  -»  S(VI) + O2                          (3-25)
27
28     have been studied by several investigators (Erickson et al., 1977;  Penkett et al., 1979;
29     Maahs, 1983).  Hoffmann and Calvert (1985), after a detailed investigation  of existing
       April 1995                               3-80       DRAFT-DO NOT QUOTE OR CITE

-------
 1      experimental kinetic and mechanistic data, suggested the following expression for the rate of
 2      the reaction of S(IV) with dissolved ozone:
 3
                                                H20]  + kjtHSOj]  + k2[SO32 "])[O3]        (3-26)
 4
 5      recommending the values kQ =2.4 x 104 M'1  s'1, k\ =3.7 x 105 M'1 sl and, k2 = l.5 x 109
 6      M"1 s"1. They also proposed that this reaction proceeds by nucleophilic attack on  ozone by
 7      SO2 • H2O, HSO^~, and 803". An increase in the aqueous-phase pH results in an increase
 8      of all three, [SO2 •  H2O], [HSO^~] and [SO^"], equilibrium concentrations and therefore in
 9      an increase of the overall reaction rate.  For an ozone gas-phase mixing ratio of 30 ppb, the
10      reaction rate varies from less than 0.001 mM h"1 (ppb SO^"1 at pH=2 (or less than 0.01%
11      SO2 (g) h'1 (g water /m3 air)'1) to  3,000 mM IT1 (ppb SC^)'1 at pH=6 (7,000% SO2 (g) h'1
12      (g water /m3 air)"1). The gas-phase SO2 oxidation rate is of the order of  1% h"1  and
13      therefore the S(IV) heterogeneous  oxidation by ozone  is significant for pH values greater
14      than 4.  The strong positive dependence of the reaction rate on the pH renders  this  reaction
15      self limiting. The production of sulfate by this reaction lowers the pH and effectively
16      decreases the  rate of further reaction. The availability of atmospheric ozone guarantees that
17      this reaction will play an important role both as a sink of gas-phase SO2 and as a cause of
18      cloudwater acidification as long as the pH of the atmospheric aqueous phase exceeds 4.
19          Hydrogen peroxide, H2O2, is one of the most effective oxidants of S(IV)  in clouds and
20      fogs (Pandis and Seinfeld, 1989a).  H2O2 is very soluble in water and under typical ambient
21      conditions  its  aqueous-phase concentration is approximately six orders of magnitude higher
22      than that of ozone.  This reaction has been studied in detail by several investigators
23      (Hoffmann and Edwards, 1975; Penkett et al., 1979; Martin and Damschen, 1981;  Cocks et
24      al.,  1982; Kunen et al., 1983; McArdle and Hoffmann,  1983) and the reproducibility of the
25      measurements suggests a lack  of susceptibility of this reaction to influence of trace
26      constituents.   The proposed rate expression is (Hoffmann and Calvert, 1985)
27
       April 1995                                3.81       DRAFT-DO NOT QUOTE OR CITE

-------
                               R   =  _ d[S(IV)]  = k[H+][H202][HS03]                   (
                                2        dt           i  + K[H+]
 1
 2     with k=1.45 x 107 M'1 s'1 and K=13 M'1 at 298 K. Noting that  H2O2 is a very weak
 3     electrolyte, and from 6.3, 6.4, and 6.5 that [H+]  [HSOJ]  =  Hso Kslpso  and that for
 4     pH>2,l+/sf[H+]—1, one concludes that the rate of this reaction is practically pH
 5     independent in the pH range of atmospheric interest.  For a H2O2(g) mixing ratio of 1 ppb
 6     the rate is roughly 300 mM tr1  (ppb SO^'1 (700% SO2(g)h'1 (g water /m3 air)"1).  The near
 7     pH independence can also be viewed as the result of the cancellation of the pH dependence
 8     of the S(IV) solubility and the reaction rate constant.  The reaction is very fast and indeed
 9     both field measurements (Daum  et al.,  1984) and theoretical studies (Pandis and Seinfeld,
10     1989b) have suggested that H2O2(g) and SO2(g) rarely coexist in clouds and fogs. The
11     species with the lowest concentration before the cloud or fog formation is the limiting
12     reactant, and is rapidly depleted inside  the cloud or  fog layer.
13          Organic peroxides have been also proposed as potential aqueous-phase oxidants  of
14     dissolved sulfur (Graedel and Goldberg, 1983; Lind and Lazrus, 1983; Hoffmann and
15     Calvert,  1985).
16          Simulations for typical continental clouds suggest that these reactions are of minor
17     importance for the S(IV) oxidation and represent small sinks for the gas-phase hydroperoxide
18     (0.2% CH3OOH h-1) and peracetic acid (0.7% CH3C(O)OOH h'1).    The S(IV) oxidation
19     by O2 is known to be catalyzed by Fe3+  and Mn2+
20
21                                            Mn2+,Fe3+
22                              S(IV) + 1 O2	> S(VI)                    (3-28)
23
24     This reaction has been the subject of considerable interest (Hoffmann and Boyce, 1983;
25     Martin, 1984;  Hoffmann and Jacob,  1984;  Hoffmann and Calvert, 1985;  Clarke and
26     Radojevic, 1987) and significantly different measured reaction rates, rate laws and pH
27     dependencies have been reported (Hoffmann and Jacob,  1984). Martin and Hill (1987a,b)
28     have demonstrated that this reaction  is  inhibited as ionic strength increases.  They explained
       April 1995                               3-82      DRAFT-DO NOT QUOTE OR CITE

-------
 1     most of the literature discrepancies by  differences in these factors during the various
 2     laboratory studies.
 3           In the presence of oxygen, iron in the ferric state, Fe(III), catalyzes the oxidation of
 4     S(IV) in aqueous solutions. Fe(II) appears not to catalyze directly the reaction and is first
 5     oxidized to Fe(III) before S(IV) oxidation can begin (Huss et al.,  1982a,b).
 6           For pH values from 0 to 3.6 the iron-catalyzed S(IV) oxidation rate is first order in
 7     iron, first order in S(IV) and  is inversely proportional to  [H+] (Martin and Hill, 1987a),
 8
                                    r _  _ d[S(IV)] _  k [Fe3+][S(IV)]                       (3_29)
                                             dt        l    [H+]
 9
10
11     This reaction is inhibited by ionic strength and sulfate and these effects are described by:
12
                                                                                          (3-30)
13
14      and
                                                                                          (3-31)
                                                1 + 150[S(VI)]2/3
15
16      where / is the ionic strength of the solution and [S(VI)] is in M.  A rate constant k = 6 s
17      has been recommended by Martin and Hill (1987a). Sulfite appears to be almost as equally
18      inhibiting as sulfate.
19           The rate expression for the same reaction changes completely above pH 3.6.  This
20      suggests that the mechanism of the reaction differs in the two pH regimes, and is probably a
21      free radical chain at high pH and a non radical mechanism at low pH (Martin et al., 1991).
22      The low solubility of Fe(III) above pH 3.6 presents special experimental problems. At high
23      pH the reaction rate depends on the amount of iron in solution, rather than on the  total
24      amount of iron present.  At this range the reaction is second order in dissolved iron (zero
25      order above the solution iron saturation point) and first order in S(IV). The reaction is still

        April 1995                                3-83      DRAFT-DO NOT QUOTE OR CITE

-------
 1     not very well understood and Martin et al. (1991) proposed the following phenomeno logical
 2     expressions (in M s"1)
                              pH4.0: -          =  lxl09[S(IV)][Fe3+]2
                                          dt

                                pH5.0-6.0:  d[S(IV)]  = lxlO-3[S(IV)]
                                               dt

                                 pHT.O:  -d[S(IV)] =  1 X KT4[S(IV)]
                                             dt
 3
 4
 5     for the following conditions:
               [S(IV)] «  10/xM,[Fe3+]>0.1//M, K0.01M, [S(VI)] < 100M,  and T=298K.
 6
 7     Note that iron does not appear in the pH 5-7 rates because it is assumed that a trace of iron
 8     will be present under normal atmospheric conditions.  This reaction is important in this high
 9     pH regime (Pandis and Seinfeld, 1989a, b; Pandis et al., 1992).
10          Martin et al.  (1991) also  found that non-complexing organic molecules (e.g. acetate,
11     trichloroacetate,  ethylalcohol, isopropyl alcohol, formate,  allyl alcohol, etc.) are highly
12     inhibiting at pH  values of 5 and above, and are not inhibiting at pH values of 3 and below.
13     They calculated that, for remote clouds, formate would be the main inhibiting organic, but
14     by less than 10%.  In contrast, near urban areas formate could reduce the rate of the
15     catalyzed oxidation by a factor of 10-20 in the  high pH regime.
16          The manganese catalyzed S(IV) oxidation was initially thought to be inversely
17     proportional to the H+ concentration.  Martin and Hill (1987b) suggested that ionic strength,
18     not hydrogen ion, accounts for the pH dependence of the rate.  These authors were  also able
19     to explain some  unusual behavior described in  the literature on this reaction and to partially
20     reconcile some of the literature rates.  The manganese catalyzed reaction obeys zero-order
21     kinetics in S(IV) in the concentration regime above 100 mM S(IV),
22
       April 1995                               3-84       DRAFT-DO NOT QUOTE OR CITE

-------
                                        _d[SOV)] = ko[Mn2+]2                          (3.32)

                                                     -4.07 J F
                                          k0=k0*10  ^
  1      with A:* =  680 M'1 s'1 (Martin and Hill, 1987b). For S(IV) concentrations below 1 mM the
  2      reaction is first order in S(IV),

                                                = k0[Mn2+][S(IV)]                        (3-33)
                                         dt
                                                    -4.07J T
                                          k = k0*10  1+Jr
 3
 4      with k*0 =  1,000 M'1 s"1  (Martin and Hill, 1987b). It is still not clear which rate law is
 5      appropriate for use in atmospheric calculations, although Martin and Hill (1987b) suggested
 6      the provisional use of the first order, low S(IV) rate.
 7           When both Fe3+ and Mn2+ are present in atmospheric droplets, the overall rate of the
 8      S(IV) reaction is enhanced over the sum of the two individual rates.  Martin (1984) reported
 9      that the rates measured were 3 to 10 times higher than expected from the sum of the
10      independent rates.  Martin and Good (1991) obtained at pH 3.0 and for [S(IV)]  < 10 mM
11      the following rate  law
12
          d[S(IV)]

                     75o[Mn(II)][S(IV)] + 2600[Fe(III)][S(IV)]
13                                                                                      (3-34)
14      and a similar expression for pH 5.0 in agreement with the work of Ibusuki and Takeuchi
15      (1987).
16          Free radicals, such as OH and HO2, either  heterogeneously scavenged by the aqueous
17      phase or produced in the aqueous phase, participate in a series of aqueous phase reactions
18      (Graedel and Weschler, 1981; Chameides and Davis, 1982;  Graedel and Goldberg,  1983;
19      Schwartz, 1984; Jacob, 1986; Pandis and Seinfeld, 1989a).
20          Pandis and Seinfeld (1989a) proposed  that under typical remote continental conditions
21      there are two main radical pathways resulting in  the conversion of S(IV) to S(VI):

        April 1995                               3.35      DRAFT-DO NOT QUOTE OR CITE

-------
                          S(IV)(+OH) -» SOgC+Oj) -* HSO^+HSOg) -* S(VI)             (3-35)
                           S(IV)(+OH) -^  SO^ SO4(+Cr,HSC>3)  -* S(VI)              (3-36)
 1
 2     with the first of these two pathways typically being faster that the second.
 3          Nitrogen dioxide has a low water solubility and therefore its low resulting
 4     aqueous-phase concentrations suggests that its oxidation of S(IV)
 5
                H0
2         -
 6                       2NO2  + HSO3- + - > 3H+ + 2NO2~  + SO|~            (3-37)
 7
 8     should be of minor important in most cases.  This reaction has been studied by Lee and
 9     Schwartz (1983) at pH 5.0, 5.8 and 6.4 and was described as a reaction that is first order in
10     NO2 and first order in S(IV), with a pH-dependent rate constant.  The evaluation of this rate
1 1     expression was considered tentative by Lee and Schwartz, in view of evidence for the
12     formation of a long-lived intermediate species.  The apparent rate constant was found to
13     increase with pH.  This reaction is considered of secondary importance at the concentrations
14     and pH values representative of clouds. However, Pandis and Seinfeld (1989b) reported that
15     for fogs occurring in urban polluted areas with high NO2 concentrations this reaction could
16     be a major pathway for the S(IV)  oxidation, if the atmosphere has enough neutralizing
17     capacity, e.g. high NH3 (g) concentrations.
18          Sulfite and bisulfite can form complexes with various dissolved aldehydes.  One
19     important example is the  reaction  of sulfite or bisulfite with formaldehyde to produce
20     hydroxymethanesulfonate ion (HMS) (Boyce and Hoffmann,  1984; Munger et al.,  1984,
21     1986; Olson and Hoffman, 1989;  Faccini et al.,  1992).
22          The  HMS formed acts as a S(IV) reservoir protecting it from further oxidation, and  its
23     formation has been advanced to explain high S(IV) concentrations that have been observed in
24     clouds and fogs.  The rates of S(IV) complexation and oxidation are highly dependent on
25     cloud pH and on the concentrations of HCHO and oxidants.  Characteristic tunes for S(IV)
26     depletion through complexation and oxidation can be compared for typical ranges of HCHO,
27     H2O2, and pH.  At pH values below about 4, the rate of reactions 6.26 and 6.27 are several
28     orders of magnitude slower than the reaction of S(IV) with dissolved H2O2.  Thus, in this

       April 1995                               3-86      DRAFT-DO NOT QUOTE OR CITE

-------
  1      range oxidation predominates over complexation.  The characteristic times of the two
  2      reactions become approximately comparable at pH around 5 so that complexation with
  3      HCHO becomes competitive with oxidation by H2O2.  When pH exceeds 6, the reactions of
  4      S(IV) with HCHO became more important than reaction with H2O2. HMS formation can
  5      inhibit S(IV) oxidation if the S(IV) complexation rate is comparable to, or greater than, the
  6      S(IV) oxidation rate and the rate of SO2 mass transport into the drop controls the rate of
  7      S(IV) oxidation.  The effectiveness of HMS as a S(IV) reservoir depends critically on its
  8      resistivity to OH  attack.
  9
10      3.4.1.3 Aqueous-phase Transformation of NO2 to HNO3 and NH4NO3
11          Aside from  reaction of N2O5 with liquid water, there does not appear to be any other
12      aqueous-phase reaction of nitrogen oxides that contributes substantially to atmospheric
13      nitrate.
14          In contrast to  the sulfate system, the nitrate system exhibits a gaseous equilibrium that
15      admits to a substantial gas-phase fraction (as nitric acid vapor) under ambient conditions.
16      Thus, the chemical  kinetics of the aqueous-phase oxidation of NO by O2 has been
17      reexammed by  two  groups (Lewis and Deen, 1994; Pires et al.,1994), with confirmation of a
18      third-order rate law,
                                       R = k[NO(aq)]2[02(aq)],                          (3-38)
19
20      analogous to the gas-phase reaction, with k = (7 ±  1) x 106 M"2 s"1 at 296 K.  Evaluation of
21      the rate of this  reaction in cloudwater confirms that the reaction rate is negligible under
22      atmospheric conditions, as indicated earlier by Schwartz and White  (1983).
23
24      3.4.2  Transport and Transformations in Plumes
25          In the 1970s, many field studies were plume studies or urban-scale studies, and most
26      models were Lagrangian  and limited to linearized treatment of chemistry and  other non-linear
27      processes.  Some of these field studies, along with regional visibility information and back-
28      trajectories from local pollution episodes, pointed to the existence of long range transport and
29      to the regional nature of air pollution and haze (Hall et al., 1973; Gillani and  Husar, 1976;
30      Wolff et al., 1977).  In response,  some of the major field studies in the 1980s had a regional

        April 1995                               3.37       DRAFT-DO NOT QUOTE OR CITE

-------
 1      scope with focus on acidic depositions, oxidants, or aerosols and visibility. That decade also
 2      saw major strides in measurement technology and in the development of increasingly
 3      sophisticated Eulerian air quality models with explicit treatment of non-linear processes.
 4      In these models, however, the treatment of plumes, particularly point-source plumes, was
 5      grossly distorted by varying degrees depending on the spatial resolution of the grid.  New
 6      interest also began to emerge in global climate change, global data, and global modeling.
 7      In the decade of the 1990s, the principal interests in modeling and measurements appear to
 8      be in two areas: global-scale issues, with particular focus on clouds and aerosols;  and,
 9      regional and sub-regional issues, with special interests in comprehensive linked study of
10      oxidants, aerosols and acidic depositions, and in multi-scale interactions (e.g., nested
11      gridding and the treatment of subgrid-scale processes related to  plumes, clouds,  and air-
12      surface interactions).
13           Topics related to field measurements are also covered in other parts  of this document:
14      methodologies for sampling and analysis of PM and acidic deposition in Chapter 4; ambient
15      air measurements of PM concentrations and properties in Chapter 7; and field studies of
16      visibility and PM in Chapter 10. The focus in this  section is on North American field
17      studies of the past 15 years or so, particularly as they relate  to the following objectives:
18      better understanding of atmospheric processes (formation, transformation,  transport, and
19      removal) which modify the concentration, size and composition  of PM; evaluation of source-
20      or receptor- oriented models of PM air quality;  and generation of model inputs.
21
22      3.4.2.1  Field Studies of Transport Processes
23           Except for the gravitational settling of coarse particles (included in dry deposition), the
24      transport of PM is  similar to that of gases.  Following their emissions, gases  and fine
25      aerosols rise due to buoyancy effects, are advected downwind by  the prevailing mean flow
26      field, and are dispersed  horizontally and vertically by ambient turbulence, wind-shear effects,
27      and cloud processes.  These dispersive mechanisms result from the interaction of large air
28      masses, or from the disturbance of the larger-scale flow in a given air mass by insolation-
29      driven surface fluxes of heat and moisture, and  by surface drag effects. The  influence of
30      these surface effects is largely confined to the atmospheric boundary layer (ABL),  the height
31      of which varies diuraally and seasonally,  peaking typically at between 1 and 3 km on

        April  1995                                3-88      DRAFT-DO NOT QUOTE OR CITE

-------
 1      summer afternoons over the continental U.S.A.  Pollutant emissions may be within the ABL
 2      or above it (depending on emission height and buoyancy), and their dispersion is markedly
 3      different in the two cases, being much more rapid and vigorous in the daytime convective
 4      boundary layer (CBL) than in the stable layers aloft or in the stable nocturnal boundary
 5      layer. Quantitative study of these transport and  dispersion processes requires, ideally,
 6      simultaneous measurements of a large number of variables related to insolation and clouds,
 7      surface characteristics and surface fluxes of heat and moisture, and dynamic 3-D fields of
 8      flow, temperature, humidity and concentrations of trace pollutants in the ambient atmosphere.
 9      Transport and dispersion processes  also have a critical influence on plume chemistry and dry
10      deposition, which are often diffusion-limited.  Meteorological measurements must therefore
11      be an integral part of any plume study,  even when the focus is on chemistry or deposition.
12      The shift to Eulerian grid modeling in the 1980s did not include measures to preserve the
13      essence  of the sub-grid-scale features of plumes, which  were instantaneously dispersed over
14      the entire emission grid cell (a volume of ~ 1012 m3 in  RADM with 80 km horizontal
15      resolution),  thereby also grossly distorting plume chemistry, aerosol formation, and pollutant
16      budgets.  There is growing awareness now of the need for more realistic treatment of plumes
17      in grid models.
18           A large body  of literature exists on studies (including field studies) of ABL structure
19      and  dynamics, and  on the characteristics of the wind, temperature and moisture fields in the
20      ABL and, to a lesser extent,  in the free troposphere aloft. Those studies are outside the
21      present scope.  Some of the recent  major advances in the knowledge about the ABL  are
22      reviewed by Briggs and  Binkowski (1985). Our scope here  is limited to field studies of the
23      transport and dispersion of PM and their precursors (e.g., SOX and NOX). Prior to 1975,
24      most such field studies were limited to the behavior of point-source plumes in the
25      7-mesoscale range (120 km), i.e.,  on plume rise and short-range dispersion. Such behavior
26      is well understood qualitatively; quantitatively, it is well enough represented in models at the
27      time scales characteristic of most commonly-used plume dispersion models (~ 1 h), but not
28      at the much shorter time scales of relevance to plume chemistry  and plume visibility. In this
29      near-source range, instantaneous plume behavior is very different from the larger scale
30      average  behavior. In an intercomparison of four plume visibility models, it was concluded
31      that  much of the  variation in visibility observed  in the Navajo power plant plume in northern

        April 1995                                3.39       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Arizona was probably due to fluctuations in source emissions and plume dispersion at scales
 2      below those resolvable by the models (White et al., 1985). Since the atmospheric residence
 3      of fine PM in the lower troposphere is on the order of days, our interest here is more on the
 4      transport and dispersion of plumes over the 0- and a- mesoscale ranges (« 20 to 200 and
 5      200 to 2,000 km).  Quantitative determination of transport over the mesoscale requires
 6      special field studies with controlled tracer releases.  Such  studies are relatively recent  and
 7      very few, and they  represent only a few isolated meteorological scenarios.
 8
 9      3.4.2.1.1  Field Measurements Related to Transport Modeling
10          Routine meteorological field measurements include surface weather observations of a
11      broad variety of meteorological variables made every three hours  at several thousand sites
12      across the  country by the National Weather Service,  as well as upper-air soundings
13      (radiosondes) of wind, temperature  and relative humidity twice a day (noon and midnight) at
14      a much more limited number of sites which, on average, are about 400 km apart. These
15      data constitute the principal raw meteorological information used in regional transport
16      models, which are either Lagrangian trajectory models or  dynamic 3D Eulerian grid models.
17      Most trajectory models are two-dimensional,  with atmospheric flow patterns being analyzed
18      on  isobaric or terrain-following surfaces, or in bulk transport layers confined to the mixed
19      boundary layer.   These simplifying assumptions concerning vertical motions lead to large
20      transport errors on  the regional scale (Kuo et al., 1985).   The vertical velocity can be
21      calculated  at grid points in a regional model domain  from  the continuity equation, but the
22      temporal and spatial resolutions of the  radiosonde data are so coarse in most areas that the
23      result would be a gross approximation only.  3D flows may be best simulated by moist
24      adiabatic trajectories, but since analysis methods cannot always  resolve the stratified nature
25      of the required moisture fields, the  most reasonable simulations of 3D transport are probably
26      dry adiabatic (isentropic) trajectories.  Danielsen (1961) presented a case study showing  a
27      separation  of =1,300 km after only 12 h  of transport as simulated by isobaric and isentropic
28      trajectories.  It was probably an extreme case.  The gridded wind field in regional Eulerian
29      air quality  models is typically generated by the application of dynamic 3D mesoscale
30      meteorological models (e.g., PSU-MM5 and  CSU-RAMS) which incorporate the routine
31      NWS observations through a dynamic  Four Dimensional Data Assimilation (FDDA)

        April 1995                               3-90       DRAFT-DO NOT QUOTE OR  CITE

-------
  1      technique.  The NWS surface weather database also includes a measure of prevailing
  2      visibility as determined by human observers.  A number of field studies have established the
  3      reliability of such subjective visibility observations (e.g., Horvath and Noll,  1969; Hoffman
  4      and Kuehnemann,  1979).  They have proved to be a very useful indicator of regional haze
  5      and its long-range  transport (Gillani and Husar, 1976), and have been used to study the long-
  6      term trends of the  spatial-temporal variability of regional haze and air quality in the eastern
  7      U.S. over many decades (Husar et al.,  1981; Sloane,  1982).
  8           Special field  studies of transport and dispersion are based on observations of the
  9      transport of pressurized  (constant density) balloons (called tetroons if their shape is
10      tetrahedral), and of the evolution of plumes resulting from pollutant emissions or controlled
11      releases of artificial tracers.  Balloons have been used in mesoscale studies in three ways: as
12      isolated Lagrangian markers of pollutant emissions (e.g., Clarke et al., 1983); in sequential
13      releases to provide one-particle diffusion estimates (e.g., Thomas and Vogt,  1990); and in
14      cluster releases to  study relative diffusion (e.g., Er-El and Peskin, 1981).  Tetroons generally
15      carry a transponder which permits continuous tracking with a radar, thus providing the
16      complete detailed 3D trajectory.  The range of the tetroon experiment is normally limited by
17      the tracking range  of the radar (< 100 km).  This range can be extended to the full range of
18      tetroon transport by including a tag which the finder can return with  information about at
19      least the terminal location.  In some studies (e.g., Clarke et al.,  1983), tetroons have been
20      tracked continuously over much longer ranges by sequential tracking  from the network of
21      FAA radars used in support of aviation. Studies based on tracers and air pollutants also
22      provide information about plume dispersion. Most early tracer studies were limited to  a
23      range of about 100 km due to the nature of the tracers then available and limitations of
24      technology.  Development of new tracers (e.g., the PFTs or perfluorocarbon tracers) and
25      new  sampling and  analysis techniques have not only extended the range in more recent
26      experiments by more than an order of magnitude, but  the new data are also more reliable.
27           Pack et al. (1978) presented a detailed review of many early studies  in which
28      observations of the transport of pollutant plumes,  tracers, or balloons were compared with
29      results of diagnostic trajectory calculations. The models commonly used then were based on
30      the kinematic approach (using objectively-analyzed wind fields based  on measured winds) and
31      a single transport layer.  The observed winds were used as input in different ways:  for

        April 1995                                3_91        DRAFT-DO  NOT QUOTE OR CITE

-------
 1     example, surface winds or adjusted surface winds representing average winds in the whole
 2     transport layer; or, upper air winds averaged over the transport layer.  The adjustment of
 3     surface winds included enhancement of the speed by as much as a factor of two, and a veer
 4     of the wind direction by as much as 40°, to account for the real-world wind speed shear and
 5     directional veer with height.  The advantage of using surface winds was due to their much
 6     higher spatial and temporal resolution, compared to the much coarser resolution of the upper-
 7     air radiosonde winds.  The early results of comparisons of calculated and observed
 8     trajectories evidenced a broad range of discrepancy (10 to 54% of the trajectory length after
 9     only  100 km, and 55 to 60% after 650 km), and also the presence of large systematic errors,
10     not always in the same direction, depending on the presence of complex flows due to fronts,
11     complex terrain, etc.  The best simulations were often obtained by the use of adjusted surface
12     winds, and such adjustments varied between studies.  The errors were found to be lowest for
13     transport in the daytime CBL, and substantially larger for transport in stably-stratified layers.
14           Moran (1992) has tabulated (his Table 2-4) basic information about a number of formal
15     /3- and a- mesoscale tracer experiments since 1973, in which the release was at surface level
16     and the measured transport range was at least 25 km (and up to 3,000 km). Table 3-16
17     summarizes, in chronological order, some of the major field studies of the past 20 years  with
18     measurements and modeling of transport extending into the a-mesoscale.  It includes the
19     major tracer studies as well as air quality and tetroon studies. The transport models in these
20     studies were driven either by routine meteorological observations or by additional
21     measurements made as part of the field studies.  The following important observations are
22     based on the studies listed in Table 3-13:
23
24         •  The routine data of the radiosonde network («400 km, 12 h) are too coarse both
25             spatially (Kahl and Samson, 1986, 1988) and temporally (Rolph and Draxler,  1990;
26             Kuo et al., 1985) for accurate simulation of long range transport.
27
28         •  The error in calculated trajectories is greatest under conditions of high speeds which
29             generally accompany complex mesoscale systems (Rolph and Draxler, 1990).
30
        April 1995                                3-92      DRAFT-DO NOT QUOTE OR CITE

-------
         TABLE 3-13. RECENT FIELD STUDIES OF a-MESOSCALE TRANSPORT AND TRAJECTORY MODEL
 VO
OJ
Study
INEL Study
Idaho Nat'l Eng. Lab

MISTT
Midwest Interstate
Sulfur Transport and
Transformation Study
VISTTA
Visibility Impairment
due to Sulfur Transport
and Transformation
in the Atmosphere
TPS
Tennessee Plume
Study



NEROS
Northeast Regional
Oxidant Study
Mt. Isa
Smelters Plume Study

Period
Feb-May 74

Summer 75
Summer 76


Jun, Jul,
Dec-79



Aug-78



15-Aug-78

Summer 79
Summer 80

Jul-79


Tracer(s)
Kr-85

Plume sulfur



Anthropogenic
aerosol, ozone



Tetroons (1 cu. m)
with transponder


"

Tetroons (6 cu. m)
Tetroons (1 and 6 cu. m)

Excess plume S
and Aitken Nuclei Count
(ANC)
Release Sites(s)
INEL (Idaho) fuel
reprocessing plant
(76 m stack)

Labadie Power
Plant near
St. Louis, MO

Los Angeles Basin




TVA
Cumberland
Steam Plant, TN

••

MD, OH, PA, TN
Columbus, OH

Mt. Isa, Australia
(Sulphide smelters,
-0.6 km apart)
Tracking/Sampling
Samplers at
1 1 Midwestern NWS
sites; 10-h day and
night samples
In-situ aircraft
measurements


Detailed air quality
and aerosol
measurements at a
Grand Canyon site

Radar to - 75 km;
terminal point based
on return tag.

«

Continued FAA
radars Radar and
return tag.
Aircraft measts. of
Total S, ANC,
COSPEC-SO2
Maximum Range
(Airshed)
-1,500 km

-300 km



-750 km




-1,000km
(KY,IN,OH,ONT)


~300km(KY)

-500 km to NE
-1,500 km to NE

-1,000km
(Semi-and region
in N. Australia)
Model
Comparison(s)
NOAA-ARL
trajectory
model with 300 m
vertical resolution
Simple particle
trajectory model


CAPITA
Monte Carlo particle
transport model


• NOAA-ATAD
• NCAR isentropic
• CAPITA Monte
Calo

3D reg'l dyn. model
NOAA, NCAR,
CAPITA, as above

Simple layered wind
trajectory model;

Ref(s)
Draxler(1982)

Gillani et al. (1978)
Gillani (1986)


Maciaset al. (1981)




Clarke et al. (1983)



Warner (1981)

Clarke et al. (1983)


Carras and
Williams (1981)

Comments
Small signal above b/g;
300 m layered approach
to permit spread by wind
dir'l shear necessary.
Quasi-Lagrangian pibal
measurements of winds
along plume transport

Evidence also of long-
range impact of Copper
smelter plumes


Part of a large plume
transport/ chemistry
study, including aircraft
measts.


Part of a large urban and
reg'l oxidant study

Exceptionally clean
plume b/g.

o
o

z
o
H

O
c
o
n
HH
H
m

-------
1
'O
VO Study
Great Plains
Mesoscale Tracer
Expt.

CAPTEX
Cross-Appalachian
Tracer Expt.

ANATEX
Across North
America
Tracer Expt.


MISERS GOLD


TABLE 3-13 (cont'd). RECENT FIELD STUDIES OF a-MESOSCALE TRANSPORT
AND TRAJECTORY MODELS
Period Tracer(s)
Jul-80 Two PFTs (PMCH
and PDCH) and two
heavy meyhanes
(ME-20, ME-21)
Sep/Oct83 PFT(PMCH)



Jan-Mar 87 3 PFTs (PMCP,
PMCH, PDCH)




l-Jun-89 Indium oxide
(vapor deposits on
particles)
Release Sites(s)
Norman, OK
(1 m AGL)


Dayton, OH
Sudbury, ONT


Glasgow, MT
St. Cloud, MN




White Sands
Missile Range,
NM
Maximum Range
Tracking/Sampling (Airshed)
Surface samplers: 17600kmtoNNE
on arc at 100 km 38
on arc at 600 km and
aircraft sampling
Surface array of — 1 , 100 km
>80 samplers at arcs(NE U.S.)
from 300-1, 100 km
and aircraft sampling
Surface network (77); -3,000 km
Towers (5); and (Eastern U.S.)
aircraft sampling.



In-situ aircraft: filter ~ 1 ,400 km
samples analysed for NM to MO
tracer and particles.
Model Comparison(s)
Different 3D regional
models


Different 3D regional
models; also
MESOPUFF II

3 single-layer LAGR,
6 multi-layer LAGR,
2 multi-layer Eulenan



Gifford's random-force
diffusion theory

Ref(s) Comments
Ferber et al. (1981) Important role of wind
Moran(1992) shear effects of nocturnal
jet.

Ferber et al. (1986) Terrain-effects found
Moran (1992) important. Enhanced
Godowitch (1989) upper air met measts.

Draxler et al. Enhanced upper air met
(1991) measts.
Rolph and Draxler
(1990)
Clark and Cohn
(1990)
Kahl et al. (1991) Dust plume from a
Mason and military test explosion.
Gifford (1992)
 •^

 %
9
O
d


I
O

-------
 1          •   Initial errors in trajectory simulations (both in direction and vertical spread) play a
 2              critical role in overall model uncertainty (Draxler et al., 1991).
 3
 4          •   Single-layer Lagrangian trajectory models do not spread the "plume" adequately,
 5              while Eulerian models spread it too much. Multi-layer Lagrangian models perform
 6              the best in terms of dispersion of point-source emissions (Clark and Cohn,  1990).
 7
 8          •   Vertical information about tracer trajectories, based on continuously-tracked tetroons
 9              and aircraft measurements, contain much useful information not captured by surface
10              sampling alone (Clarke et al., 1983).  There is, for example, evidence of cloud
11              venting of ABL pollutants  into the free  troposphere, where their residence time is
12              longer and the flow field may be quite different.
13
14          •   Terrain-induced effects played an important role in CAPTEX, and effects related to
15              the nocturnal jet were important in the Great Plains Experiment (Moran, 1992).
16              Nocturnal wind directional shear plays a major role in effectively dispersing plumes
17              which have been dispersed vertically during the preceding daytime CBL.
18
19          *   Directional wind shear plays an important role  in plume dispersion even in the CBL
20              during (3-mesoscale transport (Gillani, 1986).
21
22
23           The issue of substantial overdispersion by Eulerian models is important because the

24      state-of-the-art as well as the future direction in  mesoscale modeling (meteorological/air
25      quality/aerosol) appear to favor the Eulerian approach.  A significant source of the problem
26      must be related to the  gross initial overdispersion of plumes in regional Eulerian models,

27      particularly of elevated point-source plumes (carriers of most of the U.S. anthropogenic

28      emissions of sulfur).  The instantaneous false dilution of fresh emissions of NOX into the
29      NOx-limited surrounding environment (e.g., in the eastern U.S.) greatly distorts plume
30      chemistry and aerosol  formation. Proper sub-grid-scale treatment of plumes remains an
31      important outstanding issue in regional modeling.  Other sub-grid-scale effects in need of
32      more attention pertain  to complex mesoscale flows (e.g., storms, fronts, cloud venting,
33      complex terrain effects, etc.).  They too are an important source of model errors. A few

34      special field studies have been carried out to investigate such flows:  for example, VENTEX

35      (Ching and Alkezweeny, 1986) and PRESTORM (Dickerson et al., 1987)  for cloud venting,

36      and ASCOT (Allwine, 1993) and the  NGS Visibility Study  (Richards et al., 1991) for  flows
37      over complex terrain.  Thermal effects and drainage flows also evidently play an important
38      role in influencing paniculate air quality, as in the occurrence  of the Denver "brown cloud"
39      phenomenon (Sloane and Groblicki, 1981).

        April 1995                               3.95       DRAFT-DO NOT QUOTE  OR CITE

-------
 1           There is considerable field evidence also of synoptic scale transport (12,000 km) of
 2      airborne particles (see, for example, Gordon, 1991).  The impact of such transport is
 3      important on the global scale. That subject is beyond the present scope.
 4
 5      3.4.2.1.2  Field Measurements Related to  Dispersion Modeling
 6           Gaussian semi-empirical models have been the basis of most applied diffusion modeling
 7      since their development around 1960.  These models were based on Taylor's diffusion theory
 8      of stationary homogeneous turbulence (Taylor, 1921), and were built on a few field
 9      experiments that were quite limited in scope and technology.  The results have been
10      extrapolated far beyond the intended range  of downwind  distance and ambient conditions.
11      Some of the extrapolations were guided by statistical theory, but most were freehand
12      extrapolations (Briggs and Binkowski,  1985).  Many research-grade field studies of
13      atmospheric dispersion have since been performed, but most have been limited to the
14      -y-mesoscale range.  These have been reviewed by Draxler (1984), Irwin (1983), Briggs  and
15      Binkowski (1985) and others.  /3- and a-mesoscale studies, based on observations of the
16      dispersion of pollutant and tracer plumes have been reviewed by Moran (1992).  Pollutant
17      plumes remain vertically narrow in stable flows (e.g., elevated power plant plumes released
18      at night), but rapidly fill up the CBL after fumigation in the  daytime (see, for example,
19      Gillani et al., 1984).  Information about spreads of plumes in the elevated stable layers is
20      particularly limited.  The most common basis for estimation  of such spreads (expressed as
21      ay and  az, the RMS  variances of lateral and vertical plume spreads) over distances under
22      100 km or so is the well-known Pasquill-Gifford (P-G) curves for different  stability classes
23      (Gifford, 1961), which make use of the routine meteorological measurements to determine
24      applicable stability class.  The P-G curves were developed mostly from data collected within
25      the mixing layer.  Another set of parameterizations of elevated plume spreads was developed
26      by TVA (Carpenter et al.,  1971) based on twenty years of experience  in plume observations
27      and aerial monitoring. These require the temperature profile to establish atmospheric
28      stability.  More recently, Smith (1981) analysed aircraft measurements  in elevated power
29      plant plumes in different parts of the U.S., mostly in the stable layers,  and  determined that
30      the P-G curves overestimated plume spread in stable layers quite substantially both vertically
31      and horizontally, and that the TVA approach tended  to underestimate the horizontal spread,

        April 1995                               3-96      DRAFT-DO NOT QUOTE OR  CITE

-------
  1      but possibly overestimate the vertical spread.  Evidently, there continues to be uncertainty
  2      about plume spreads even at distances under 100 km.  Of particular interest is horizontal
  3      plume dispersion, both because it is generally far greater over the mesoscale, and because it
  4      is highly variable.  Close to the source, plume spread is largely by progressively larger
  5      turbulent eddies, but after the plume dimension substantially exceeds the scale of these eddies
  6      (typically less than 1 km), dispersion is increasingly by directional wind shear with height.
  7      Such shear is small  for the  vertically thin nocturnal plume, moderate for the plume in the
  8      CBL, but maximum for the daytime plume which, after maximum vertical spread in the
  9      CBL, enters the nocturnal regime which is often characterized by strong directional shear
10      effects (Gillani et al.,  1984).  The average crosswind spread rates of plumes from  a large tall
11      stack power plant emitted within the CBL on summer days in the Midwest were observed to
12      be in the range 0.25 to 1.0 km per km of downwind transport until the plume attained a
13      width of about 30 km. Thereafter, further plume spread within the CBL was typically much
14      slower (Gillani and Pleim, 1994).
15           A common approach in Lagrangian studies of dispersion over long distances has been
16      to use semi-empirical  "mesoscale" dispersion coefficients by analogy with parameterizations
17      of microscale turbulent spread.  An important consequence of Taylor's statistical theory was
18      that,  in stationary homogeneous turbulence, sy  grew linearly with time at first (for t ~ TL,
19      the Lagrangian time scale,  = 1 to 2 min in the CBL), and then asymptotically as t1/2 within a
20      few kilometers.   Observations  of a few a-mesoscale field studies  have been interpreted to
21      suggest that the  regime of linear time dependence may apply also at long distances (see, for
22      example, Pack et al.,  1978), with the characteristic time scale (TL) here being related to the
23      diurnal and/or inertial  scale («24 h).  Others have proposed parameterizations of mesoscale
24      ay which use powers of t ranging  from 0.85  to 1.5  (see, for example, Carras and Williams,
25      1988). Thus, there  is  no consensus about simplistic modeling of  mesoscale diffusion over
26      scales exceeding 24  h.  Given the wide range of conditions that plumes can experience
27      during long range transport in  different air masses,  over a variety of terrain types, and over
28      multiple diurnal cycles during different seasons, such a controversy is not surprising.  For
29      transport in the first 24 h, the time and height of emission are critical influencing variables.
30      Thus, for example, crosswind spreads after 24  h of transport of two plumes released from
31      the same tall-stack power plant at  0800 and 2000 on a given day are likely to be very

        April 1995                                3.97       DRAFT-DO NOT QUOTE OR CITE

-------
 1      different. During the next diurnal cycle, however, these spreads, as a fraction of downwind
 2      distance travelled, are likely to converge.  Alternate semi-empirical approaches of
 3      representing mesoscale plume dispersion include simulation of relative dispersion of
 4      hypothetical co-emitted conservative particles. In conjunction with instantaneous wind data
 5      (e.g., pibal  soundings), such models have  proved to be satisfactory over /3-mesoscale
 6      distances (e.g., Gillani, 1986), but more work is needed to establish their application over
 7      long distances based on hourly-average gridded wind data such as are produced by the
 8      meteorological preprocessors of regional Eulerian models.  Overall, based on field evidence,
 9      paniculate air quality is significantly influenced by regional transport and dispersion, but
10      quantitative simulation of these processes is still subject to considerable error.
11
12      3.4.3  Transformations  in Plumes
13      3.4.3.1  Gas-to-Particle Conversion in Plumes
14          A number of field studies of gas-to-particle conversion have been conducted in the
15      plumes of large point-sources of SOX and NOX (e.g., coal- and  oil-fired power plants and
16      metal smelters).  Fewer studies have focused on urban-industrial plumes. These studies have
17      focused principally on quantifying the rates of aerosol formation and, to a lesser extent, on
18      investigating the mechanisms.  Mechanistic studies are more difficult, particularly when
19      multiple mechanisms are co-active,  as is commonly the case.  In the NAPAP emissions
20      inventory for base year 1985 (Placet et al., 1991), about 70% of the U.S. anthropogenic
21      emissions of SO2, and about 25% of the corresponding emissions of NOX, were attributed to
22      large point-sources with stack heights exceeding  120 m (probably less than  150 individual
23      sources). The contribution of such sources is even higher in  the eastern U.S., particularly in
24      the Ohio and Tennessee River Valleys.  Clearly, these large emissions  are very important in
25      the context  of regional aerosols.  Fortunately, many of these  sources are located in rural
26      areas,  and their plume chemistry can be studied in isolation from the complications of
27      interactions with other plumes.  Much of the remaining anthropogenic emissions of SOX and
28      NOX are contributed by urban-industrial area sources.
29
30
        April 1995                                3-98       DRAFT-DO NOT QUOTE OR CITE

-------
  1      3.4.3.1.1   Plume Studies
  2           Power plant and urban plume studies of SO2-to-sulfate transformations published before
  3      1980 have been reviewed by Newman (1981) and in the earlier 1982 PM/SOX Air Criteria
  4      Document (U.S. Environmental Protection Agency, 1982).  Only a brief overview of those
  5      studies is provided here; the main focus here is on plume  studies published after 1980.  Since
  6      the plume mass is airborne, the most meaningful plume studies are based on measurements
  7      made from instrumented aircraft.  Early studies (pre-1975) often reported SO2 oxidation rates
  8      as high as 50% h"1.  They  are now generally considered to be flawed due to limitations in the
  9      measurement technology then available. This technology  has made major strides since. For
 10      example, the development of the filter pack (Forrest and Newman, 1973) has proved to be a
 11      useful method of simultaneous collection of high-volume samples of SO^ and paniculate
 12      sulfur.  Such samples,  however, only provide average concentrations over entire plume
 13      cross-sections or, at best, over long crosswind plume traverses.  The development of
 14      continuous monitors for both SO2 and paniculate sulfur (Huntzicker et al., 1978; Cobourn
 15      et al., 1978) made it possible to study sulfate formation with crosswind plume detail. Such
 16      detail during a single plume traverse contains a nearly instantaneous snapshot of the  full
 17      spectrum of chemistry  between the high-NOx regime in plume core to the low-NOx regime at
 18      plume edge (Gillani and Wilson, 1980). With cross-sectionally averaged measurements, such
 19      a spectrum can only be discerned in measurements ranging from near-source to far
 20      downwind.  The technology of continuous measurements of nitrogen  species  with high
 21      sensitivity has also evolved greatly since 1980.
 22          The period between 1974 and 1981 was very active in terms of plume studies focused
 23      particularly on estimating the rate of oxidation of SO2. Studies by Brookhaven National
 24      Laboratory (Newman et al., 1975a,b; Forrest and Newman, 1977a,b) and TVA (Meagher
 25      et al.,  1978) in coal- and oil-fired power plant  plumes as well as a nickel smelter plume
 26      generally yielded low oxidation of SO2 (seldom exceeding 5%  over 50 km and several hours
27      of plume transport, with  an uncertainty of about a factor of two). These investigators found
28      the oxidation rate to be highest close to the source, where  it appeared to be correlated with
29      plume paniculate loading, and  interpreted the oxidation to  be due to a heterogeneous second-
30      order mechanism which became quenched as the plume diluted (Schwartz and Newman,
31      1978).  These results were  in sharp contrast to those of Husar et al. (1976) for a coal-fired

        April 1995                               3.99      DRAFT-DO NOT QUOTE OR CITE

-------
 1      power plant plume, also over about 50 km of plume transport, which showed the oxidation
 2      rate to be slow during an early induction period, increasing thereafter to as much as 5 % h'1.
 3      No mechanistic interpretation was proposed by these authors.  This controversy was resolved
 4      by the subsequent findings of Gillani et al. (1978) resulting from two case studies which
 5      were remarkable for their coverage of downwind range exceeding  300 km and 10 to  12 h of
 6      transport of a coal-fired power plant plume during daylight as well as dark.  The authors
 7      found the oxidation rate of SO2 to be strongly correlated with sunlight, and also  with the
 8      extent of plume dilution, and background ozone concentration (considered to be a surrogate
 9      for background reactivity).  Maximum measured paniculate sulfur as a fraction of total
10      plume sulfur ranged as high as 18%.  The daytime conversion rate in the plume was slow at
11      first, but increased as the plume diluted,  reaching maximum values on the two days of
12      1.8 and 3.0% h"1 in the afternoon.  Such rates are consistent with  theoretical rates based  on
13      the SO2-OH reaction (Calvert et al., 1978; Hov and Isaksen, 1981).  The entire  plume
14      transport on both occasions was in fairly dry environment (R.H. <  70%). Presumably,  the
15      mixing of plume NOX and background VOC led to photochemistry which generated the
16      necessary oxidants for gas-phase  oxidation of SO2. The measurements of VOC in the
17      background were both sparse and of limited reliability. The study also found the formation
18     of substantial excess  of ozone in  aged plumes.  The interpretation  based on plume-
19     background interaction satisfactorily explained the results of the BNL and TV A studies in
20     which the measurements of low oxidation of SO2 were all in coherent stable elevated plumes
21      during early morning and evening hours  (low sunlight and little plume dilution), as well as of
22     Husar et al., whose measurements were  in the more polluted and  convective summer daytime
23     CBL.
24          As of the end of the 1970s, a number of factors had been  implicated as being  relevant
25     to plume sulfur chemistry.  Gillani and Wilson (1980) conducted a systematic investigation of
26     the dependence of ozone and aerosol formation in power plant plumes on a variety of
27     possible influencing factors, based on the plume data of five case  studies.  They found that
28     temperature variations in the range 28 to 33  °C, and R.H. variations in the range 50 to 80%
29     did not have an appreciable influence; the importance of sunlight, plume dilution and
30     background composition was reconfirmed.  Eatough et al. (1981,  1982) have observed a
        April 1995                               3-100      DRAFT-DO NOT QUOTE OR CITE

-------
  1     positive temperature dependence of a linear SO2 oxidation rate in power plant and smelter
  2     plumes in western U.S. in the temperature range 0 to 30 °C.
  3          Gillani and Wilson (1980) also presented direct evidence and interpretation of the role
  4     of plume-background interactions in plume photochemistry within the context of a common
  5     pattern of diffusion-limited plume chemical evolution through three stages in a moderately
  6     polluted environment.  In the "early" stage, the plume is narrow and dominated by a high-
  7     NOX regime in which ozone  and other oxidants are sharply depleted by reaction with plume
  8     NO and SO2; the VOC-NOX  chemistry, SO2 oxidation, and aerosol formation are inhibited in
  9     the plume in this stage.  As the plume spreads and dilutes with a background characterized
 10     by relatively high VOC/NOX ratio, the VOC/NOX ratio increases also in plume edges.  This
 11     "intermediate" stage of plume chemistry is characterized by rapid formation of ozone and
 12     aerosols in plume edges, leading to an observed  excess there of ozone over the background
 13     (ozone "wings")  while the plume core still has an ozone deficit.  Sharp "wings" of Aitken
 14     nuclei concentration have also been observed in plume edges at times,  indicating directly the
 15     nucleation of new aerosol (Wilson, 1978;  Gillani et al., 1981).  With continuing dilution, the
 16     plume ultimately develops a condition of low-NOx, high VOC/NOX ratio and, in the summer,
 17     an ozone "bulge" throughout. In this "mature" stage,  the rate of oxidation of SO2 to sulfates
 18     (and presumably  also of NOX to secondary products) reaches  its  peak.
 19          Gillani et al. (1981) provided a quantitative interpretation of the above  observations by
 20     developing an empirical parameterization of the gas-phase conversion rate of SO2 to sulfate
 21      in terms of measured variables representing sunlight, mixing  and background reactivity.   The
 22     parameterization was verified based on the "dry" data  of three different power plant plumes
 23      over ten days of measurements  in two different summer periods.  Crosswind-resolved
 24      reactive plume models capable of facilitating plume-background interactions and including
 25      detailed simulation of chemical kinetics have been developed  and applied by  Hov and Isaksen
 26      (1981), Stewart and Liu (1981),  and Gillani (1986).  These models can depict the  observed
 27      behavior of ozone in the three plume stages.  Their applications  have shown that the
28      evolution of OH in the plume (a measure of oxidation  potential)  mimics the above description
29      of ozone evolution (Hov and Isaksen,  1981), and that plume oxidant  and aerosol formation
30      are very sensitive to background VOC and their ingestion into the plume (Gillani, 1986).
31      However, these models continue to remain unevaluated adequately owing to a continuing lack

        April  1995                               3_10i     DRAFT-DO NOT QUOTE OR CITE

-------
 1     of data characterizing the composition of plume background (especially VOC) and the
 2     crosswind detail of important intermediate and secondary species (e.g., OH, HO2, HNO3,
 3     etc.).
 4           A number of plume studies have verified the sunlight dependence of the SO2 oxidation
 5     process,  observing higher seasonal conversion rates during summer, and higher diurnal rates
 6     during midday (Husar et al., 1978; Lusis et al., 1978; Roberts and Williams,  1979; Meagher
 7     et al., 1981; Hegg and Hobbs, 1980; Gillani et al., 1981; Forrest et al., 1981; Williams
 8     et al., 1981; Wilson,  1981; Wilson and McMurry, 1981; Liebsch and de Pena, 1982).
 9     In these  studies, the peak daytime conversion rate was typically between 1 and 5% h"1 in the
10     summer  (higher under humid conditions), and much lower in winter.  Wilson (1981)
11     reviewed the data of twelve power plant and smelter plumes in the  U.S., Canada and
12     Australia, covering measurements during day and night, and summer and winter. The main
13     conclusion was that diurnally, midday conversion rates were relatively high and  quite
14     variable  (1 to  10% h"1), while the nighttime conversion rates were generally low (under
15     0.5% h"1).  Also, the rates were found to be lower in winter than in summer.
16     Geographically, the measured plume conversion rates in the arid and relatively clean
17     southwestern U.S. environment were found to be particularly low (0.5% h"1) at all times,
18     including summer midday.  Williams et  al. (1981) also found the rates to be low in a smelter
19     plume in the arid,  clean environment of north central Australia ( = 0.15% h"1 averaged over
20     24 h of transport).
21           Gillani et al.  (1981) were able to formulate the parameterization of the gas-phase
22     conversion rate by isolating case studies performed entirely in dry conditions when liquid-
23     phase contributions were negligible.  They also observed that for all cases when the plume
24     had  any  history of wet exposure  (clouds, fogs or high humidity), the oxidation of SO2
25     invariably proceeded at a rate faster than that predicted by the gas-phase parameterization.
26     Whereas the typical range of the peak summer daytime conversion  rate was  1 to 5 % h"1 in
27     Project MISTT (Missouri, Illinois), it was closer to 1 to 10% h"1 in the more humid
28     conditions of the Tennessee Plume Study (Tennessee, Kentucky).  In the wetter daytime
29     situations, evidently, liquid-phase chemistry was  superposed over the underlying gas-phase
30     chemistry.  Gillani and Wilson (1983) focused their study on the plume data of such "wet"
31     situations.  They attributed to liquid-phase chemistry the part of the total measured

       April 1995                               3-102      DRAFT-DO NOT QUOTE OR CITE

-------
  1      conversion rate which was in excess of the rate estimated by the gas-phase parameterization.
  2      The liquid phase was found to be due to clouds, fogs and light rain, or due to wetted
  3      aerosols under conditions of high ambient humidity (R.H.  > 75%).  The liquid-phase
  4      contribution to the conversion rate was found to be in excess of 40% of the total in two-
  5      thirds of the cases analysed, being as high as 8% hf1  averaged over the whole plume over
  6      6 h of transport in the most extreme case (clouds and light rain).  Similar increases in
  7      conversion rates in power plant plumes interacting  with high humidity have also been
  8      observed by others (e.g., Dittenhofer and de Pena, 1978; Eatough et al., 1984).
  9           Determination of the liquid-phase conversion  rate involves quantification not only of the
10      kinetics, but also of the  discrete  and variable extent of plume-cloud interaction.  Gillani et al.
11      (1983) formulated a parameterization of the conversion rate for plume-cloud interaction in
12      which the physical extent of such interaction was represented probabilistically, and the higher
13      liquid-phase conversion  rate  was applied only for the in-cloud portion of the plume. The
14      application of the parameterization to a case study  corresponding to summer daytime plume
15      transport within the CBL,  in patchy contact with fair-weather cumulus above, permitted
16      estimation of the average in-cloud conversion rate averaged over 7 h (1000 to 1700) to be
17      12% h"1.  Considering that the corresponding average liquid water content in the clouds was
18      certainly less than 1 g m"3 (1  ppm), much higher actual oxidation rates within individual
19      droplets are indicated.  Gas-phase photochemistry at a much slower rate was concurrently
20      quite active in the more  extensive drier parts of the plume below, producing ozone and other
21      oxidants which contributed to gas-phase as  well as  liquid-phase sulfur chemistry.  It was not
22      possible to relate the in-cloud kinetic rate to the critical variables controlling it, such as cloud
23      liquid water content, H2O2 concentration, or droplet pH, because such measurements were
24      not made.  The role of concurrent gas-phase photochemistry is indeed essential to provide the
25      oxidizing agents of liquid-phase chemistry.  Clark et al. (1984) found the contribution of
26      liquid-phase chemistry in a power plant plume to be negligible during long-range transport
27      over water in a shallow  stratocumulus-filled boundary layer, with limited plume dilution,  low
28      insolation, and little photochemistry.
29           A quite different approach based on aerosol growth laws applied to aerosol size
30      distribution data was taken by McMurry et al. (1981) and McMurry and Wilson (1982) to
31      study relative contributions of the principal mechanisms of gas-to-particle conversion.

        April 1995                                3_103      DRAFT-DO NOT  QUOTE OR CITE

-------
 1     Theory predicts different growth laws for different chemical mechanisms of aerosol
 2     formation.  The authors examined the functional dependence of calculated particle diameter
 3     growth rate on particle diameter.  By matching field data with theoretical growth laws, it was
 4     possible to differentiate between mechanisms.  Application of this approach indicated gas-
 5     phase chemistry and condensation of the product to be the predominant mechanism of aerosol
 6     formation in several power plant plumes in eastern and western U.S., with increasing
 7     contribution of heterogeneous mechanisms with increasing humidity (McMurry et al.,  1981);
 8     in a case study of the urban plume of St. Louis, 75% and 25% of the aerosol formation were
 9     attributed to homogeneous and heterogeneous mechanisms, respectively, while most of the
10     aerosol formation in the ambient air in the Great Smokey Mountains where relative
11     humidities were high (up to 95%) was attributed to the droplet-phase mechanism (McMurry
12     and Wilson, 1982).
13          In an overview of empirical parameterizations of sulfur transformations in power plant
14     plumes, Gillani (1985)  estimated that on a 24-h average basis, sulfate formation rates in a
15     large power plant plume in the U.S. Midwest in July 1976 were likely to be 0.8 ± 0.3% h'1
16     by gas-phase  reactions (midday peak ~ 2.6%  h"1) and at least half as much by liquid-phase
17     reactions. Winter  rates were estimated to be an order of magnitude lower than the summer
18     rates for the gas-phase  mechanism, but comparable for the liquid-phase mechanism. Since
19     1981, no new field studies of chemistry in large point-source plumes have been conducted in
20     the eastern U.S.  A comprehensive plume study with state-of-the-art aircraft measurements of
21     primary and secondary sulfur and nitrogen species, as well as  VOC and ozone, is planned to
22     occur in the summer of 1995 as part of the Southern Oxidant Study (SOS) Nashville Field
23     Measurement Program.
24          Information about field measurements of nitrate formation in point-source combustion
25     plumes is much more meager.  Summertime plume measurements suggest that nitrate
26     formation is principally in the form of nitric acid vapor (Hegg  and Hobbs,  1979; Richards
27     et al.  1981), and that oxidation of NOX to HNO3 may proceed  about three times faster than
28     the rate of oxidation of SO2 (Richards et al., 1981; Forrest et al., 1981).  Richards et al.
29     (1981) observed that along the transport  of the Navajo Generating Station in Arizona,  there
30     was adequate ammonia to  neutralize the sulfate formed in the plume, but not enough to form
31     ammonium nitrate.  Forrest et al.  (1981) found NH^/SO^f to increase with downwind

       April 1995                               3-104     DRAFT-DO NOT QUOTE OR CITE

-------
  1      distance and was mostly less than 2 (not enough to fully neutralize the sulfate), but
  2      sometimes more than 2, indicating a possibility of the formation of some ammonium nitrate.
  3      Eatough et al. (1981) observed that in the western desert region, the neutralization of sulfuric
  4      acid in plumes was due not only to ammonia, but also to other basic material (e.g., metal
  5      oxides and CaCO3).
  6           Field information about secondary formations in urban plumes is scantier than for
  7      power plant plumes for sulfur  compounds, but possibly slightly more for nitrogen
  8      compounds.  White et al. (1976, 1983) reported slow formation of ozone and aerosols at first
  9      in the St. Louis urban plume,  but faster rates  farther downwind.  Average  sulfate formation
 10      rates between successive downwind measurement locations on summer days were estimated
 11      at 2 to 4% h"1.  Isaksen et al.  (1978) applied a reactive plume model to  a subset of the
 12      St. Louis data, and estimated peak rates for the formation of sulfuric and nitric acid of 5 and
 13      20%  h"1, respectively. Based  on the same data set, Whitby (1980) estimated that about
 14      1,000 tons of secondary fine aerosol may be produced in the plume in one summer
 15      irradiation day.  Alkezweeny and Powell (1977) estimated peak sulfate formation rates in the
 16      St. Louis plume at  10 to 14%  h"1.  Miller and Alkezweeny (1980) reported sulfate formation
 17      rates in the Milwaukee urban plume on two summer days  in very different air masses to
 18      range from 1% h"1  (clean background) to 11% h"1 (polluted background).  The most
 19      extensive studies of NOX chemistry in urban plumes have been reported by Spicer and
 20      co-workers.  They have reported results for the Los Angeles, Phoenix, Boston and
 21      Philadelphia urban plumes.   In the Los Angeles studies, the transformation rate of NO2-to-
 22      products was estimated at 5 to 15% h"1  (Spicer,  1977) and 5 to 10% h:1  (Spicer et al., 1979).
 23      The sum of transformation plus removal rates  was estimated for the Phoenix and  Boston
 24      plumes at <5%  h"1 and 14 to  24% h"1, respectively.  The low rate for Phoenix was
 25      attributed partly to thermal decomposition of PAN after its formation in the plume.  In a
 26      study of the Detroit plume, Kelly (1987) estimated the NOX transformation rate at 10% h"1,
 27      with 67 to 84% of the products being in the form of HNO3. Measured concentrations of
28      nitric acid, however, were much lower because of its higher removal rate.  All of the above
29      urban plume  studies, and most of the power plant plume studies, have been daytime studies.
30      Field measurements of nighttime chemistry of nitrogen oxides in plumes are almost non-
31      existent.

        April 1995                               3_105      DRAFT-DO NOT  QUOTE OR CITE

-------
 1     3,4.3.1.2 Background Field Studies
 2          Attention is now focused on studies of aerosol formation in background air. The plume
 3     studies have shown that the rates of oxidation of SO2 and NOX in the background represent
 4     approximately the upper limit of the conversion rates in the plume.  In non-humid,
 5     moderately polluted conditions, they range typically between 1  and 5% h"1 for midday SO2
 6     oxidation in summer in the eastern U.S. (depending on the variability of VOC/NOX and the
 7     composition of VOC),  and up to 1 %  h"1 in the cleaner parts of the Western  U.S. Winter
 8     rates are about an order of magnitude lower.  By contrast, observed  NOX to nitrate
 9     conversion rates are about three times faster in summer than in winter (Parrish et al., 1986).
10     Aerosol nitrate formation depends strongly on availability of NH3  and on temperature.
11     Background aerosol is generally more aged and its  acidity more neutralized than plume
12     aerosol.
13          The situation is more complex in humid conditions.  Field measurements of the
14     compositions of cloudwater, rainwater and the precursor clear-air aerosol have shown that
15     strong acidity is substantially greater  in cloud and rain water than  in the clear-air aerosol
16     (Daum et al., 1984; Lazrus et al., 1983; Weathers  et al.,  1988).  This is indicative  of the
17     contribution of aqueous-phase chemistry to  cloudwater acidity in excess  of that due  to
18     scavenged aerosol.  Based on climatological data of clouds and SO2  distribution, and
19     assuming aqueous-phase oxidation of SO2 by ozone, Hegg (1985)  estimated  contribution of
20     the aqueous mechanism to global tropospheric sulfate production to be at least 10 to 15 times
21     greater than that due to the gas-phase mechanisms.  Applications of  more comprehensive
22     global models have given  estimated aqueous-phase  contributions of 40 to 95 % of the total
23     sulfate production (Langner and Rodhe, 1991 and references therein).  Regional models for
24     North America suggest 50 to 80% of the sulfate deposited in precipitation to be  formed in
25     clouds (Fung et al., 1991; McHenry  and Dennis, 1991).
26          A number of ambient studies have attempted  to study aqueous  chemistry based on in
27     situ measurements in clouds. Determination of the rates and mechanisms of aqueous-phase
28     chemistry is particularly ambiguous for several  reasons.  First, it is  difficult to distinguish
29     between the contributions  of in situ chemistry and  aerosol scavenging to the observed
30     concentration of the solute in the droplet phase.  Also, aqueous chemistry rate depends not
31     only on the change in concentration,  but also on the change in time.  It is difficult enough to

       April 1995                               3-106      DRAFT-DO NOT QUOTE OR CITE

-------
  1     determine the difference in concentration of even one reactant or product species, but
  2     determining the corresponding time difference is even more difficult (Schwartz, 1987; Gervat
  3     et al.,  1988;  Kelly et al.,  1989).  In stratiform clouds,  in particular, it is not always possible
  4     to determine  what constitutes pre-cloud air corresponding to specific cloud water samples
  5     (Gillani et al., 1994).  Finally,  it is difficult,  based on  field data, to attribute the inferred
  6     chemistry to  specific mechanisms (oxidation by H2O2 or O3, etc.).  The conclusions
  7     regarding rates and mechanisms of aqueous chemistry based on measurements in clouds are
  8     therefore quite uncertain, and have been a source of considerable controversy (e.g., Hegg
  9     and Hobbs, 1982, 1983a,b versus Schwartz and Newman, 1983).  One important finding in
 10     support of in-cloud  oxidation of SO2 by H2O2 , however, is the almost universal mutual
 11     exclusion of these two  species in non-precipitating stratiform clouds (Daum et al., 1984;
 12     Daum, 1988). In such clouds, there is generally enough time available for the species to
 13     react fully until the  one with the lower concentration  in the precursor air is depleted. The
 14     implication is that the aqueous-phase oxidation of SO2 by H2O2 takes precedence over other
 15     competing reactions.
 16          Most field studies have been limited to estimating  the amount or fraction of sulfate
 17     formed by the aqueous pathway, rather than the rate of formation.  Liu et al. (1993) have
 18     summarized the results of a number of cloud  studies between 1979  and 1991.  In these
 19     studies, a number of different approaches have been used to resolve the contributions of
 20     aerosol scavenging and in  situ chemistry to the observed cloudwater sulfate.  The study of
 21      Liu et al. (1993), which was part of the first intensive (summer 1988) of the Eulerian Model
 22     Evaluation Field Study  (EMEFS), used three different approaches for estimating the
 23      scavenged fraction of observed sulfate, and attributed 27 to 55% of cloudwater sulfate to in
 24     situ production.  The inferred results for the aqueous-phase production of sulfate in the
 25      collective studies  vary widely.  In winter studies, such production is low (e.g.,  Strapp et al.,
 26      1988),  while  in summer studies,  it is generally higher (e.g., Mohnen and Kedlacek,  1989).
 27      Many studies  implicate  H2O2 as the principal  oxidant (e.g., Van Valin et al., 1990), while
28      others implicate ozone (e.g., Hegg and Hobbs, 1986).
29           There is a variety of evidence for and against the  formation of HNO3 in the cloud
30      environment (e.g., Lazrus  et al., 1983; Daum et al., 1984; Hegg and Hobbs, 1986; Leaitch
31      et al., 1986a). The  heterogeneous mechanism involving N2O5 has received attention mostly

        April 1995                                3_107      DRAFT-DO NOT QUOTE OR CITE

-------
 1     as the "nighttime" mechanism (Lazrus et al., 1983; Richards, 1983) owing to the short life of
 2     the NO3 radical (precursor of N2O5) in sunlight. To account for the comparable measured
 3     amounts of sulfate and nitrate deposited  in winter storms in Ontario,  Barrie (1985) suggested
 4     the possibility of the N2O5 mechanism for wintertime formation of nitrate in clouds.  Leaitch
 5     et al. (1988) found substantial enhancement of NO"3 in and near clouds on 8 of 12 days of
 6     winter measurements in central  Ontario under freezing conditions and low insolation.
 7     On these occasions, variations in NOySO^" were associated with H+/SO4~ in the cloud water,
 8     implicating HNO3. Also, the observed levels of NO"3 could not be simulated in a model
 9     without invoking the N2O5 mechanism.  Based on a detailed examination of the nighttime
10     behavior of the NO3 radical,  Noxon (1983) concluded that  there was a significant loss of
11     NO3 compared to  N2O5 by an unknown scavenger (wet particles?).  In measurements at a
12     rural  site in central Ontario in August 1988 as part of EMEFS, Li et al. (1993) observed a
13     gradual increase in the concentration of aerosol nitrate (NO3~) from 1800 to midnight, and
14     then a gradual decrease.  In a diagnostic model study, they concluded that the observations
15     could be explained by heterogeneous reactions of NO3 and  N2O5 on wet particles.  They
16     attributed more than 80% of the NO3" formation to NO3 and about 10%  to N2O5 , and less
17     than 5% to HNO3.
18           In visibility studies, the water content of aerosols is of crucial importance.  The
19     estimation  of visibility impairment involves use of models in conjunction with ambient data
20     of both aerosols and relative humidity. Frequently, both sets of data are not available
21     concurrently for all stations in a monitoring network such as IMPROVE (Interagency
22     Monitoring of PROtected Visual Environments).  In such cases, gaps in information must be
23     filled by the  use of empirical relationships between average visibility impairment caused by
24     soluble aerosols and  average relative humidity derived from the available concurrent  data.
25     Such an application based on data at the 36 national IMPROVE sites is described by  Sisler
26     and Malm  (1994).
27           Another important area  which critically involves water uptake by soluble aerosols
28     relates to aerosol-cloud interactions.  Such interactions are  a critical link in cloud formation
29     and the global water cycle, in cloud optics and the global energy budget, in pollutant
30     redistribution by clouds,  in pollutant wet removal from the atmosphere,  and in atmospheric
31     chemistry.   Of particular importance is the process of aerosol scavenging by clouds.

       April 1995                              3-108      DRAFT-DO NOT QUOTE OR CITE

-------
 1      Interstitial aerosols in clouds may become incorporated into cloud droplets by "activation"
 2      (droplet nucleation), Brownian diffusion, inertial impaction, coalescence, and phoretic
 3      effects.  Of these microphysical cloud processes, aerosol activation is by far the most
 4      important. A soluble particle (the CCN) is activated when water vapor supersaturation
 5      around it (S) exceeds a critical value  (Sc) which depends principally on particle dry size (D0)
 6      and composition (commonly expressed in terms of the water-soluble solute fraction, s).  The
 7      works of Kohler (1936), Junge and McLaren (1971) and Hanel (1976)  provide the underlying
 8      theory for condensation of water on aerosols based  on assumptions of internally mixed
 9      aerosols.  Based on properties of representative continental and marine CCN, Junge and
10      McLaren predicted that Sc would be sensitive to CCN size, but to CCN composition only for
11      10.1.  Fitzgerald  (1973) confirmed the insensitivity to e in the range 0.15 to 0.35 based on
12      simultaneous measurements of CCN size,  e and CCN activation spectra (functional
13      dependence of activated fraction of aerosol on S) for S between 0.35 and 0.75%.
14           More recently, based on extensive year-long measurements of CCN spectra for
15      continental aerosols (representative of eastern U.S.  background), separated into narrow size
16      bands within the accumulation mode, Alofs et al. (1989) derived a simple semi-empirical
17      expression relating Sc to D0 and e applicable down  to S = 0.014%.  They also showed,
18      based on their own data and a literature review, that for continental aerosols in industrialized
19      regions, e »  0.5 is a reasonable approximation, indicating that the activation of such
20      aerosols is unlikely to be sensitive to  particle composition.  Based on their expression for Sc
21      and using s = 0.5, a supersaturation  of about 0.1% (characteristic for  stratiform clouds)
22      would be adequate to activate most of the accumulation mode particles  exposed to a cloud.
23      Cumuliform clouds with higher S would activate many Aitken mode particles also.  In
24      cumulus clouds, peak supersaturation is typically attained near cloud base, which is where
25      maximum activation is likely to occur.  The cloud module of the Regional Acid Deposition
26      Model (RADM) is based primarily  on a cumulus parameterization, and makes the assumption
27      of 100%  cloud scavenging efficiency  for sulfates formed from the oxidation of SO2 (Chang
28      etal., 1990).
29           The principal interest in quantitative field studies of aerosol-cloud interactions is the
30      scavenging of acidic aerosol mass by  clouds. The focus of measurements in these studies
31      (from aircraft or at fixed mountain  sites) was on gross spatial averages  (over 10s of km) of

        April 1995                              3.109      DRAFT-DO  NOT QUOTE OR CITE

-------
 1      species mass concentrations (mostly of sulfate and nitrate) based on batch samples collected
 2      in cloud water, and in cloud and clear air (Scott and Laulainen, 1979; Sievering et al., 1984;
 3      Daum et al., 1984, 1987; Hegg et al., 1984; Hegg and Hobbs, 1986; Leaitch et al., 1986b;
 4      Pueschel et al.,  1986).  In some studies, continuous measurements  of aerosol size spectra
 5      were used to derive spatially-averaged aerosol volume concentrations (Leaitch et al.,  1983;
 6      Hegg et al., 1984; Heintzenberg et al., 1989) based on which, aerosol volume scavenging
 7      efficiency was inferred.  In one study, continuous measurements of light scattering
 8      coefficient were used as a surrogate for aerosol  mass concentration (ten Brink et al.,  1987).
 9      In these studies, inferences of the efficiency of aerosol scavenging were generally based on
10      comparisons of species mass or volume concentrations (or their surrogates) in cloud water
11      and/or cloud interstitial air with those in putative pre-cloud air.  Such inferences can be
12      confounded by incorrect identification of pre-cloud air, non-Lagrangian sampling, extended
13      sampling periods and resultant averaging of spatial inhomogeneities (including clear air
14      pockets within clouds), and inadequately resolved contributions of aqueous-phase chemistry.
15      Not surprisingly, the  results  of the  above studies varied quite widely.  Most commonly,
16      however, mass scavenging efficiency was found to be high (>0.8).
17           The above studies based on spatially-averaged particle mass concentrations could not
18      address the issue of main concern with respect to radiative transfer, namely, the partitioning
19      of cloud particles  between droplets and interstitial aerosol in terms  of their local number
20      concentrations.  Field studies focused on aerosol scavenging  based  on particle number
21      concentrations are relatively scarce.  In the study of Leaitch  et al. (1986) for stratiform and
22      cumuliform clouds, the authors took special care to ensure Lagrangian adiabatic
23      interpretation by comparing the instantaneous cloud droplet number concentration at a single
24      location within the adiabatic updraft core near cloud base with the below-cloud aerosol
25      number concentration. They found that activation efficiencies so defined were generally high
26      when pre-cloud AMP concentrations were less than about 750 cm"3, but dropped off non-
27      linearly at higher particle loading.  Raga and Jonas (1993) made a  similar observation when
28      comparing droplet concentrations near cloud top with the sub-cloud aerosol concentrations on
29      the assumption that the latter represented the pre-cloud condition.
30           Gillani et al.  (1994) demonstrated that such an assumption was not generally valid in
31      stratiform clouds which are layered and may include  sharp inversions decoupling the layers

        April 1995                               3-110     DRAFT-DO NOT QUOTE OR CITE

-------
  1      from each other and from the sub-cloud layer.  For such clouds, the adiabatic assumption
  2      made in 1-D cloud models is not generally valid.  To circumvent this difficulty with respect
  3      to identification of pre-cloud air, Gillani et al. defined fractional activation (F) in terms of
  4      local variables only, as the ratio of cloud droplet concentration (activated particles) to total
  5      particle concentration (droplet concentration + concentration of unactivated accumulation-
  6      mode particles, 0.17 to 2.07 jum diameter).  In their study (aircraft measurements  in and near
  7      stratiform clouds near Syracuse, NY in the Fall of 1984), continuous in situ measurements
  8      were available for particle number concentrations in 15 size classes each for the droplets and
  9      for dried (by heating the probe  inlet air) interstitial aerosols.  Thus, they were able to
10      determine F at  a high spatial resolution throughout the clouds studied (continental  stratiform).
11      It was determined that accumulation-mode particles larger than 0.37 ^m  were efficiently
12      activated in the cloud under all measurement conditions, but that particles in the range 0.17
13      to 0.37 jum were often activated only partially.  Partial activation generally correlated with
14      high local total particle concentration (> 600 cm"3) and with low temperature lapse rate
15      (surrogate for cooling rate with ascent, dT/dt  = w. dT/dz, where  w= the mean long-wave
16      updraft speed),  the two conditions most responsible for limiting supersaturation.  It is
17      important to note that w  is a most difficult quantity to measure, and is not generally available
18      in field measurements.  Under  the most polluted conditions in a stable stratus, fractional
19      activation of the accumulation-mode particles was as low as 0.1 in the core  of the cloud.
20      Statistically, based on ten days  of measurements in the Syracuse study, it exceeded 0.9  in
21      36% of the data in cloud interior, but was below 0.6 in 28% of such data.   It was generally
22      quite low in cloud edges.  Evidently, the assumption made in RADM of total activation is
23      questionable for stratiform clouds.
24           Simple parameterizations  of fractional activation in clouds have been developed based
25      on 1-D adiabatic Lagrangian models (e.g., Twomey,  1959; Ghan et al.,  1993), and generally
26      highlight the significance of particle loading and updraft  speed  (model calculated).  The 1-D
27      adiabatic approach is useful near cloud base and in updraft cores, but it breaks down near
28      cloud edges and in the upper portions of clouds  where entrainment and mixing effects are
29      substantial.  It is also questionable in the presence of additional complexities such  as cloud
30      layering (Gillani et al., 1994) and lifting and sinking motions (Baker and Latham,  1979;
31      Pruppacher and Klett, 1978). These complex  effects  result in three-dimensional spatial

        April 1995                               34 H       DRAFT-DO NOT QUOTE OR CITE

-------
  1      inhomogeneities and multi-modal droplet size spectra which are uncharacteristic of the simple
  2      adiabatic model.
  3           Noone et al. (1992) studied activation in ground fogs.  They were able to infer size-
  4      segregated volume and number scavenging efficiencies of aerosols (using a counterflow
  5      virtual impactor) in the fog under conditions of very high particle loading and extremely low
  6      supersaturations.   For such highly-polluted fog conditions, they found high activation
  7      efficiencies (>0.8) only for particles larger than 0.8 pcm.
  8           In  most cloud and fog studies which include considerations of particle composition, use
  9      is made  of the  concept of water-soluble mass fraction  (s).  This implicitly assumes internally
10      mixed particles.  As was shown by Zhang et al. (1993), there may really be two e's, one
11      (£m) for  the "more" hygroscopic particles, and one (gj) for the  "less"  hygroscopic aerosols.
12      In the diagnostic  modeling study of Pitchford and McMurry (1994), the two-e concept was
13      implemented.   For clouds and fogs, this implies that Sc may be different for different
14      particles in the same size range.
15           The interaction between aerosols and clouds modifies not only the clouds, but also the
16      aerosols. The  condensation-evaporation cycling of aerosols through non-precipitating clouds
17      generally results in growth of the nuclei due to microphysical and chemical processes during
18      their  in-cloud residence (Hoppel, 1988; Hoppel et al., 1990).
19
20
21      3.5  DRY DEPOSITION
22      3.5.1 Theoretical Aspects of Dry Deposition
23           Dry deposition is commonly parameterized by the deposition velocity, Vrf (m s"1) which
24      is defined as the coefficient relating the pollutant deposition flux F (g m'V1) and the
25      pollutant concentration c (g/m3) at a certain reference  height above the surface, i.e.,
26
                                               F = Vdc                                  (3-39)
27
28      The deposition velocity can be expressed as the inverse of a sum of "resistances" in three
29      layers adjacent to the surface (Sehmel, 1980; Hicks, 1982):
30
        April 1995                               3-112      DRAFT-DO  NOT QUOTE OR CITE

-------
 1           1.     The aerodynamic layer, i.e., the layer in which atmospheric turbulent fluxes are
 2                 constant (typically extending to about 20 m above the ground). In this layer,
 3                 pollutant transfer,  whether gas or particle,  is controlled by atmospheric
 4                 turbulence.
 5           2.     The surface (or quasi-laminar) layer, a thin layer (~ 1 mm) just above the
 6                 surface in which transport occurs by molecular diffusion.  In this layer, gases
 7                 transfer to the surface by molecular diffusion and particles undergo Brownian
 8                 diffusion and inertial impaction.
 9           3.     The earth/canopy/vegetation surface, the actual pollutant sink
10
11           For gases, the deposition velocity is a function of these three types of resistance as
12      follows:
                                         VH = (r  + r  + r V1
                                           d   v a    s    CY
13
14      where ra is the atmospheric resistance  through the aerodynamic  layer,  rs is the surface layer
15      resistance, and rc is the canopy/ vegetation resistance.  All resistances  are in units of s m "J.
16           The aerodynamic resistance ra can be expressed (Wesley  and Hicks, 1977) by:
17
18
19      where zs is the reference height (m) (~ 10 m),  z0 is the roughness length (m), k   is the von
20      Karman constant (0.4), «* is the friction velocity (m s"1), and fh  is the stability correction
21      factor.  Roughness lengths vary from about 10"5 m for very smooth surfaces (ice, mud flats)
22      to 0.1 m for fully grown root crops, to 1 m for a forested area, to 5-10 m for an urban core
23      (Seinfeld,  1986).
24           The surface layer resistance can be parameterized as a function of the Schmidt number
25      Sc = v/D,  where v is the kinematic viscosity of air (m2/s) and D is the molecular diffusivity
26      (m2/s)of the species, as
        April 1995                                3-113      DRAFT-DO NOT QUOTE OR CITE

-------
                                              rs = d,-                                     (3-42)
                                               s     'ku*
 1
 2      where dlt d2 are empirical parameters (d1 =~1.6 - 16.7, and d2  =0.4-0.8, with a
 3      suggested choice of d\ =  5, d2 = 0.66).
 4           The canopy resistance rc  for a gaseous species can be parameterized (Yamartino et al.,
 5      1989) as:
                                   rc =  [LAAI/rf + LAI/rcut + l/r^"1                      (3-43)
 6
 7      where LAI is the leaf area index (i.e., the ratio of leaf surface area divided by ground surface
 8      area), /y  is the internal foliage resistance, rcut is the cuticle resistance, and r  is the ground
 9      or water  surface resistance.  Values for /y are discussed by O'Dell et al. (1977).  The
10      resistance rcut is parameterized by Pleim et al. (1984).
11           For gaseous pollutants,  solubility and reactivity are the  major factors affecting surface
12      resistance and net deposition  velocity. For particles, the factor most strongly influencing the
13      deposition velocity is the particle mass or, assuming similar densities, the particle size.
14      Particles  are transported toward the surface by  turbulent diffusion, which for larger particles
15      is enhanced by gravitational settling.  Across the quasi-laminar surface layer very small
16      particles  (< 0.05 /urn diameter) are transported primarily by  Brownian diffusion, analogous
17      to the molecular diffusion of gases.  The larger particles possess inertia, which may enhance
18      the flux through the quasi-laminar sublayer.
19           The downward pollutant flux is  the sum of the turbulent diffusive flux and a flux due to
20      gravitational settling,  i.e.,
                                        F(z) = Fd  +  VgC = VdC                           (3-44)
21
22      where Vg is the gravitational  settling velocity of the particle.  Whereas in the formulation of
23      the algorithm for gases the analogy with electrical resistance  is straightforward, it is less so
24      for particles.  This is because at any height within the aerodynamic layer and surface layer
25      the flux of trace gases is diffusive only and hence a function of the concentration gradient.
26      Consequently, when equating the fluxes through each layer under the steady-state
27      assumption,  the deposition velocity may be cast in a form proportional to the inverse of a
        April 1995                               3414      DRAFT-DO NOT QUOTE OR CITE

-------
  1     sum of resistances.  Nevertheless, the electrical resistance analogy can still be employed for
  2     particles.  The gravitational settling velocity is merely represented by the reciprocal of an
  3     additional resistance acting in parallel with the diffusive resistance.
  4          As noted earlier, for particles, the resistance in the vegetation layer (rc) is usually
  5     assumed to be zero, since particles that penetrate  the surface layer are assumed to stick to the
  6     surface.  The expression for deposition velocity in terms of the resistances, modified to
  7     include gravitational settling, is
  8
                                     Vd = (ra  + rs + r^Vg)'1  + Vg                        (3-45)
  9
 10          Therefore, the deposition velocity of particles may be viewed in terms of electrical
 11     resistance as the reciprocal of three resistances in series (ra, rs, and /y^V) and one in
 12     parallel (1/VJ.  The third resistance in series is denoted here as a virtual resistance in view
                    o
 13     of the fact that it is a mathematical artifact of the  equation manipulation and not a physical
 14     resistance.  Equation (3-109) is usually implemented with ra (particles) equal to ra (gases),  in
 15     which ra is computed by Equation  3-105, and the surface layer resistance is
 16
                                      rs = (Sc~2/3 - l(T3'Vu *-1                         (3-42)

 17
 18     where Sc is the Schmidt number based on D, the Brownian diffusivity of the particle in air,
 19     and St is the Stokes number, St=F'u2lgn.   The surface layer resistance incorporates the
                                         o
20     effects of both Brownian  diffusion,  through the Schmidt number, and inertial impaction
21      effects, through the Stokes number.
22          The gravitational settling velocity Vg is a function of the particle size, shape and
23      density. For spherical particles (Seinfeld, 1986),
24
                                           v  -  dp2g(pp  - pa)C                              (3.46)
                                            g "
25
        April 1995                                3415      DRAFT-DO NOT QUOTE OR CITE

-------
 1      where d  is the particle diameter (m), p  is the particle density (g /m ), p& is density of the
 2      air (g /m3), m  is the viscosity of air (g rrf's"1), and C  is the slip correction factor
 3
                             C  = 1  + (2X/dp)[1.257 + 0.4exp(-0.55dp/A)]                (3-47)
 4
 5      where X. is the mean free path of air molecules (k = 6.53 x  10"6 cm at 298K)
 6           Figure  3-12 shows particle deposition velocities based  on wind tunnel measurements.
 7      Deposition velocities are presented as a function of particle  diameter, particle density, and
 8      surface roughness height.  Particle deposition velocities exhibit a characteristic minimum as
 9      a function of particle size.  For the smallest particles, deposition velocity increases as
10      particle size  decreases because diffusion  by Brownian motion increases as particles get
11      smaller.   For the largest particles, gravitational settling becomes important as particles get
12      larger so  the deposition velocity increases as particles increase in size.  A characteristic
13      minimum in deposition velocity results in the range of 0.1 to  1.0 jim diameter where neither
14      Brownian diffusion nor gravitational settling  is strong enough to control  removal.
15           It is possible to obtain a rapid estimate  of the atmospheric lifetime  of particles with
16      respect to removal by dry deposition.  If the  aerosol can be  assumed to have a uniform
17      concentration between the ground and a height h, then the residence time relative to
18      removal by dry deposition is  h/Vd.  For example, for a 1000 m atmospheric layer, and a
19      particle deposition velocity of 0.1 cm/s, the estimated residence time is 11.5 days.
20
21      3.5.2    Field Studies of Dry Deposition
22           In spite of many field measurements and considerable  progress since 1980  in our
23      understanding  of dry deposition processes and their quantification, uncertainties remain
24      substantial.  The problem is extremely complex  involving a large multiplicity of factors,
25      and their complex interactions,  which influence dry deposition of atmospheric particles and
26      their precursors (see, for example,  a tabulation of some of these in Davidson  and Wu,
27      1990).  These  factors relate to characteristics of the atmosphere, nature of the
28      depositionsurface, and properties of the depositing species.  It is impossible in field studies
29      to measure all the pertinent variables  over large enough spatial and  temporal  domains.  In


        April 1995                               3-116      DRAFT-DO NOT QUOTE OR CITE

-------
             _  5 ton truck (8-11 ms'1 speed)
             0.1   0.2  0.4     124     10   20   40
                                      Diameter, urn
                         100  200
Figure 3-12.    Extrapolations from correlations of windtunnel measured  deposition
               velocities for z = 1 m, densities of 1, 4, and 11.5 g cm"3. VT represents
               terminal settling velocity.

Source:  Sehmel, 1980, as presented by Nicholson, 1988.
April 1995
3-117
DRAFT-DO NOT QUOTE OR CITE

-------
 1      essence, knowledge  of dry deposition is limited by the inability to make the necessary
 2      measurements in other than special circumstances.  This was a key statement of the NAPAP
 3      Workshop on Dry Deposition in Harpers Ferry, West Virginia  (Hicks et al., 1986).  The
 4      Workshop report also  noted that there is presently a lack of fundamental knowledge
 5      concerning the chemical and biological processes influencing dry deposition, and there are
 6      serious hazards associated with scaling input information down from grid level  to local, and
 7      scaling up the results of local measurements  to broader domains.  Information contained in
 8      the Workshop report and in subsequent research publications on the subject were reviewed
 9      by Davidson and Wu (1990, henceforth to be referred to as "DW90").  That review
10      summarizes the results of a large number of field studies published since earlier reviews by
11      McMahon and Denison (1979), Sehmel (1980), Hosker and Lindberg (1982) and Galloway
12      et al. (1982).  It also includes summaries of dry deposition processes, wind tunnel studies
13      and empirical models,  techniques for measuring deposition in the field, and comparisons of
14      field data and model results. The summary presented in this section is based largely on
15      DW90.
16          A large number of techniques have been used in measurements  of dry deposition.
17      They are generally grouped into two  classes:  surface analysis  methods, which are based on
18      examination of contaminant accumulations on natural or surrogate surfaces, and
19      atmospheric flux methods,  which involve ambient measurements of the species  of interest
20      and other related variables.  These methods provide the deposition  flux out of which the
21      deposition velocity is inferred. Surface  analysis methods include foliar  extraction (by
22      washing individual leaves),  throughfall and stemflow (wet measurements  above and within
23      the canopy),  watershed mass balance, tracer techniques, snow sampling, collection on
24      surrogate surfaces, etc.  These methods may  provide useful data on the flux of  coarse
25      particles, but fail to  simulate the physical processes which control the deposition of sub-
26      micron particles to natural  surfaces, and to give meaningful data on trace gas deposition.
27      Deposition on surrogate  surfaces may not mimic that on natural surfaces.  Atmospheric flux
28      methods include micrometeorological methods (eddy correlation and vertical gradients),
29      aerometric mass balance in a box over the depositing surface, tracer techniques, etc.
30      Micrometeorological methods also include what has come to be known as the inferential
31      approach in which measured concentrations are combined with specified or calculated

        April 1995                               3-118      DRAFT-DO NOT QUOTE OR CITE

-------
  1      deposition velocities based on meteorological  data and surface information.  This approach
  2      is used in long-term monitoring programs in which only simple measurements are possible
  3      at remote sites (e.g. weekly average species concentrations and routine meteorological
  4      measurements).  For details of the various methods, see DW90 and the Workshop report.
  5
  6      3.5.2.1   Measured Deposition Velocities
  7           Measurements of dry deposition in the field and in chambers have primarily involved
  8      six categories of contaminants:  sulfur species, nitrogen species, chloride species, ozone,
  9      trace elements and atmospheric particles.  The results of many of these studies published
 10      between  1978 and 1987 are reviewed in DW90, which includes extensive  tabulations of the
 11      studies and their results.  Of the reported studies on Sulfur Species, 20 pertain to SO2.
 12      They give deposition velocities ranging from nearly 0 to 3.4 cm/s.  The variations are due
 13      to differences in seasonal  and diurnal conditions, aerodynamic transfer, surface
 14      characteristics (especially  stomatal resistance),  measurement  methods, etc. Daytime values
 15      are generally higher, as expected (lower aerodynamic and stomatal resistances).
 16      Micrometeorological methods were used in 16 studies whose average values of vd gave a
 17      grand average of 0.95 ± 0.62 cm/s.  Four studies provided an average value  of 0.13 ± 0.09
 18      cm/s for  deposition velocity on snow.  For particulate sulfur,  34 studies are included, with
 19      10 also including particle  size measurements.   A graph  also includes results  of earlier
 20      studies, and gives values of vd in the range 0.01 to 10 cm/s.   Results for vd in cm/s based
 21      on different methods are as follows:  0.55  ± 0.65 for micromet methods, 0.26 ± 0.25 for
 22      surrogate surface exposures, 0.23 + 0.24 for foliar extraction, and 1.00 ± 0.41 for
 23      throughfall.  Since the micromet method is believed to be more specific for submicron
 24      particles  while the surrogate surface method is biased in favor of  larger particles,  the
 25      difference in the results of those methods is opposite to that  expected.  The surrogate
 26      surface and foliar extraction results are  close,  but each has a large variance.  Throughfall
 27      values are the largest probably partly because they  include deposition of SO2.  Evidently,
28      measurement methods themselves are an important  variable because they do not measure
29      the same  thing.
30           Twenty two species are reported for Nitrogen  Species, including NO2, NOX, HNO3,
31      NO3''NH3, and NH4+.  The inferred values of vd (cm/s) are:  0.012 to 0.5 for NO2 (2

        April 1995                               3_H9      DRAFT-DO NOT QUOTE OR CITE

-------
  1      studies), -2.6 to 0.3 for NOX (4 studies), 0 to 2.9 for HNO3 (4 studies), 0.13 to 1.3 for NO3
  2      ' (7 studies),  1.9 ± 1.55 for NH3 (1 study), and 0.06 to 1.0 for NH4+ (4 studies).  The zero
  3      value for nitric acid was for snow in a chamber study; otherwise,  the values for nitric acid
  4      are the highest, indicating low surface  resistance.  The values for  particulate nitrate are
  5      somewhat larger than for sulfate; this may reflect larger particle size associated with nitrate.
  6      Four studies are reported for chloride-containing particles, giving  values of 1.0 to 5.1 cm/s;
  7      a value for HC1 of 0.73 cm/s on dew was obtained in  one study.  The highest values for
  8      chloride were in winter, related to road salt.  Based on 11 studies  using micromet methods,
  9      vd of ozone on vegetation ranged between  nearly 0 and 1.5 cm/s (average of 15 values =
10      0.39 ± 0.21).  Nighttime values were lower, but the day-night difference  was less for ozone
11      than for NO2.
12           Results of 19 studies included measurements for  21  trace elements,  with particle size
13      data in 15 studies.  For these data, crustal element enrichment factors (EF) were
14      determined.  Values of EF « 1 indicate  crustal  sources, while EF >  1 (enriched) indicate
15      non-crustal sources such as anthropogenic, natural combustion (volcanism, forest fires),
16      biogenic, sea-spray, etc..  Large enrichment factors were found for  Ag, As,  Cd, Cu, In, Pb,
17      Sb, Se and Zn. Ni and V were marginally enriched.  Other elements were mainly soil-
18      derived.   vd for these  elements were generally higher (>1 cm/s), while they were generally
19      less than  1 cm/s for the enriched elements  (smaller, submicron particles).   A figure
20      including these as well as data of earlier studies is presented,  showing a positive correlation
21      between vd and MMD (mass median diameter).  For Pb, the values  ranged between 0.1 and
22      1.0 cm/s.  Friedlander et al.  (1986) have used CO as a tracer  for automobile emissions to
23      estimate the deposition velocity for Pb,  by comparing the ratio Pb/CO in  ambient air to that
24      in a tunnel.  They found the former to  be lower, indicating deposition compared to its value
25      in fresh emission (tunnel).  Based  on these data, they estimated vd for Pb to be 0.26  cm/s,
26      which is consistent with the range given above.  DW90 also report the results of 5 other
27      field studies with micromet measurements  of dry deposition for submicron particles,  and
28      particle size measurements also.  vd was generally less than 1 cm/s,  in general agreement
29      with results  for sulfate and the enriched trace elements.
30           DW90 have also presented results of comparisons between measured values ofvd
31      with predicted values  based on six model calculations.  These results are from published

        April 1995                               3_120      DRAFT-DO NOT QUOTE OR CITE

-------
 1      studies with size distribution data for aerosol sulfate  and trace elements.  The measured
 2      values of vd are for the full size range; the model value is the concentration-weighted
 3      average of the calculated values for all size  classes.  For sulfate, the predicted values were
 4      generally smaller than the measured values.  Good agreement  was, however,  not expected
 5      because of differences in ambient conditions and surface conditions between  values used in
 6      the model compared to the corresponding  measured values.  Similar comparisons for 24
 7      trace elements were also  tenuous:  out of  11 of the 24 elements for which more than one  or
 8      two data points only were available, the measured  values were in the predicted range; for
 9      Al, Ca and Fe, the predictions were low, while for Zn, the predictions were too high.  For
10      the other 13 elements  with sparse measured data, the agreement was generally much poorer.
11
12
13      3.6  WET DEPOSITION
14      3.6.1  Introduction
15           Although detailed physico-chemical models are needed to describe the details of
16      in-cloud and below-cloud scavenging of particles, there has been a benefit in using
17      comparatively  simple  formulations of precipitation scavenging that provide a convenient
18      picture of the process  as  a whole.  These simple methods  are not designed  to explain
19      detailed variations  in wet deposition with time or space, but they are useful in describing
20      average deposition rates over large areas.  Two alternative  techniques have become popular.
21      The first relates concentrations  of material  in precipitation  to the quantity available in the
22      air, thus describing the overall efficiency of precipitation as a removal path.  By relating
23      concentrations in precipitation to those in  the air, dimensionless scavenging ratios  can be
24      determined.  The second  common method is based on the first-order removal of airborne
25      gases or particles as rain falls through the atmosphere.   Concentrations in the air will
26      decrease exponentially and a scavenging rate can then be determined.
27           Below-cloud  scavenging rates for particles of about 3x10"5 s"1 appear  to be typical;
28      in-cloud scavenging leads to rates typically ten times larger (Hicks and Meyers, 1989).
29      Hygroscopic particles  are scavenged more readily than hydrophobic ones.
30           Based on the wet flux  W the wet deposition velocity may be defined as
31

        April 1995                               3-121       DRAFT-DO NOT QUOTE OR CITE

-------
                                          Vw =       A   ~Ah                             (3-48)
                                            w    c(x,y,0,t)
 2      where the last equality assumes that the pollutant is uniformly distributed between z - 0
 3      and z =h.  The wet deposition velocity Vw can be computed by
 4
                                               Vw = wrPo                                   (3-49)
 5
 6
 7      where wr is the washout ratio (i.e., the dimensionless  ratio of the concentration of material
 8      in surface-level precipitation to  the concentration of the material in surface-level air) and p0
 9      is the precipitation intensity (mm hr"1).  For example, if \vr - 106 and p0 = 1  mm h"1, then
10      Vw = 28 cm  s"1,  which gives, for h = 1,000 m, A = 2.8 x 10'4 s'1.  Seinfeld  (1986)
11      provides a detailed discussion of precipitation scavenging  of particles, including the
12      calculation  of collision efficiencies and scavenging rates.
13           Scavenging  ratios  relate concentrations in precipitation to those in  air.  Although such
14      ratios depend on  many factors, they provide a simple  way to include wet deposition
15      processes in air quality  models.  The washout (or "scavenging") ratio is
16
                                             Wr =  [c]fainPa                                 (3-50)
                                                    LCJaerosol
17
18      with [c]rain  in mg g"1, [c]aerosol  in mg m"3,  and ra (=1,200 g m"3), the density  of air.  The
19      definition of this  ratio presumes that the aerosol  measured at ground level is vertically
20      uniform and that  there are no factors limiting the collection of aerosol by the droplets, such
21      as solubility.  Scavenging ratios of about 400 appear to be appropriate in the  case of
22      particles well mixed in the lower atmosphere but originating near the surface,  while  values
23      of about 800 appear characteristic of material derived from the free troposphere (Hicks and
24      Meyers, 1989).
25
26

        April 1995                                 3-122      DRAFT-DO NOT QUOTE OR CITE

-------
 1      3.6.2   Field Studies of Wet Deposition
 2           The removal  of pollutants from the atmosphere by precipitation is the terminal step of
 3      a three-step sequence.  In the first step, the pollutant or its precursor(s) must be delivered
 4      by transport processes to the precipitating cloud .or to the air below it; in the second step,
 5      the species must become incorporated  into the precipitating droplets either within the cloud
 6      (in-cloud scavenging) or below it (washout).  We have already addressed field
 7      measurements  of transport and in-cloud scavenging in liquid water clouds.  The focus here
 8      will be on field studies of the third step.  Wet deposition  measurements are made
 9      principally to meet three objectives:  (1) to determine the regional spatial-temporal
10      distribution and chemistry of wet deposition; (2) to study pathways and mechanisms of
11      pollutant wet removal from the atmosphere;  and, (3) to generate data for diagnostic
12      evaluations of precipitation scavenging modules.  The first of these objectives is best
13      studied based on data of routine monitoring  programs. These are  reviewed in detail in
14      SOST Report No.  6 of NAPAP (1991), and are not covered here.  Our focus here is on
15      recent research field  studies aimed at objectives (1) and (2) above.
16           The scavenging coefficient and the scavenging ratio, in common use  in the Lagrangian
17      models of the  1970s, represent highly lumped representations  of the complex of processes
18      involved in wet removal.  They are empirical entities  which, by themselves, contain little
19      mechanistic information.  While reporting their measurements of scavenging ratio during a
20      year-long study in  Paris, Jaffrezo and Colin  (1988) included a table (Table 3-14) which
21      summarized  not only their own data but also those of other earlier studies.  The various
22      results are not directly comparable owing, at least  partly,  to differences in  measurement
23      methods.  Of particular interest in their study is the interpretation of elemental  composition
24      data. They were able to separate the measured elements  into three groups which differed in
25      terms of their solubility  and also, by the  mechanisms  of their scavenging.  The measured
26      concentrations in precipitation and in air were nearly proportional  for the insoluble species
27      Al, Si, and Fe; this was interpreted to imply that their scavenging  was mostly a local
28      mechanism (below-cloud impaction).  At the other extreme, the local concentrations of the
29      very  soluble species Na and Cl in the two phases were least correlated, indicating a more
30      complex and progressive process  of enrichment of one medium relative to  the other (in-
        April 1995                                3.123      DRAFT-DO NOT QUOTE OR CITE

-------
                          TABLE 3-14.  SCAVENGING RATIOS (DIMENSIONSLESS)
t— '












1
?3
-t*.

O
J>
H
6
o
^
s
o
CJ
O
H
W
Reference n G.M.
Cl 78 2,941
S 82 743
Na 81 444
K 82 951
Mg 81 596
Ca 82 1,048
Zn 69 767
Al 82 291
Si 82 373
Fe 82 184
Ti 9 305
Mn 7 146
1. Jaffrezo and Colin, 1988
2. Harrison and Pio, 1983
3. Arimotoet al., 1985
4. Buat-Menard and Duce, 1986
5. Lindberg, 1982
6. Gatz, 1977
7. Chanet al., 1986
8. Peirsonet al., 1973
9. Cawse, 1981
10. Savoieet al., 1987
* Non-sea sulfate
G.M. Geometric mean
A.M. Arithmetic mean
Med. Median
S.D. Geometric standard deviation
Med.
(1)
2,917
753
530
970
682
1,097
707
283
405
194













S.D.
4.73
1.98
3.17
2.30
2.39
2.49
2.65
2.72
2.35
2.51
1.30
1.36











A.M.
7,710
940
744
1,325
816
1,579
1,226
459
533
267
378
171











A.M.
(2)
600
700
560
620
850
1,890

















G.M. G.M. A.M. A.M. G.M. A.M. A.M. A.M.
(3) (4) (5) (6) (7) (8) (9) (10)
350 1,400 2,300 4,100
1,000 370*
360 2,100 2,900 5,500 490
300 2,000 548
400 457
320 1,100 352 2,100
790 820 179 612 1,050 1,030
580 1,300 756 620 430

390 600 253 468 890 270

250 2,100 3,600 370 756 760











n
i—i
H
W

-------
  1      cloud processes).  The remaining soluble  species (SO", K, Ca, Zn, and Mg) showed an
  2      intermediate behavior.  Earlier data at the same site of the relationship between scavenging
  3      ratio and particle mass median diameter (MMD),  which showed a minimum in the
  4      scavenging ratio for MMD = 1 to 2 ptm (reported as Figure 6-1), were judged to be
  5      supportive of the above interpretation.
  6           A significant  effort in NAPAP in the 1980s  was devoted to development of wet
  7      removal characterizations that directly reflected the cloud physics, attachment, reaction, and
  8      precipitation processes (Hales, 1991).  The PLUVIUS II models, prepared under the
  9      auspices of NAPAP, was a reactive storm model  based on multi-phase material balance,
10      and served as the basis for the development of the one-dimensional  RADM  Scavenging
11      Module, RSM.  A  parallel activity in NAPAP was DOE's PRECP (Processing of Emissions
12      by Clouds and Precipitation)  field measurements program which comprised a series of six
13      individual intensive field studies with the  objective  of systematically measuring scavenging
14      characteristics  for different classes of storm systems important to regional acid deposition.
15      In these, studies, the emphasis was on in situ aircraft measurements.   What follows is a
16      brief review of such research field studies. It is based  substantially  on Hales (1991).  In the
17      context of precipitation scavenging studies, it is useful  to bear in mind that pollutant
18      particles, on average, undergo a  number of repeated cycles in and out of non-precipitating
19      clouds before finally being removed by precipitation.
20           In situ aircraft measurements in clouds and precipitation are of crucial  importance in
21      mechanistic/diagnostic  studies.  Current technology  permits continuous aircraft
22      measurements  of NO, NO2, NOy, HNO3, PAN, SO2, O3, H2O2, liquid water content
23      (LWC),  and size-segregated  aerosol and cloud/rain droplet concentrations with quite high
24      sensitivity and precision.  In addition, filter samples and cloudwater samples can provide
25      mass concentrations of the major ions in  aerosols  and droplets at a temporal resolution of a
26      few minutes.  Ground monitoring of precipitation  in recent studies has included use of the
27      NAPAP-developed  Computer-Controlled  Automated Rain Sampler (CCARS) which is a
28      combination rain gauge and sequential precipitation chemistry  sampler, controlled and
29      monitored by a programmable microprocessor.  Such samplers permit capture  of statistically
30      valid footprints (multiple sequential event samples)  of deposition during the  course of a
31      storm.  Upper-air meteorological measurements with fine vertical resolution  of wind

        April  1995                                3_125       DRAFT-DO  NOT QUOTE OR CITE

-------
 1     components, temperature and moisture are also important.  These can be made using radar
 2     profilers and doppler radars.
 3          Field studies have been conducted in and below point-source plumes (meso-y scale)
 4     and urban plumes (meso-p  scale).  In the former, precipitation scavenging of S and N
 5     compounds was found to be minimal (Granat and Soderland, 1975; Dana et al., 1976;
 6     Drewes and Hales, 1982), indicating low precipitation scavenging efficiency for SO2 and
 7     NOX from fresh plumes.  Hales and Dana (1979) found appreciable removal of S  and N
 8     compounds from the urban plume of St. Louis by summer convective storms.  Patrinos and
 9     Brown (1984),  Patrinos (1985) and Patrino et al.  (1989) found efficient  scavenging of these
10     compounds from the urban plumes of Philadelphia and Washington, DC by frontal storms.
11     H2O2 data in rain showed considerable  spatial variability in the plumes.
12          The major regional-scale field studies include OSCAR (Oxidation  and  Scavenging  by
13     April Rains, April 1981), PRECP (mid-1980s), and the DOE-FBS  (Frontal Boundary
14     Study).   OSCAR (Chapman et al.,  1987) included a nested array of ground level sampling
15     (an extended regional precipitation chemistry network in northeastern  United States, with an
16     embedded high-density network in northeast  Indiana)  as well as three research aircraft.  The
17     focus was on scavenging by extratropical cyclonic storms.  The aircraft  made clear air
18     measurements  before and after frontal passage, as well as measurements within the storm,
19     in the vicinity of the high-density network.  Measurements were made during four storms.
20     OSCAR data have been used for regional model  development and evaluation.
21          The six PRECP  studies, conducted between 1984 and  1988, were targeted at
22     scavenging measurements  in different types of storm systems. Three  studies were focussed
23     on convective  storms  (II, V, and VI) in summer,  and the other three on  extratropical
24     cyclonic and frontal storms during other seasons; five were  conducted east of the
25     Mississippi River, and one  in the Oklahoma-Kansas-Colorado  area. All of them included
26     two or more research  aircraft, and all also included at least limited area precipitation
27     chemistry networks (PRECP IV had three multiscale networks ranging from a coastal "rain-
28     band" network  to a truly regional scale  network).  The network in PRECP VI was a highly-
29     density  network within an 80-km RADM grid cell, designated to provide information about
30     regional sub-grid scale variability.  Two of the studies were conducted jointly with other
31     meteorology-focussed  measurement programs; PRECPII with the NSF PRESTORM study,

       April 1995                              3-126     DRAFT-DO NOT QUOTE OR CITE

-------
  1      and PRECP IV with the NSF-NASA-NOAA Genesis of Atlantic Lows Experiment
  2      (GALE).  Such synergism  resulted in particularly strong meteorological  data in these two
  3      studies.  PRECP  I was  intended to be an exploratory study, but generated  a database of
  4      which at least one storm has been extensively studied (Saylor, 1989). PRECP VI, on the
  5      other hand was designed as the grand finale aimed at generating a definitive database for
  6      evaluation of the RADM Scavenging Module, but failed to meet its main objective owing
  7      to the extreme  drought  of the summer of 1988.
  8           Overall, the studies have developed a substantial database of mechanistic-diagnostic
  9      information suitable for diagnostic model studies.  PRECP II definitively demonstrated the
10      cloud venting phenomenon transporting boundary layer  pollutants to considerable heights in
11      the free troposphere (Dickerson  et al., 1987).  PRECP III provided a significant new
12      mechanistic insight regarding scavenging in orographically enhanced storms, e.g., the
13      observation of an unexpected entrainment mechanism that occurs as  orographic lifting
14      occurs, and which enhances chemical wet removal appreciably (Hales, 1991).  PRECP V,
15      focussed on studying vertical  profiles of chemical species in and around convective storms,
16      resulted in one study (Daum et al.,  1990) which showed that while SO2 was more
17      concentrated in the lower parts of the ABL, H2O2 was concentrated near the top,
18      underscoring the  importance  of mixing in facilitating aqueous-phase  of SO2 by H2O2.  The
19      same study also found that in the low-NOx background,  H2O2 was correlated with humidity.
20           The Frontal Boundary Study (DOE) was conducted in fall 1989 as part of a global
21      study  of the fate  of energy-related pollutants.  The focus was on pollutant redistribution and
22      removal  by stable frontal storms occurring  subsequent to pollution episodes associated  with
23      high-pressure stagnation.  Aircraft soundings ahead of, within, and following the passage of
24      the front showed  considerable spatial variability in precipitation amount and composition
25      (Hales, 1991).
26           The data of the above studies constitute a substantial mechanistic-diagnostic  database
27      for model evaluation. In addition to these research studies, a number of research-grade
28      precipitation  chemistry  networks were also operated in the 1980s.  They include the
29      Canadian CAPMON, and the  U.S. MAP3S and UAPSP, as well as the shorter-term EPRI-
30      OEN and the EPA-ME35.  Applications of the research  network measurements for source-
31      receptor pathway  studies are discussed by Hales et al. (1987).

        April 1995                               3_127     DRAFT-DO NOT QUOTE OR CITE

-------
 1     3.6.2.1  Overview of SO2 and NOX Wet Scavenging
 2           Hales (1991)  has presented a useful overview of our understanding of SO2 and NOX
 3     wet scavenging  based on field measurements which is very pertinent here, and is recapped
 4     below.  SOX:  SO2 is emitted principally from point sources.  It is moderately soluble in
 5     water, and its solubility decreases with increasing  acidity of the solution.  It is not
 6     efficiently scavenged from concentrated  fresh plumes,  but this efficiency improves  as the
 7     plumes dilute.  It is essentially  insoluble in ice and cold snows, but tends to be more
 8     efficiently scavenged by wet slushy  snow and snows composed  of graupel formed by
 9     rimming of supercooled  cloud water.  Only a small fraction of the SO2 emission  is removed
10     as unreacted S(IV) which constitutes about 20% of S in precipitation  in the eastern U.S. in
11     sold seasons (significantly in the form of hydroxymethane sulfonate ions), and virtually
12     none in summer (high acidity of droplets). Sulfate removal is also small  from fresh plumes
13     (not much there), but increases substantially  with plume dilution as more  is formed in the
14     plume.  It is scavenged efficiently by clouds  and rain.   Roughly 1/3 of the S emitted
15     annually in North America is believed to be  removed by precipitation.
16           NOX:  Point  sources are a  relatively smaller contributor of NOX, but still quite
17     substantial.  Both  NO and NO2  have low solubility in water.  Virtually no NOX is removed
18     from fresh plumes.  HNO3 formed by  gas-phase oxidation of NO2 is very soluble in water
19     and is the principal source of NOjin precipitation.  NO3, N2O5, and HO2NO2 are also
20     believed to be significant intermediates.  Since all of the intermediates  are secondary
21     products, NOX scavenging increases  with plume dilution and oxidation. Mesoscale studies
22     show much variation in the efficiency  of wet scavenging of SOX and NOX, depending on
23     storm type and history of plume chemistry.  About 1/3 of the anthropogenic NOX emissions
24     in the U.S. are estimated to be removed  by wet deposition.  The distinct seasonal character
25     of SOX wet  deposition is absent in the case of NOX wet deposition. Some likely  reasons are
26     as follows:  HNO3 has a strong affinity  for ice as well as liquid water; its formation has no
27     direct dependence on H2O2 which peaks in summer; and, there are mechanisms for the
28     formation of HNO3 in low winter sunlight.
        April 1995                               3-128      DRAFT-DO NOT QUOTE OR CITE

-------
  1      3.7   PHYSICAL AND CHEMICAL CONSIDERATIONS IN
  2            PARTICULATE MATTER SAMPLING AND ANALYSIS
  3      3.7.1  Size Cut-Point For Separating Fine and Coarse Particulate Matter
  4      3.7.1.1 Background
  5           In 1979 EPA scientists, in a paper entitled  "Size Considerations for Establishing a
  6      Standard for Inhalable Particles"  recommended that total suspended particulate matter
  7      (TSP), as defined by the high volume sampler, be replaced by the fraction obtained with a
  8      sampler having a precise upper cut-point (originally 15 /^m, but later changed to 10 /*m);
  9      and that "a second particle size cut-point of < 2.5 /xm diameter be incorporated in the air
 10      sampling devices" (Miller  et al.,  1979).   This study  found that "the existence  of a bimodal
 11      distribution  with fine and coarse  modes has been clearly demonstrated by.... mass-size
 12      distribution  studies and by number distribution studies. These size distribution  studies
 13      suggest 1  to 3 /^m as the most appropriate range for a cut-point for fine and coarse
 14      aerosols.  However, practical considerations of reducing plugging of impactor orifices
 15      indicate that 2.5 /xm is a more appropriate cut-point, especially for particle size fractionating
 16      devices such as the dichotomous  sampler" (Miller et al, 1979).
 17           The  cut-point of 2.5 jum, which has been used in many studies since 1979, was chosen
 18      not because  it was ideal but because it was the smallest cut-point deemed feasible for a
 19      dichotomous sampler at that time. Current technology has demonstrated the feasibility of
20      dichotomous samplers with cut-points at 1 pirn, or even lower if desired.  Impactor and
21      cyclone technology can also  be used for cut-points below 2.5 fj.m. Therefore, it is
22      appropriate at  this time to review existing data on size distribution of ambient aerosols  so
23      that policy makers may consider  whether a change to  a smaller cut-point should be
24      considered.  This is especially important in view of the possibility of a major increase in
25      both research measurements, exposure assessment, and regulatory monitoring of fine
26      particles, as  well as of PM]0.
27
28      3.7.1.2 Size Measurements
29          Information  on the size of fine and coarse particles comes  from two basic techniques,
30      (1) particle-counting techniques that  measure the size of individual particles and convert the
31      particle number distribution to a particle  volume  distribution  and (2) particle-collecting

        April  1995                              3429     DRAFT-DO NOT QUOTE OR CITE

-------
 1      techniques that use aerodynamic separation, collection of material in specific size ranges,
 2      and gravimetric or chemical analysis to determine the total mass or the mass of specific
 3      components in the size ranges collected.  Particle counting has the potential advantages of
 4      not causing as much  disturbance to the gas/particle equilibrium.  However, considerable
 5      care must be taken to avoid heating the sample or diluting it with clean or drier air than
 6      that present in the atmosphere.  With particle counting techniques it may also be possible to
 7      avoid problems of particle bounce.  However,  several expensive  and complex instruments
 8      are required to cover the desired range of 0.001 to 100 /zm. Because  sizes can be measured
 9      very precisely, the size ranges covered can be  very small and an almost continuous function
10      of number versus size can be obtained.
11          Particle collecting techniques have the advantage of  obtaining size-differentiated
12      samples for chemical analysis.  The equipment used  is simpler and less expensive.
13      However, aerodynamic separation does not provide as distinct a  classification by size.
14      Large particles may bounce from their intended collection surface and be counted  in smaller
15      size ranges.  Also, the requirement for long sampling times may result in averages of
16      distributions that change with time.  Particle collection techniques provide a limited number
17      of size  cuts and yield discontinuous functions  of mass versus particle  size.
18          Both techniques, however, clearly indicate the natural division of ambient air particles
19      into fine and coarse modes with a minimum between 1.0 and 3.0 /nm  diameter. Size
20      distributions obtained with  particle counting techniques tend to show a lower, broader, and
21      more distinct minimum than distributions obtained with particle collection techniques  such
22      as impactors.  The position of the minimum  between the accumulation and coarse  mode
23      may vary from study to study.  The peak of the fine particle mode tends to increase in size
24      with increasing  concentration and with increasing relative  humidity.  Several good reviews
25      of particle size distribution are available:   on physical properties of sulfur aerosols (Whitby,
26      1978), on the size distribution of urban aerosols (Lippmann, 1980) on sizes of particulate
27      sulfate  and nitrate in the atmosphere (Milford  and Davidson,  1987); and on the size
28      distribution of coarse mode aerosol (Lundgren and Burton, 1995).
29
30
        April 1995                                3-130      DRAFT-DO NOT QUOTE OR CITE

-------
  1      3.7.1.3  Appropriate Display of Size-distribution Data
  2           Size-distribution data, if not properly displayed, can give misleading information on
  3      the position and shape of peaks and valleys and can lead to incorrect conclusions, especially
  4      in regard to the position, width, and separation of fine and  coarse modes.  For this reason
  5      many workers use a histogram display  obtained as follows.   The mass, number, surface, or
  6      volume in each size range is  divided by the difference of the logarithms of the diameters at
  7      the upper D; and the lower DJ.J ends of the size range, and plotted as rectangles  of width
  8      log Dj-log DJ.J  and height,  i.e. mass/ (log Du-log De) on a log diameter scale.  This is
  9      normally shown as AC/A log  D , dM/d log D , or normalized, for example, as AM/M*A log
10      D .   Such histogram  plots  are especially useful for impactor data, which normally yield
11      fewer size intervals than particle-counting  techniques.  Examples of such displays are
12      shown in Figure 3-13 (Wilson et al., 1977) and Figure 3-14 (John et al.,  1990).
13           It is frequently desirable to draw a smooth line through  the data in order to identify
14      modes and the mass mean diameters (MMD)  and widths (o ) of modes.  This can be done
                                                                &
15      by fitting the  data to two or more lognormal distributions,  as  was done in Figure 3-13 (also
16      see Dzubay and Hasan,  1989; and Whitby-DISFIT (TSI, 1993);  or by using an inversion
17      process such as originally  developed by Twomey, as was done in Figure  3-14 (John  et al.,
18      1990; Winklmayr et al., 1989).  In this type of presentation the area in each rectangle or the
19      area under a portion of a curve  is proportional to the mass  in that size range (or  the
20      quantity of any other parameter plotted on a linear scale).  Plotting mass  per impactor stage
21      versus impactor  stage number, or drawing  lines connecting  the midpoints of size range at
22      the heights of the mass in  each  size range, does not provide such quantitative information.
23      Once  the characteristics  of the impactor have  been demonstrated, and once good fits to
24      lognormal distributions have been obtained, repeated measurements of the same species
25      may be shown by curves fitted to inversion or lognormal distributions such as the example
26      in Figure 3-15 (John  et al., 1990).
27           In impactor measurements, the maximum size of the upper stage and the minimum
28      size  of the lower stage (or after filter) are not well defined.  Therefore, an arbitrary choice
29      must be made in order to define the A log Dp. This choice can have a remarkable  influence
30      on the perceptions  of the positions, height, and width of modes.  A particularly
        April 1995                               3-131      DRAFT-DO NOT QUOTE OR CITE

-------
          Q."
          O

          O
                           Electrical aerosol analyzer
              0.002    0.01
                                100
      Figure 3-13.   An example of histogram display and fitting to log-normal functions for
                    particle-counting size distribution data.  Instruments used and the
                    range covered by each  are shown.  Counts are combined into
                    reasonably-sized bins and displayed.  Lognormal functions, fitted to the
                    data, are shown with geometric mean sizes (MMD)  of each mode and
                    the width (
-------
•fe
*
£
Q
8s
T3
8
300
250
200
•1 CO
lOU
100
50
n
-
-
_
H




n



(a)

' — •— T~
i 	 1 —
                    0.01
0.1          1          10
  Aerodynamic Diameter (urn)
            100
                300
                250
                200
              §" 100
              •O
              (J  50
              •O
                  0
                    0.01
                                  Inverted Size Distribution
                                (b)
0.1          1          10
  Aerodynamic Diameter (urn)
            100
                300
                250
                200
              o 100

             I  50
                  0
                    0.01
                                   Lognormal Fit
                                (c)
0.1          1          10
  Aerodynamic Diameter (urn)
            100
Figure 3-14.   An example of an effective display of impactor data:  (a) histograph
              showing mass found on each impactor stage and upper and lower cut
              points of each stage, (b) inverted size distribution, (i.e., a smooth
              distribution that would give the observed distribution considering the
              actual efficiency of each stage; cut points are not exact;  each stage allows
              some large  particles, which it should collect, to pass through to the next
              stage and collects some small particles which it should pass on to the next
              stage), (c) the solid line is the distribution obtained by fitting a sum of
              several lognormal functions to the inverted distribution. The dashed lines
              show the lognormal functions obtained from the fitting process.

Source:  John et al., 1990.
April 1995
         3-133
DRAFT-DO NOT QUOTE OR CITE

-------
               a1
               o>
300
250
200

150
               Q
               O)
               I 1°°
                   50
                    0
                                                                •0600-0930
                                                                 1000-1330
                                                                 1400-1730
                                                                 1800-0100
                      0.01
                 0.1           1           10
                  Aerodynamic Diameter (|im)
                100
       Figure 3-15.  Size distributions of sulfate, Long Beach, June,  1987, showing use of fitted
                     log-normal distributions to describe diurnal variations in size and
                     concentration.
       Source:  John et al., 1990.
 1     3.7.1.4 Comparison of Particle-counting and Particle-collection Techniques
 2          Unfortunately, there have been few efforts to compare results of the two particle-sizing
 3     techniques.  One such effort is shown in Figure 3-17 (Durham et al., 1975).  The differences
 4     between the two techniques, as evident in the figure, are qualitatively observed in individual
 5     studies using either of the two techniques.  Particle counting techniques usually give a lower
 6     and wider minimum. Typically particle counting leads to volume distributions plotted versus
 7     geometric size (or more properly, geometric size inferred from mobility or optical size);
 8     whereas impactor separations give mass versus aerodynamic size.  In Figure 3-17 both
 9     geometric and aerodynamic scales are given.  This figure illustrates the problems involved in
10     defining particle "size" and serves as a reminder that each particle sizing technique gives a
11     different "size".  The upper scale, used for impactor data, is given in  aerodynamic diameter.
       April 1995
                             3-134
DRAFT-DO NOT QUOTE OR CITE

-------
            a) Yugoslavia, Winter B, Author's Original Endpoints, 0.1 and 20 |im
             300
                                                               MMD    aa  %Mass
                                                               0.256   1.28   39.8
                                                               1.93    1.09   19.9
                                                              13.1     2.31   40.3
Authors
original-,
curve  1
                                      1.0                  10.0
                                    Aerodynamic Diameter, Dp (urn)
                                               100.0
           b) Yugoslavia, Winter B, Replotted with New Endpoints, 0.4 and 11  u,m
             300
                                                            Mode MMD   oa  %Mass
                                                              1   0.671  1.09  44.6
                                                              2   1.83  1.28  25.4
                                                              3   7.89  1.20  30.1
                                      1.0                  10.0
                                    Aerodynamic Diameter, Dp (u,m)
                                               100.0
             3-16.  Effect of changing endpoints.  This example of impactor data shows how
                   the lack of a well-defined upper and lower size limit can affect the
                   perception of location of fine and coarse particle modes. The curve drawn
                   by authors of the report, and a histogram with an upper limit of 20 /tin
                   and a lower limit of 0.1 urn diameter, are shown in Figure 4a.  In Figure
                   4b a histogram with a lower limit of 0.4 /tin and an upper limit of 10 /*m
                   is shown.  Notice how the author's free hand curve and histogram suggest
                   a fine particle MMD around 1.5  /im diameter. A quite different idea of
                   the location of the modes is given when different endpoints are chosen and
                   when the data is fitted to a 3-lognormal mode distribution.  Much of the
                   material found between 1.0 and 5.0 /tni is probably smaller  particles
                   caught on the glass fiber impactor stages which have very poor separation
                   efficiencies.

      Source:  Sega and Fugas, 1984.
1     The aerodynamic diameter of a particle is the diameter of a particle of density, p = 1.0,

2     which would behave similarly with respect to impaction as the particle in question.  For
      April 1995
              3-135
DRAFT-DO NOT QUOTE OR CITE

-------
               CO
               E
               O)
               _o
c
.o
.n
*h-
w
b
o
1
§
I
I
I
                     0.01
140 -


120 -


100 -


 80 -
                    60 -
                    40 -
     Aerodynamic Diameter, Dae, urn
         0.1             1.0
                                                      10
                                                           TIM
Denver-Welby, Nov.12,1971.
   Pollution Decay
-O- MAAS
     A*y ,»„   1 CFM Andersen
    • After Filter I

     ?Lage|-,.  } 2.54 CFM Andersen
     After Filter I
                                      0.1              1.0
                                      Particle Diameter, Dp, urn
                                         10
                                                                            140
                                                                            120
                                                                            100
                                                                            80
                                               60
                                               40
                                                                            20
                                                                                 O)
                                                                                 c
                                                                                 o
                                                                                 I
                                                                                 b
                                                                                 i
                                                    I
                                                    o
                                                    O
                                                    GO
      Figure 3-17.  These size distributions, obtained during an EPA study of the Denver
                    brown cloud represent one of the few efforts to compare particle-counting
                    and particle-collection size-distribution measurements. Note that impactor
                    data is given in aerodynamic diameter  and particle-counting data is given
                    in geometric diameter derived from the number distribution and estimated
                    density.

      Source: Durham et al., 1975.
1     spheres, the aerodynamic diameter, Da, equals \[p Dp, where p is the density of the particle
2     and D is the geometric diameter.  Since coarse particles are expected to have a greater

3     density than fine particles, converting the volume, geometric-size distribution to a mass,

4     aerodynamic-size distribution would increase the apparent size of the volume distribution

5     above 1 /zm and widen the minimum.  For small particles, below 0.5 pm,  or at reduced
6     pressures  where the mean free path of the gas molecules is of the same order, or larger than

7     the particle diameter, the Stokes diameter, which is more closely related to the diffusion
      April 1995
                  3-136
                                           DRAFT-DO NOT QUOTE OR CITE

-------
  1      coefficient, is a more useful parameter.  The relationships between Stokes, aerodynamic and
  2      geometric diameter are discussed in Section 3.1.7.1.
  3           The particle size distribution shown on the bottom of the graph was derived from a
  4      combination of a mobility counter and several optical counters. The "mobility size",
  5      obtained from the electrical aerosol  analyzer (EAA) in earlier  studies and the differential
  6      mobility analyzer (DMA) in more recent studies is dependent on the particle shape but not
  7      the density.  For irregularly shaped particles the "mobility" size gives the  Stokes diameter,
  8      which is the geometric diameter of a sphere with the same aerodynamic drag.  For a sphere
  9      the Stokes diameter and the geometric diameter are the same.  By comparing the mobility or
10      Stokes diameter to the aerodynamic diameter it  is possible to measure the density of spherical
11      particles (Stein et al., 1994).
12           The  "optical" size of a particle depends on the particles  shape and refractive index, and
13      on the characteristics of the optical counter. The amount of light scattered by a particle at  a
14      wavelength near the particle size varies rapidly  with changes in size, wavelength, refractive
15      index, and scattering angle.  Therefore,  several different optical counters may be needed to
16      cover the range of atmospheric particle sizes. Because of non-linearities in the response of
17      laser or narrow wavelength optical counters to size changes it  is especially difficult to
18      measure particles in the 0.5 to 1.0 size range (Hering and McMurray,  1991; Kim, 1995).
19      Since the amount of scattered light depends strongly on the refractive index it would be
20      useful to calibrate optical counters with particles of the same refractive index as those in the
21      atmosphere.  Hering and McMurray (1991) used a differential mobility analyzer to select
22      particles of a uniform geometric diameter.  The light scattering of these monodispersed
23      atmospheric  particles, as measured by a Particle Measuring System LAS-X optical counter,
24      was compared to that of spheres of polystyrene  latex (a substance frequently used to calibrate
25      optical counters) and oleic acid of the same geometric diameter.  The atmospheric aerosols
26      scattered less light than polystyrene  latex sphere (refractive index  = 1.9 -  O.Oi), but about
27      the same amount of light as oleic acid spheres (refractive index =  1.46 - O.Oi) of the same
28      geometric  size. Relating the variety of sizes measured by particle counters and impactors,
29      and combining them into a single size indicator, is a major task which has not yet been
30      adequately addressed.
        April 1995                                3-137      DRAFT-DO NOT QUOTE OR CITE

-------
 1           The greater width of the coarse modes, as measured by the impactor in Figure 3-17,
 2     may be attributed to the use of glass fiber filter paper for the impactor surface.  It is now
 3     recognized that the use of glass fiber filter material,  as contrasted to a flat surface, causes a
 4     severe reduction in the effectiveness of the cut.  Large particles  bounce off the glass fiber
 5     (Vanderpool et al., 1987) giving much reduced collection efficiencies; whereas fine particles
 6     penetrate into the fiber and some are captured in  stages that should have near zero collection
 7     efficiencies (Rao and Whitby, 1978). Many studies  that used the Anderson High Volume
 8     Fractionating Sampler also used glass fiber filters. The use of glass fiber filters as impaction
 9     collection surfaces causes any given size range to contain both larger and smaller particles
10     than predicted and thus tends to spread out the modes and fill in the minima.  An example  of
11     the smoothing effect of glass fiber collection surfaces, and especially the collection of fine
12     particles on upper stages, can be seen in Figure 3-16.  Nevertheless, the bimodal nature of
13     the ambient aerosol is still captured.
14
15     3.7.1.5  Review of Size Distribution Data
16     Early Studies
17           In  1978, when EPA scientists debated the best  cut-point to separate fine particles from
18     coarse particles, there was limited information available.  Particle-counting data from
19     California studies had been summarized by Whitby and Sverdrup (1980) and are shown in
20     Figure 3-18.  With the exception of one distribution from Pomona, all distributions showed a
21     minimum near 1 /zm and indications of significant amounts of coarse particle material
22     between 1.0 and 2.5  jum.  (The region between 1 and 2.5 /zm will be referred to as the
23     intermodal region.)  Other studies of size distribution (McMurry et al., 1981) in the
24     Southeastern United States, provided similar information (Figure 3-19).
25           Results from several  impactor studies were  also available,  some of which suggested two
26     modes.  However,  much of the impactor data were considered unreliable in regard to the
27     existence and position of modes (Whitby et al., 1974).  However, one of the more extensive
28     and reliable studies available (Patterson and Wagman,  1977) provided confirmation of the
29     Whitby bimodal observations.  In this study, mass and composition measurements were made
30     for four different levels of visibility. The histograms for mass,  sulfate, and iron for two
31     levels of visibility are shown in Figure 3-20.  It  is clear that the major portion of the fine

       April 1995                               3-138      DRAFT-DO NOT QUOTE  OR CITE

-------
                                                      Richmond
                                                      San Francisco Airport
                                                      Fresno
                                                      Hunter Liggett
                                                      Harbor Freeway
                                                      Pomona
                                                      Goldstone
                                                      Clean Continental
                                                        Background
                                                   1.0    2.5
0.01
      Figure 3-18.
Grand average volume-size distributions from the Aerosol
Characterization Experiment (ACHEX) in 1972.  A size distribution for
clean continental aerosol is shown for comparison.  Note that with the
exception of the Pomona size distribution, all distributions show a distinct
minima near 1.0 fim diameter.  A line has been added at 1.0 /on, 2.5 /on,
and 10 /on diameter to indicate how much of the coarse particle mode is
observed between 1.0 and  2.5 jtm diameter.
      Source:  Whitby and Sverdrup, 1980.
1     mass is below 0.6 ^m and the major portion of the coarse mass is greater than 3 pm in

2     diameter.  These  impactor data, coupled with the more extensive number-size  distributions

3     data of Whitby and Sverdrup (1980) led to a preference for a 1 ^m cut-point but an


      April 1995                              3-139      DRAFT-DO NOT QUOTE OR CITE

-------
          70
          60
          50
       =  40
        Q.
       Q
       S 30
          20
          10
© 16 Km Downwind-13:23 15:21
  Average of 18 Distributions
  5O2 = 78 ppb

® 23 Km Downwind-16:18 17:07
  Average of 8 Distributions
  SD2 = 34 ppb
               A Background
            0.01
                    0.1
                            10
                                             Dp,
      Figure 3-19.  Volume-size distribution taken in the midwestern U.S. near the
                   Cumberland Power Plant in Tennessee.  Note that coarse mode decreases
                   and fine mode increased as the mobile sampling van moved downwind
                   farther from urban influence but allowing more time for reaction as the
                   power plant plume mixed with background air and SO2 was converted to
                   sulfate and NOX to nitrate.

      Source: McMurry et al., 1981.
1     acceptance of 2.5 /zm on the assumption, then considered to be the case, that 2.5

2     represented the minimum cut-point that was attainable with a dichotomous sampler (Miller et

3     al., 1979).
      April 1995
                             3-140
DRAFT-DO NOT QUOTE OR CITE

-------
                  Background visibility
                     M.ot=44.8^g/nnP
                                                 Visibilit level A

2.0
g> 1.5
5 1.0
0.5

1 III
Mass
-
-
• n

Illll 1 IIIMlll 1 Illllll
-






-
-
-
1 1 1 1 1 'i "i i 1 1 1 1 1 1 i IIIIIM
z.o
2.0
1.5
1.0
0.5

1 1 1 ' 1'"' ' ' ' ' 	 1 1 1 Illl
Mass
• — -
-
-


-
	 ^r— n
1 1 1 1 1 1 III 1 IIIMlll 1 Illllll
          0.1 0.20.5    125     10  2050
                       Diameter, urn
                                 100   0.1 0.20.5    1
                   25     10  2050
                  Diameter, urn
                                 100
2.80
|2.10
-; 1.40
& 0.70
n
' Illllll 	 "1111 1 1 1 1 1 III
L
-



1
I I I i I I II
Sulfate ~
-
-
r~i i i IIIMII i i i i 1 1 ii
2.80
2.10
1.40
0.70
n
i i i i 1 1 in i i i i 1 1 M i i iiiliil
n Sulfate ~
1 1
1
I | Illl Ml "' 1 1 llflllll 1 Illllll
         *0.1 0.20.5    1  25     10  2050    100  T).1  0.20.5     1   25     10  2050
                       Diameter, urn                                Diameter, \im
                                                                        100
       1.20

       1.00

       0.80

    ff 0.60

       0.40

       0.20

         0
- Iron (Fe)
1.20

1.00

0.80

0.60

0.40

0.20
- Iron (Fe)
          0.1  0.20.5    1   25     10 2050    100  0.1  0.20.5     1   25    10  2050   100
                      Diameter, urn                                Diameter, \im
Figure 3-20.   Examples of size distribution histograms for total mass, sulfate, and iron
               obtained at two visibility levels using an Anderson impact or. Arbitrary
               choice of 0.1 and 100 for lower and upper limits cause the extreme
               rectangles to be long and low.  Note the separation into fine and coarse
               modes in mass and that sulfate and iron clearly belong in the fine and
               coarse mode respectively.

Source: Patterson and Wagman,  1977.
April 1995
                              3-141
      DRAFT-DO NOT QUOTE OR CITE

-------
 1     Recent Work
 2          In the intervening 15 years, there has been very little additional work in which
 3     particle-counting techniques, covering the entire size range, have been used to measure
 4     ambient aerosols.  Most of the particle-counting studies have focused on fine and ultrafine
 5     particles, diameter < 1.0 urn.  There have however been a number of impactor studies that
 6     provide insight into the size of the fine and coarse modes and into what material is found
 7     between them.
 8          There are only a few impactor size distribution studies that cover the full size range
 9     from 0.01 to 100 am (Lundgren and Hausknecht, 1982; Lundgren et al.,  1984; Burton and
10     Lundgren,  1987; Vanderpool et al, 1987).  Lundgren and co-workers used a  mobile unit, the
11     wide range aerosol classifier (WRAC), to measure mass-size distribution in ten size ranges
12     from <0.4 to >50 pun.  Two distributions, averages for Philadelphia and Phoenix, are
13     shown in Figure 3-21.  Both clearly indicate a fine particle mode with an MMD near 0.5 am
14     for Philadelphia and 0.3 um for Phoenix. Both show a coarse particle mode  with an MMD
15     near 20 am in diameter.  However,  there is a significant amount of material  found in the
16     intermodal region, 1 to 2.5 am.  Although the intermodal mass is not a significant fraction of
17     the total suspended particulate mass  or even of  TSP, as would be measured by a high-volume
18     sample (upper cut-point around 25 am),  it does represent a major portion of  the coarse
19     fraction of PM10.
20          The existing size-distribution data were recently reviewed by Lundgren  and Burton
21     (1995), with emphasis on the coarse mode.  They concluded that the coarse mode could be
22     reasonably well described by a lognormal distribution with a mass mean diameter  (MMD) of
23     15 to 25 am and a mode spread (ae) of approximately two.  This allows one  to calculate, for
                                       &
24     a freshly-generated coarse mode aerosol, that about 1% of the mass would be less than
25     2.5 am and only about 0.1 % would be less than 1.0 /xm in diameter.  This conclusion is
26     confirmed by data from Whitby in which a wind change allowed a measurement of fresh
27     coarse mode aerosol (National Research  Council, 1979).  As can be seen in Figure 3-22, the
28     intermodal mass, 1.0 to 2.5 am, was not affected, even though the mass at 20 am increased
29     substantially.
        April 1995                               3-142      DRAFT-DO NOT QUOTE OR CITE

-------
                      90.0
                                           Philadelphia-WRAC
                    o.
                   Q

                   o
                   T3
45.0-
                                         1.0              10.0
                                        Aerodynamic Diameter, Dp
                                                   100.0
                   o
                   •a
                      90.0'
                      45.0-
                                             Phoenix-WRAC
          Mode
            1
            2
            3
                                       MMD    og   %Mass
                                       0.188   1.54   22.4
                                       1.70    1.90   13.8
                                      16.4    2.79   63.9
                                         1.0              10.0
                                        Aerodynamic Diameter, Dp (urn)
                                                   100.0
       Figure 3-21.   Impactor size distribution measurement generated by Lundgren et al.
                     with the Wide Range Aerosol Classifier:  (a) Philadelphia and
                     (b) Phoenix.  Note the much larger, small size tail to the coarse mode in
                     the dryer environment of Phoenix.

       Source: Lundgren et al., EPA Report.
1          Another extensive set of studies covering the full size range, but limited to the Chicago

2     area, has been reported by Noll and coworkers (Lin et al., 1993, 1994).  They used an

3     Anderson impactor for smaller particles and a Noll Rotary Impactor for larger particles.

4     Results of Lin et al.  also indicate a bimodal mass distribution.  For the shorter time interval

5     measurements (8 or 16 h), the average  MMD for the fine mode was 0.42 /mi, with a ag

6     around two.  The average MMD of the coarse mode was 26+8 /mi, with a oa varying from
                                                                             &

7     2.0 to  3.5.  As shown in Figure  3-23, the results of Noll and coworkers  (Lin et  al., 1993,

8     1994) also suggest that some coarse mode material is found in the intermodal region, 1.0 to

9     2.5 urn.
      April 1995
                         3-143
DRAFT-DO NOT QUOTE OR CITE

-------
                                  Hunter-Liggett
                                  9-14-72
                                    0.1
1
                                           Particle Diameter,
               2.5
10
      Figure 3-22.    Example of aged and fresh coarse mode particle size distributions.
                      A sudden wind change brought fresh wind-blown dust to the sampler,
                      operated as part of the South Coast Air Quality Study.  Note that there
                      is only a very small change in the intermodal mass, 1.0 to 2.5  urn
                      diameter, although there is a major increase in the mass between 2.5
                      and 10 fim in diameter.

      Source: National Research Council, 1979.
1     3.7.1.6 Intermodal Region

2     Coarse mode

3          The question then arises, what portion of the coarse mode material found in the

4     intermodal  region is real and what portion is artifact? As discussed in Section 3.3.3.2.4, the

5     optical size may differ from the geometric or aerodynamic size.  Optical counters are

6     normally calibrated with latex particles, or other particles of a specific refractive index.

7     Atmospheric particles with different refractive indices would be incorrectly sized if the
      April 1995
3-144      DRAFT-DO NOT QUOTE OR CITE

-------
          20.0
      20.0-
                                                                   1.0
                               10.0
           100.0
                 Aerodynamic Diameter, Dp (urn)
             Aerodynamic Diameter, Dp (urn)
          40.0
     30.0-
           0.0
             0.1
10.0
                                          100.0
                 Aerodynamic Diameter, Dp
             Aerodynamic Diameter, Dp
       Figure 3-23.    Size distributions reported by Noll from the Chicago area using an
                      Anderson impactor for the smaller particles and a Noll Rotary Impactor
                      for the larger particles.
       Source:  Lin et al., 1993.
1      difference in refractive index resulted in a difference in the amount of light scattered by the
2      particles (Wilson et al., 1988; Liu et al., 1992; Hering and McMurry, 1991).  For particle
3      counters using lasers, particles of different sizes within the 0.5 to 1.0 /mi range may give the
4      same light scattering (Hering and McMurry, 1991; Kim 1995).
5           In the case of impactors, it is possible that it an artifact may arise from particle bounce,
6      from fragmentation of larger agglomerates, or from loosening of material from other surfaces
7      by impacting particles.  The problem of particle bounce in impactors has been treated
8      theoretically and practically in many studies (Wang and John, 1987, 1988). Most workers
9      coat the coarse particle stages with a grease or oil to reduce  bounce.  However, as the
      April 1995
3-145      DRAFT-DO NOT QUOTE OR CITE

-------
 1      surface becomes covered with aerosols, a particle may impact another particle instead of the
 2      surface and either bounce to a lower stage or cause deagglomeration and reentrainment of
 3      previously collected particles (John et al., 1991; John and Sethi, 1993).   As impactor
 4      collection plates become loaded  or as inlet upper size cut surfaces become dirty, the
 5      magnitude of the effect increases (Ranade et al., 1990; John and Wang, 1991). One result is
 6      a lowering of the effective cut point of the inlet and the impactor stages. Thus, it is
 7      uncertain how much of the mass found in the intermodal size range is real and how much is
 8      due to artifacts.
 9           There are several reasons to believe, however, that some of the intermodal mass may
10      be  real.  For example, Lundgren and Burton (1995) point out that the lifetime of particles in
11      the atmosphere is a strong function of their aerodynamic size. Thus, while freshly generated
12      coarse mode aerosol may have a MMD of 20 /mi, with only  1% below  2.5 jum, as  the
13      aerosol ages the larger particles  will rapidly fall out, leaving  a distribution enriched with
14      particles in the small-size tail of the distribution.
15           A second explanation has to do with the possible multimodal nature of dust generated
16      by  wind or vehicular traffic. A  study  by  the U.S. Army (Pinnick et al., 1985) measured the
17      size distribution of dust generated by heavy vehicles driven on unpaved  roadways in the arid
18      southwestern United States. A variety of light-scattering instruments were used and were
19      recalibrated for the refractive index of the soil particles.  The occurrence of strong surface
20      winds (gusts of 15 to 20 m s"1) during the study permitted, in addition to the vehicular-
21      generated dust, some measurements of windblown dust.  There were some differences
22      between sandy soil and silty soil, and between dust generated by vehicular traffic  and by
23      wind. However, all situations produced a bimodal  size distributions.  The upper mode had
24      an  MMD that ranged from 35 to 53 /mi, with ag from 1.37 to 1.68. Of particular interest,
25      however, was a second mode having an MMD that varied from 6.2 to 9.6 /mi, with a #„
                                                                                         6
26      from 1.95  to 2.20.  (One measurement in silty soil had an MMD of 19.4 /mi.) The MMDs
27      of the smaller coarse particle modes are significantly smaller than those coarse mode MMD's
28      observed by Lundgren or Noll.  An example of vehicular generated dust is shown in
29      Figure 3-24.  Note that the differential mass is plotted on a logarithmic  scale. These results
30      suggest that in arid areas, significant soil material, generated by traffic or wind, may be
31      found in the intermodal region.

        April 1995                               3-146      DRAFT-DO NOT QUOTE OR CITE

-------
           O\
2.
8
U)
I
O
O

z
o
H

O
G
O
H
W

O
90
n
^-^
H
W
       s
       §•

       I'
       en
       n>

       (H

O
O
       cn
       m
       a
       nT
       en
        .
Q.

8
a
en_
N'
     R"  s-   B.  »
     ^-  O   Mi  SI"
O
5
8
»•
en
     0)
     O

     3
     en
     £

     o
     2

     13
     GO
           a
            o
               en
                cy
                0*
                o>
                s?
ode
3-   3


«   tr
cr

i
ffi
O


CD

O)
             T3
              s»
              a.
              Rg
                    en
                    en
     >—  a   a-
     ^^   ^2   ^3
                    O
                    n>
                    3
o    l—

I    I
3    M
O    en
^   a,

<    g1
     o
     R
     S'
     en

     S

    €
    >O
s-   3
                         rt   N
                         »-•   fB
         «•   a
         y    n-
                                                                        S
              n>   c.
         2   g   a
              O

              1
              o.
              n>
              o'
              en
              sr

              en
              O)
              D-
                         en
                         O
              O)

              §

              S.
              n>
                       O

                       3
                       CD
     P>
 p—   a

 S   ^
a   g
 c*.   ^
 O   o
 S.   I
 §   S
 en   _
                   O   "
OSB9
                  rT

                            I
                                  O
                                  D.
                       en
                       en
                            CL
                            O
              I  ^
              en   p
          3
                       2—   f^
                                                                                  Average peak differential mass concentration

                                                                                              dM/dr(ngnv3nm3)

-------
 1     Efficiencies typically reach a minimum between 0.1 and 1 /wm and increase for particles
 2     smaller than 0.1 /mi.  Thus, although most  of the paniculate mass is captured, the particles
 3     that do escape are in the smaller size range.  Data from U.S. EPA, plotted in Figure 3-25,
 4     (U.S. Environmental Protection Agency, 1993) show the size distribution of fly ash from a
 5     pulverized coal power plant and the size distribution of the material escaping from the
 6     various control devices. The small-size tail of the coarse mode escapes preferentially and
 7     likely contributes material in the intermodal region.
 8          Cheng et al. (1985) reported experimental measurements from an atmospheric fluidized-
 9     bed coal combustor.  Size distribution measurements,  made downstream of a cyclone and
10     again downstream from baghouse filtration of the material leaving the cyclone, are shown in
11     Figure 3-26 (Cheng et  al.,  1985).  Electron microscope photographs confirmed a fine particle
12     mode of spherical particles between 0.1 and 0.25 /mi,  presumably formed from evaporation
13     and condensation of volatile species from the coal matrix; and irregular-shaped chunks  from
14     the coarse mode with a peak concentration between 1  and 3 /tin in diameter.
15          A fourth reason comes from a study of the size of particles  collected in various types of
16     samplers.  Burton et. al (1995) used two techniques to measure the size of individual
17     particles collected on filters.  Particles with diameters between 1 and 2.5 /mi were found to
18     account for 18 to 20%  of the coarse fraction of PM10.
19          A fifth piece of evidence comes from studies in  which measurements are made of the
20     elemental composition  of PM2 5 and PM10 or the coarse fraction of PM10.  Elements
21     representative of soil type material have been found in the PM2 5  fraction.  In a study in
22     Philadelphia that used dichotomous  samplers,  an amount of soil-type material equal to 5% of
23     the coarse mode fraction of PM10 was found in the PM2 5 fraction (Dzubay et al., 1988).
24     Since the virtual impactor used in the dichotomous sampler minimizes particle bounce and
25     reintrainment, this would appear to be the small-size tail of the coarse mode in the 1 to 2.5
26     pirn size range.  Similar results have been reported from the IMPROVE network  (Eldred et
27     al., 1995). Elemental  analysis suggested that soil-derived material, equal to 20% of the
28     coarse fraction of the PMj0 sample, was found in the PM2 5 sample.
29          Thus, one can conclude that coarse mode material is found in the intermodal region.
30     There are reasons to suspect that a portion of this material is an artifact but that a portion is
31     real coarse mode material having an aerodynamic diameter between  1.0 and 2.5 /mi. In

       April 1995                               3-148      DRAFT-DO NOT QUOTE OR CITE

-------
o.0-9
O 0.8
OJ „-,
O 0.7
< 0.6
*
in 0.5
Vt **•**
Ui
CO 0.4
CO 0.2
5 0.1
^]
No Controls
. 100% Emitted















	 •— I






~~*




















No Controls
Cyclones
Scrubber
ESP
Baghouse
                                                                       % of Flyash
                                                                       in 1.0 to 2.5 |im
                                                                 of Controlled
                                                                   Emissions
                                                                       6
                                                                       3
                                                                      51
                                                                      29
                                                                      53
                                 of Total
                                 Flyash

                                  6
                                  0.59
                                  3.06
                                  0.23
                                  0.11
                                  Scrubber
                                  6% Emitted
                             1   2.5    10

                          Log diameter, Dp
                   1  2.5    10

                Log diameter, Dp
                          100
      Figure 3-25.      Size distribution of emissions from a pulverized-coal power plant and
                        the particle size distributions remaining after several types of control
                        devices (EPA, AP-42, 1993).
1     either event, this can lead to a misunderstanding of the source of the particles, to

2     inappropriate control strategies, or to exposure studies that fail to differentiate correctly

3     between fine and coarse particles.
      April 1995
3-149
DRAFT-DO NOT QUOTE OR CITE

-------
                               0.05  0.1              1.0     2.5
                                       Stokes diameter,
                              0.05 0.1              1.0    2.5
                                       Stokes diameter, urn
      Figure 3-26.      Size distributions from a fluidized-bed pulverized coal combustor, a
                        after initial clean up by a cyclone collector, and (b) after final clean
                        up by a baghouse.
      Source: Cheng et al., 1985.
1     Fine Mode
2          This section discusses the source of fine mode material found in the intermodal region.
3     Early particle-counting data suggested that, with a few exceptions, significant mass of fine
4     mode material would not be found above 1 /xm (see Figures 3-13, 3-18, 3-19, and 3-20).
5     However, impactor studies, on some occasions, have observed significant mass on stages
6     with a cut point of 1 /zm.  In  some instances, the MMD of the fine mode was as large as
7     1 /*m (John et al., 1990).  The change in relative humidity produced by a few degrees
      April 1995
3-150
DRAFT-DO NOT QUOTE OR CITE

-------
  1      change in temperature can significantly modify the MMD of an ambient aerosol size
  2      distribution.  As the RH approaches 100%, accumulation mode aerosols, with dry sizes
  3      below 1.0 pm in diameter, may grow larger than 2.5 /zm in diameter, be rejected by PM2 5
  4      samples, and be counted as coarse particles.
  5           Before examining additional field data demonstrating the effect of relative humidity on
  6      particle size, it is useful to review some basic information on the interaction of water vapor
  7      with the components of fine particles.  Sulfuric acid (H2SO4) is a hygroscopic substance.
  8      When exposed to water vapor a H2SO4 droplet will absorb  water vapor and grow in size
  9      until an equilibrium exists between the liquid water concentration in the H2SO4 solution
10      droplet and the water vapor concentration in the air.  The amount of water in the sulfuric
11      acid droplet will increase and decrease smoothly as the RH increases and decreases.
12      Ammonium sulfate, (NH4)2SO4, however, is deliquescent.  If initially a crystal in dry air, it
13      will remain a crystal until a certain RH is reached; at this point it will absorb water and
14      become a solution droplet. The RH at which this happens, =80% RH in the case of
15      NH4)2SO4, is called the deliquescent point.  At RH's above the deliquescent point the
16      (NH4)2SO4 droplets are hygroscopic, gaining or  losing water reversibly as the RH increases
17      or decreases.  If the RH decreases below the deliquescent point the solution droplet becomes
18      supersaturated and unstable to crystallization.  However, sub-micron sized droplets  will
19      remain supersaturated until a significantly lower RH, known as  the crystallization or
20      efflorescent point is reached.  The crystallization point decreases with decreasing droplet size
21      and decreasing purity (Whitby, 1984).  Thus, for a deliquescent substance, a plot of droplet
22      diameter or water content as  a function of RH will have two lines, one for increasing RH
23      and another for decreasing RH.  An example of this phenomena, known as hysteresis,  is
24      shown in Figure 3-27.  Table 3-15 shows the RH at the deliquescent and crystallization
25      points for some compounds found in the atmosphere.
26           Much experimental and theoretical effort has gone into understanding this process.  The
27      basic theory was elucidated by Hanel (1976).  Much experimental work has been done on
28      atmospheric species  (e.g., Tang  and Munkelwitz, 1977, 1993; Richardson and Spann, 1984).
29      Ammonium nitrate, NH4NO3, because of its volatility, is difficult to handle but has been
30      studied successfully by Richardson and Hightower (1987).   The aerosol equilibria models
       April 1995                               3-151      DRAFT-DO NOT QUOTE OR CITE

-------
       2.0
    o
    Q.
   Q
       1.5
       1.0
          0
   H
                  L
(NH4)2 S04
                  30
        50          70
          RH, %
                   8

                   7

                   6

                   5

                   4

                   3
                                       1
                                       0
                90
o
Q.
Figure 3-27.     Particle growth curves showing fully reversible hygroscopic growth of
               sulfuric acid (H2SO4) particles, deliquescent growth of ammonium
               sulfate [(NH4)2 SOJ particles at about 80% RH, hygroscopic growth
               of ammonium sulfate solution droplets at RH greater than 80%; and
               hysteresis as the droplet remains supersaturated as the RH decreases
               below 80%  RH until the crystallization point is reached.

Source: National Research Council, 1979.
April 1995
           3-152
DRAFT-DO NOT QUOTE OR CITE

-------
                TABLE 3-15.  RELATIVE HUMIDITY OF DELIQUESCENCE AND
       	CRYSTALLIZATION FOR SEVERAL ATMOSPHERIC SALTS3
        Compound         	Deliquescence	Crystallization0
        (NH4)2S04                  79.9 ± 0.5                  37 + 2
        NH4HSO4                   39.0 ± 0.5
        NH4NO3                    61.8
        NaCl                       75.3 + 0.1                  42
       "Taken from Tang and Munkelwitz (1993) unless otherwise indicated.
       bTang and Munkelwitz (1977).
       cShaw and Rood (1990) and references therein.
 1     developed by Seinfeld and co-workers allow calculation of the water content of bulk solution
 2     as a function of relative humidity.  (Kim et al., 1993a,b).
 3          The water content of a sub-micron sized droplet,  and therefore it size, depends not only
 4     on the dry size but is a result of a balance between surface tension and solute concentration
 5     (Li et al., 1992).  Pure water is in equilibrium with its vapor when the RH equals 100% and
 6     is therefore, stable, i.e. the rate of evaporation equals the rate of condensation. The water in
 7     a solution will be in equilibrium with water vapor at a  lower water vapor concentration
 8     because the presence of solute molecules or ions lower the rate of evaporation. Therefore, a
 9     solution will absorb water and become more dilute, increasing the water vapor concentration
10     needed for equilibrium until the solution water vapor concentration required for equilibrium
11     matches the ambient water vapor concentration or RH. As the droplet size decreases the
12     surface tension increases and the vapor pressure of water required to maintain equilibrium
13     increases. Therefore, the smaller the dry size of the particle, the less the amount of growth
14     as RH increases.
15          Theoretical calculations of the growth of various  sizes of ammonium sulfate particles
16     and an experimental verification of such calculations, using a simulation of the humidification
17     process  in the human lung, are shown in Figure 3-28.  Note the very rapid increase in the
18     amount  of water and in the diameter of the aerosol particle as the relative humidity
19     approaches 100% RH.  Considering the difficulty  of measuring relative humidity accurately
20     between 99 and 100%, theory and experiment are in reasonable agreement.  As can be seen

       April 1995                              3-153     DRAFT-DO NOT QUOTE OR CITE

-------
             O"  4
             .g"
              co
             DC
              2
              CD
                                I             I
                               Theoretical Prediction at 22° C
                       o o o o o  Experimental Measurements
                                                             RH-99.8%
                                                                        216
                                                                        125
                                                          64
                                                                        27
                 ..  I  I   I   I  I   I   I  I   I  I   I   I  I   I   I  I   I  I   I   I  lQ

                                                                      200
                 50           100          150
                NH4 HSO4 Dry Particle Diameter (nm)
tf
O  4
.g"
rr

I  3
(5
                       	  Theoretical Prediction at 22° C
                       o o o o o  Experimental Measurements
                                                               RH-99.5%
                      i  i   i  i
                                I
                                I  i   i   i  i   I  i   i   i  i
                                                                        125
                                                                        64
                                                                        27
                               50           100          150
                              NH4 HSO4 Dry Particle Diameter (nm)
                                                         200
Figure 3-28.     Theoretical predictions and experimental measurements of growth of
                 NH4HSO4 and (NH4)2SO4 particles at RH between 95 and 100% RH.
Source:
April 1995
                           3-154      DRAFT-DO NOT QUOTE OR CITE

-------
  1     the effect of surface tension is most important for particles with dry size less than 100 nm
  2     (0.1 fj.m).  This phenomena may be of importance in considering the biological effect of
  3     water-soluble pollutants.  Accumulation mode particles will be diluted when exposed to
  4     humidification in the lungs.  Ultrafine or nuclei mode particles will not be diluted as much.
  5     In the atmospheric aerosol the number distribution will almost always be dominated by
  6     particles below 100 nm (see Section 3.2.1).  However, aerosols generated in the laboratory
  7     for exposure studies probably lack the smaller particles found in the atmosphere.  This
  8     provides a hypothesis  for the difference in effects observed in epidemiological studies and
  9     laboratory exposure studies. The importance of this more concentrated, ultrafine droplet
 10     component of the atmospheric  atmosphere may have been neglected because most of the
 11     experimental studies of hygroscopicity have used near-micron-sized particles.  However, in
 12     the modeling of deposition of hygroscopic particles,  workers,  such as Martonen (1993), have
 13     corrected the experimental curves of particle size as  a function of RH,  based on
 14     measurements of near micron-sized particles, to account for the effects of surface tension on
 15     ultrafine particles.
 16          In addition to the laboratory studies discussed above there are some measurements on
 17     the effect of RH changes  on atmospheric aerosol.  McMurray and co-workers have made use
 18     of a Tandem Differential  Mobility Analyzer (TDMA) system (Rader and McMurry, 1986) to
 19     measure the change in particle  size with changes in relative humidity at Claremont, CA, as
 20     part of the Southern California Air Quality Study (SCAQS) (McMurry  and Stolzenberg,
 21      1989) and at the Grand Canyon National Park, AZ, as part of the Navajo Generating Station
 22      Visibility Study (Zhang and McMurray, 1993; Pitchford and McMurry, 1994). One  mobility
 23      analyzer is used to isolate a narrow size distribution.  After humidification the size
 24      distribution of this fraction is measured.  An example is shown in Figure 3-29. Note that
 25      Figure 3-29 is a number size distribution not a mass  size distribution. Particle growth with
 26      increasing RH is evident.  However, between 70 and 91% RH the distribution splits into
 27      less-hygroscopic and more-hygroscopic components.  Pitchford and McMurry (1994)
 28      attribute this splitting to external mixing, i.e. there are two relatively distinct classes of
29      particles, both containing  some soluble and some non-soluble  material, with the more
30      hygroscopic component containing significant more soluble and hygroscopic material.   A
31      summary of the results of these studies is given in Table 3-16 (Zhang and McMurray, 1993).

        APril 1995                              3_155      DRAFT-DO NOT QUOTE OR CITE

-------
0
0
"^
 
-------
t
>— *
vo



U>
1
-J
O
?
H
6
0
25
O
H
O
O
3
g
O
H
m
TABLE 3-16. SUMMARY OF HYGROSCOPIC GROWTH FACTORS

Dry Size (/xm)
0.05
0.2
0.4
0.5

Dry Size (/^m)
0.05
0.10
0.20
0.30
0.40
Values are mean + standard deviations.




1987 SCAQS, Claremont,
More hygroscopic peak
Dp(90 + 3% RH)
Dp(0% RH)
1.14 + 0.05
1.23 ± 0.08
1.63 + 0.11
1.59 ± 0.08
1990 NGS Visibility Study, Grand
More hygroscopic peak
Dp(89 + 4% RH)
Dp(0% RH)
1.36 ± 0.08
1.42 ± 0.08
1.49 + 0.11
15.1 ± 0.09
1.43 + 0.10





CA
Less hygroscopic peak
Dp(87 + 2% RH)
Dp(0% RH)
1.03 + 0.03
1.02 ± 0.02
1.04 + 0.05
1.07 + 0.03
Canyon, AZ
Less hygroscopic peak
Dp(89 + 4% RH)
Dp(0% RH)
1.14 ± 0.10
1.17 + 0.09
1.17 + 0.10
1.14 ± 0.10
1.07 + 0.03






-------
 1     The difference in growth rates may be due both to size and to variation in composition as a
 2     function of size.  The lower growth factor for 0.2 yum particles in  Claremont relative to the
 3     Grand Canyon may be due to a higher concentration of non-soluble organic material in
 4     Claremont.
 5           Some experimental examples of the significant effect of relative humidity on ambient
 6     aerosol size distributions are shown in Figure 3-30 (Lowenthal et al.,  1995).  In this  work,
 7     supported by the Electric Power Research Institute, impactor collection and ion
 8     chromatographic analysis were used to measure sulfate size distributions over short enough
 9     periods to demonstrate the effects of changing relative humidities.  The results suggest that
10     the lognormal distribution is preserved as relative humidity increases,  but that the MMD
11     increases.  This effect is especially pronounced as the  relative humidity approaches 100%.
12           There are also studies of the behavior of ambient aerosols as the relative humidity is
13     reduced by heating the sampled air. Shaw and Rood  (1990) report a study using a heated
14     integrating nephelometer in which crystallization RHs  of 4  to 67% were observed.  Similar
15     studies in Washington, D.C. by Fitzgerald et al.  (1982) found no evidence of crystallization
16     or efflorescence when RH was reduced to 30% RH.
17           Further experimental evidence of the effect of decreasing relative humidity on aerosol
18     size distribution is provided by impactor data reported by Berner (1989).  One impactor
19     sampled aerosol in its humidified state directly from the atmosphere.  The inlet of a second
20     impactor was  warmed  «7 °C above the ambient temperature of ~5 °C in order to
21     evaporate most of the particle-bound water before collecting the aerosol.  The water and
22     other volatile  material in both the "wet"  and the  "dry" samples would evaporate in the
23     laboratory prior to weighing the impactor stages.  As can be  seen in Figure 3-31 in the
24     ambient air most of the non-volatile mass was above 1.0 /xm  with significant amounts above
25     2.5 nm. However, after heating the size of the aerosol was reduced so that most of the non-
26     volatile mass was below 1.0 /mi.  Berner treated the distributions as monomodal and derived
27     growth factors of 4.9 for fog and 4.1 for haze.  If the observations are treated as
28     multimodal, good bimodal, or as shown in Figure 3-31, trimodal fits are obtained. This
29     splitting into "more" and "less" hygroscopic modes at high relative humidity has been
30     observed by McMurry and co-workers (McMurry and Stolzenberg, 1989;  Zhang and
31     McMurry, 1993)  (Figure 3-29) and Lowenthal et al.,  (1995) (Figure 3-30). In  some cases,

       April 1995                               3_158      DRAFT-DO NOT QUOTE OR CITE

-------
     Q
     O)
     o


     O
         0
           • RH - 99% 8/12/90, 0200 hr

           + RH < 50%
            Sulfate Size Distributions
          0.01
                                 Diameter
     00
     Q
     D)
     _g


     O
     •D
         0
             RH = 95% 8/4/90, 0200 hr

           H- RH < 50%
            Sulfate Size Distributions     + •+  •
                i    i
          0.01
0.1                1
    Diameter (|im)
10
Figure 3-30.    Example of growth in particle size due primarily to increases in

              relative humidity from Uniontown, PA.


Source:  Lowenthal et al., 1995.
April 1995
     3-159     DRAFT-DO NOT QUOTE OR CITE

-------
                  80.0
                                 Bologna Haze, Wet (Berner, 1989)
Mode
1
2
3
MMD
0.204
1.95
3.50
oa %Mass
1.69 9.9
1.97
2.65
23.5
66.5
                0.1             1.0             10.0
                  Aerodynamic Diameter, Dp (um)
                                                                                 100.0
                 100.0
               0" 50.0-]
               o
                  o.o
                                  Bologna Haze, Dry (Berner, 1989)
   Mode  MMD   og  %Mass
    1    0.130  1.42   10.8
    2    0.589  1.34   57.4
    3    1.65   1.36   31.8
                                                                % Dry mass lost
                                                                upon heating
                    0.01
                0.1              1.0
                 Aerodynamic Diameter, Dp (urn)
                                                                  10.0
                                                                                 100.0
                  70.0
               |
               §"
               T3
                  35.0-
                                 Bologna Fog, Wet (Berner, 1989)
  Mode  MMD   oa  %Mass
   1    0.310  2.09   30.8
   2    1.34  1.93   36.4
   3    5.31   1.91   32.8
                    0.01
                0.1              1.0
                  Aerodynamic Diameter, Dp
                                                                  10.0
                                                                                 100.0
                 200.0
               E
               o

               I
               •
                 100.0-
                  0.0
                                 Bologna Fog, Dry (Berner, 1989)
   Mode  MMD   aa  %Mass
    1    0.145  1.39   17.8
    2    0.524  1.36   65.4
    3    1.56   1.32   13.9
                                                                % Dry mass lost
                                                                upon heating
                                                                10.9%
0.01
                                    0.1             1.0
                                     Aerodynamic Diameter, Dp
                                              10.0
                      100.0
Figure 3-31.       Mass size distribution of non-volatile aerosol material.  The aerosol
                    was collected at ambient conditions, "wet",  or after evaporation of
                    water, "dry".  A - haze; B - fog.


Source:  Berner, 1989.
April 1995
                          3-160
DRAFT-DO NOT QUOTE OR CITE

-------
 1      reported by Pitchford and McMurray (1994), splitting into three modes of varying
 2      hygroscopicity was observed.  However, the separation into two "more" hygroscopic modes
 3      may represent, as suggested by Berner, variations in relative humidity extremes during
 4      different parts of the overnight sampling period.
 5           In measuring light scattering with the integrating  nephelometer, the aerosol community
 6      has been very concerned about the difference in relative humidity and temperature in the
 7      ambient air and in the volume of air in which particle scattering is actually measured (Covert
 8      et al., 1972; Fitzgerald et al., 1982).  Temperature  differences between the measurement
 9      volume and ambient air of 1 or 2 °C can change the relative humidity and change the
10      observed light scattering.  Great efforts have been made to minimize this temperature
11      difference.  However, researchers have not been nearly as careful in considering temperature
12      and relative humidity effects when measuring size distribution, either with impactors or
13      particle counters,  even though effects have been reported in the early literature (Wagman
14      et al., 1967; Sverdrup et al., 1980).
15           A recent paper by Cass and coworkers (Eldering  et al., 1994) provides some insight
16      into how differences in RH resulting from heating can  cause differences between
17      particle-counting distributions and impactor distributions.  Particle size distributions were
18      obtained by counting particles by mobility (electrical aerosol analyzer) and light scattering
19      (optical particle counter).  An example is shown in  Figure 3-32. Almost no particles were
20      found between 1.0 and 2.5 /^m diameter.  When these particle number data were converted
21      to total expected light  scattering, they agreed with measurements made by  a heated, but not
22      an unheated, integrating nephelometer; and when converted to expected mass, agreed with
23      filter measurements of dry mass.  Eldering et al. (1994) conclude that even the moderate
24      heating occurring in mobility and optical counters was  enough to change the size of the
25      particles, especially when the ambient air was close to  100% RH.  It seems likely that most
26      particle counting systems produce some heating of the aerosol, and thus some reduction of
27      the measured particle size from that existing in the ambient air.   On the other hand, if
28      particle-size measuring devices were located in air conditioned or heated trailers or
29      laboratories, the  temperature of the sampled air would be changed and the measured particle
30      size distribution would be different from that existing in the ambient air (Sverdrup et al.,
31      1980).

        April 1995                               3_16l      DRAFT-DO NOT QUOTE OR CITE

-------
    100.00
     80.00--
"E   eo.oo
 0.
 Q
 §>  40.00
 T3
 >
     20.00--
      0.00
Figure 3-32.
                    0.01
                                          Diameter ((im)
                        Example of particle-counting volume distribution obtained in
                        Claremont, CA. Compare to Figures 3-14 and 3-31.  Heating of the
                        sampled air by the mobility and optical counters are believed to have
                        resulted in a distribution representative of a lower than ambient
                        relative humidity.
       Source:  Eldering et al., 1994,


 1          During the high relative humidities that occur at nighttime, growth of hygroscopic
 2     components can result in the growth of some fine mode aerosol to diameters greater than
 3     1.0 /mi and perhaps even above 2.5 /mi.  As can be seen in Figure 3-28, dry ammonium
 4     sulfate particles having a dry  diameter of 0.5 /mi will grow to  =2.5 /mi at a relative
 5     humidity between 99 and 100%.   When the relative humidity actually reaches 100%, the
 6     particles will continue to grow to  maintain the relative humidity at 100%,  and eventually
 7     become fog droplets that are large enough to be collected in the fraction larger than 2.5 /mi.
 8     Ammonium sulfate particles with  dry sizes greater than 0.5 /mi would also grow into the
 9     larger than 2.5 urn size range.
10          The addition of water to hygroscopic particles, discussed  in the previous section, is a
11     reversible process.  Particles  absorb water and grow as RH increases; as RH decreases some
April 1995
                                               3-162
DRAFT-DO NOT QUOTE OR CITE

-------
  1      of the particle-bound water evaporates and the particles shrink. However, the large amount
  2      of liquid water associated with hygroscopic particles at high relative humidity provides a
  3      medium for liquid phase transformation process. A number of atmospheric process, which
  4      convert SO2 to sulfate or NOX to nitrate, can take place in water solutions but not in the gas
  5      phase.  These processes are not reversible but  lead to an accumulation of sulfate or nitrate
  6      and lead to an increase in the dry size of the particle.  Of course as more sulfate or nitrate  is
  7      added to the particle it will absorb more water so that  the wet size will also increase.
  8           The first observation and clear discussion of these combined effects of relative humidity
  9      on growth and SO2  conversion to sulfate are given  by  Hering and Friedlander (1982) as
 10      shown in Table 3-17.  Using a low pressure impactor,  they observed that days with higher
 11      relative humidity had higher sulfate concentration and higher MMD's compared to days with
 12      lower relative humidity. Hering and Friedlander (1982) named the small mode the
 13      condensation mode and suggested that it was formed by the gas phase conversion of SO^ to
 14      sulfate and subsequent nucleation, coagulation, and growth by condensation.  They named
 15      the larger mode the  droplet mode.  They discussed  possible mechanisms for formation of this
 16      mode. They ruled out coagulation as  being too slow.  Reactions in fog droplets  were ruled
 17      out on the basis that fog has too few particles per unit  volume to give the number of particles
 18      found in the droplet mode.  They concluded that growth occurred due to reaction of SO2 in
 19      the liquid water associated with the particle.
20
21
        TABLE 3-17.  COMPARISON OF  SULFATE CONCENTRATION AND MASS MEAN
             DIAMETERS OF AEROSOLS FOR DAYS WITH HIGHER AND LOWER
                                     RELATIVE  HUMIDITY

Minimum RH, %
Maximum RH, %
Sulfate Concentration, /ig/m3
Mass Mean Diameter, pm
Low RH Days
17-35
45 -68
3-9
0.20 + 0.02
High RH Days
26-66
69 - 100
3 -52
0.54 ± 0.07
 1           In a series of papers McMurray and co-workers make use of the aerosol growth law,
 2      originally developed by Heisler and Friedlander (1977), to study the mechanism and rates of

       April 1995                              3_163      DRAFT-DO NOT QUOTE OR CITE

-------
 1      sulfate formation in ambient air (McMurry et al., 1981; McMurry and Wilson,  1982,  1983).
 2      They were able to apportion growth to condensation and droplet mechanisms and observed
 3      droplet growth in particles up to 3 /nm in diameter.
 4           A process of aerosol growth due to increasing relative humidity (Figure 3-33) has also
 5      been utilized by Cahill et al. (1990) to explain observations of sulfate  size changes during the
 6      1986 Carbonaceous  Species Methods Comparison Study in Glendora, CA. Cahill used a
 7      DRUM sampler to measure sulfate in nine size ranges. By tracking the mass of sulfate in the
 8      0.56 to 1.15 /xm size range Cahill et al.  could follow the expansion and contraction of
 9      aerosol particles containing sulfate.  Because of the relative high time  resolution of the
10      DRUM sampler (4 h except for an 8-h increment each night from midnight to 8 a.m.), Cahill
11      et al. (1990) could follow this process as the relative  humidity increased during  the night and
12      decreased during the day.  This data indicates that during the  "Poor Period" (low visibility)
13      particles grow as relative humidity increases.  However, they  did not return to the smaller
14      size observed during the "Fair Period" (good  visibility). This could be due to a combination
15      of growth due to reaction of SO2 to sulfate within the particles or failure  of the  droplet to
16      crystallize thus maintaining particle-bound water in a supersaturated state.
17           John et al. (1990), in studies in the Los  Angeles area, observed a number  of sulfate size
18      distributions with MMD near 1.0 ^m. A histogram of the sulfate MMDs from  his study is
19      shown in Figure 3-34.  John et al. (1990) have provided a qualitative explanation to account
20      for these large MMDs for fine mode aerosol.  In analyzing their data John et al. plotted
21      sulfate mass as a function of sulfate MMD and found two distinct regions, as shown in
22      Figure 3-35.  Distributions with particles near 0.2 /*m diameter are probably still dry; the
23      particles have not reached their deliquescent point. As the relative humidity increases they
24      reach their deliquescent point and grow rapidly into the 0.5 to 0.7 pirn size range.   During
25      the formation of fog, the hygroscopic particles act as fog condensation nuclei, and with
26      relative humidity at 100%, grow into 1 to 10  fim fog droplets. Sulfur dioxide dissolves  in
27      the fog droplets and is rapidly  oxidized to sulfate by atmospheric oxidants such  as H2O2 or
28      O3,  or by catalysis by Fe or Mn.  These particles lose some of their water as the relative
29      humidity decreases below 100% RH, but will have substantially more  sulfate than prior to
30      activation.  Similar processes occur in clouds  (Swartz, 1984).
        April 1995                               3-164      DRAFT-DO NOT QUOTE OR CITE

-------
             50
    40

E
(O

S  30
             2°
             10
              0
                  D Poor Period, 8/12-8/16
                  • 0000 - 0800
                                    a
                                                                D
                                                                  D
                                         on
                                      n
                                             n
                                                   n
                       o
                      o°-
                  o    o
                  o
                                     o
                                                             O Fair Period, 8/17-8/20
                                                             •0000-0800
                             20
                                 40            60
                                Relative Humidity (%)
               80
100
      Figure 3-33.
             Relative humidity versus sulfur, during the 1986 Carbonaceous Species
             Methods Comparison Study, for particles with Dp>0.56 /im. The
             approximate trajectories followed during each day by the Dp > 0.56
             sulfur size fraction are shown for period P and period F.  Note that
             even when the humidities are low, 30 to 50 %, the period P aerosols
             remain coarser by a factor of three than those of period F.  The water
             content incorporated in the aerosols during the 0000- to 0800-h time
             periods is lost only slowly, giving a strong hysteresis effect in sulfur
             size.
      Source:  Cahill et al., 1990.



1          In an analysis of data from the IMPROVE network Cahill and co-workers (Eldred et

2     al., 1995) report that 20% of the total sulfate is found in the coarse fraction of PM10.

3     Studies in Philadelphia using dichotomous samplers have also reported that 20% of the total

4     sulfate  is found in the coarse fraction (Dzubay et al., 1988).  Cahill and coworkers suggest
      April 1995
                                   3-165
DRAFT-DO NOT QUOTE OR CITE

-------
              CD


              £
35

30

25

20

15

10

 5

 0
                             Summer  Atl Sites  SO*'
                                                          (a)
   0.1                   1
          Aerodynamic Mode Diameter
                                                            10

               400
               300
            o
            §
            O  200
            (D
            Summer  All Sites  SOn •
         (b)
                    0.1                  1
                           Aerodynamic Mode Diameter (urn)
                                           10
Figure 3-34.     Data from the South Coast Air Quality Study (John et al., 1990).
                Plots show (a) frequency of sulfate modes of various sizes as a function
                of mode mass mean diameter (MMD) and (b) average sulfate mode
                concentration as a function of mode MMD.  Note that although there
                are only a few instances when the MMD is near 1.0 /an diameter, it is
                these situations that give rise to the highest sulfate concentrations.
                Modes with MMDs above 2.5 /im diameter may be due to collection of
                fog droplets containing sulfate or reaction of SO2 in liquid droplets of
                NaCl due to NaCl sea spray droplets in which SO2 has dissolved and
                reacted to form sulfate and release HC1 gas.
April 1995
                     3-166
DRAFT-DO NOT QUOTE OR CITE

-------
                    1,000 -
                 <0

                 ^
                 0)
                 •-

                 I
                 o
                 O
                 o>
      Figure 3-35.
   0.1                                     1
        Aerodynamic Mode Diameter (jim)

Log-log plot of sulfate mode concentration versus mode diameter from
Claremont during summer SCAQS (John et al., 1990).  The solid lines
have slopes corresponding to mode concentration increasing with the
cube of the mode diameter. A transition between the two modes is
believed to occur at approximately the sulfate mode concentration
indicated by the horizontal dashed line.
1     that sulfate particles may grow larger than 2.5 pm in diameter and thus be sampled in the

2     PM10 fraction but not the PM2 5  fraction.  It is possible for SO2 to react with basic

3     carbonatecoarse particles to form a sulfate coating or to dissolve in wet NaCl particles, from

4     oceans, lakes, or salt placed on streets to dissolve ice, and be converted to sulfate with the

5     release of HC1.  However,  there is substantial evidence that some  fine sulfate, and therefore

6     possibly other fine mode material, may be found in the size range above 1.0 p,m and even

7     above 2.5 pm diameter, due to the growth of hygroscopic  particles at very high relative

8     humidity.
      April 1995
                       3-167
DRAFT-DO NOT QUOTE OR CITE

-------
1
2
3
4
5
6
7
     A similar process probably accounts for the large size of the fine mode observed in
Vienna (Berner et al., 1993).  Winter and summer size distribution are shown in Figure 3-36.
Berner et al. reported that fog occurred during the night time during the winter study.  In
this European study, as in American studies, instances of fine mode size distributions with
MMDs near or above 1 ^m seem to occur only when fog or very high relative humidity
conditions have been present.
               40.0
             Q.
            Q
            O)
            .o
            TJ
         20.0"
                                          Vienna, Summer
                                     Aerodynamic Diameter, Dp (^
               50.0
             Q.
            Q
             CD
            _O
            T3
         25.0-
                                            Vienna, Winter
                                                                            10
                                      Aerodynamic Diameter, Dp (urn)
       Figure 3-36.    Typical results of size-distribution measurements taken with a Berner
                      impactor in a Vienna street with heavy automotive traffic (Berner et al.,
                      1993):  (a) measurements taken during summer at three different
                      elevations,  (b) measurements taken during winter at three different
                      elevations,  fog was frequently present during the winter sampling
                      period.
       April 1995
                                        3-168
DRAFT-DO NOT QUOTE OR CITE

-------
  1      3.7.1.7.  Conclusions
  2           This review of atmospheric particle-size-distributions was undertaken to provide
  3      information which could be used to determine what cut-point; 1.0 /urn, 2.5 pirn, or something
  4      in between; would give the best separation between the fine and coarse particle modes. The
  5      data do not provide a clear or obvious answer. Depending on conditions, a significant
  6      amount of either fine or coarse mode material may be found in the intermodal region
  7      between 1.0 and 2.5 pun.  However, the analysis does demonstrate the important role of
  8      relative humidity in influencing the size of the fine particle mode and indicates that
  9      significant fine mode material is found above 1.0 ptm only during periods of very high
10      relative humidity.
11           Thus,  a PM2.5 pirn sample will contain most of the fine mode material, except during
12      periods of RH near 100 %.  However, especially in conditions of low RH, it may contain
13      5 to 20 % of the coarse mode material below 10 />im in diameter.   A PM1.0 /an  sample will
14      prevent misclassification of coarse mode material as fine but under high RH  conditions will
15      result in some of the fine mode material being misclassified as coarse.
16           A reduction in RH, either intentionally or inadvertently, will reduce the size of the fine
17      mode.  A sufficient reduction in RH will yield a dry fine particle mode with very little
18      material above 1.0 /mi.  However, reducing the RH by heating will result in loss of
19      semivolatile components such as ammonium nitrate and semivolatile organic  compounds.  No
20      information was found on techniques designed to remove particle-bound water without loss of
21      other semivolatile components.
22
23
24      3.8   SUMMARY
25         •  The  scales of transport of most interest with respect to atmospheric aerosols are the
26            13- and a-mesoscales, which have been emphasized in this section.
27
28         •  Field measurements and model evaluations of transport and dispersion on these scales
29            are quite limited.  Quantitative simulations of transport processes on these scales are
30            believed to be subject to considerable error.
31
32         •  The  main sources of this error are believed to be related to sub-grid-scale processes
33            related to plumes, clouds, complex terrain and complex mesoscale flow systems.


        April 1995                               3-169      DRAFT-DO NOT QUOTE OR CITE

-------
 1            Routine surface meteorological measurements are extensive enough, but upper level
 2            measurements are too coarse.
 3
 4         •  New multi-scale models currently under development are expected to provide
 5            substantially finer spatial resolution where most needed, as well as special treatment
 6            of some of the remaining sub-grid-scale effects (e.g., plume-in-grid treatment for
 7            major point source emissions).  Transport simulation should be much improved in
 8            these new models.
 9
10         •  Meteorology  influences paniculate air quality in a variety of other ways also (Sloane,
11            1983). Over a period of a week or so, air masses stagnating in high pressure cells
12            over source regions may develop considerable haze which may then be transported
13            over a long range (Gillani and Husar, 1976; Samson, 1977).  Also, UV radiation,
14            temperature and humidity are well known to be important factors  in atmospheric
15            chemistry and aerosol formation.
16
17         •  The primary focus in this section has been on field measurements  related  to two
18            important transformation processed:  gas-to-particle conversation (particle formation),
19            and growth of hygroscopic particles  in humid air, clouds, and fogs.
20
21         •  Another major focus has been on transformations in plumes of major point sources
22            and urban-industrial complexes which are the carriers of most particles of
23            anthropogenic origin.
24
25         •  Field measurements of homogeneous gas-phase and heterogenous  aqueous-phase
26            chemistry are reviewed for sulfur, nitrogen, and organic compound.
27
28         •  Gas-phase chemistry in (NOX) plumes depends principally on plume dilution with
29            background hydrocarbons and oxidants, and on sunlight. Large diurnal and  seasonal
30            variations exist in the rates of oxidation of SO2 to sulfate and NO2 to inorganic
31            nitrate. For SO2, the gas-phase rate in diluted point-source plumes varies typically
32            between  1 and 3% h"1 during summer midday conditions in the eastern United States,
33            and up to about  1% h"1 in the cleaner conditions of southwestern  United States.  In
34            urban plumes, the upper limit appears to be closer t 5% h"1 under the more polluted
35            conditions. For NO2, the rates  appear to be about three times faster for both types of
36            plumes.  Winter rates are about an order of magnitude lower, on  average.
37            Conversion rates in the background are comparable to the peak rates in diluted
38            plumes.  Neutralization of H2SO4 formed from SO2 oxidation increases with plume
39            age and background  NH3 concentration.  If the NH3 concentration is more than
40            sufficient to completely neutralize H2SO4 to (NH4)2SO4, then some of the HNO3 may
41            be  converted  to NH4NO3, depending also on temperature.  NH4NO3 is observed
42            commonly in summer in Los Angeles at  Riverside,  downwind of  a major source of
43            ammonia.
44
45         •  Contributions of aqueous-phase  chemistry in plumes are highly variable, depending
46            on  availability of the aqueous phase  (wetted aerosols, clouds, fogs,  and light rain) and
47            the photochemically  generated oxidizing  agents (particularly H2O2 in the  case of

        April 1995                              3-170      DRAFT-DO NOT QUOTE OR CITE

-------
  1             sulfur).  In-cloud conversion rates for sulfur can be several fold larger than the
  2             underlying gas-phase rates, and both are low in stable and low-sunlight conditions.
  3             Nitrate in the aqueous phase is due mainly to dissolution of nitric acid formed in the
  4             gas phase during daylight; there are indications of its formation under dark conditions
  5             by heterogeneous processes involving the NO3  radical and N2O5.
  6
  7         •   Variable amounts of secondary organic aerosols have been observed in urban smog,
  8             particularly in the Los Angeles basin. They are most common on summer days at
  9             downwind sites, and have been observed to comprise as much as 70% of the total
10             OC. More typically, however, primary OC is  dominant.
11
12         •   There is considerable recent evidence suggesting that aerosol composition is
13             externally mixed, at least partially, with "more" hygroscopic and "less" hygroscopic
14             components co-existing in monodisperse populations. This observation has important
15             implications concerning the water content of atmospheric particles, and their growth
16             in humid conditions.
17
18         •   One- to  three-quarters of the aerosol  mass in the eastern United States atmosphere is
19             estimated to be water-soluble.  At RH >  75%, the water content of such aerosol in
20             creases with RH.  Also, the critical supersaturation for activation of such aerosol is
21             relatively insensitive to its chemical composition.  The accumulation mode of such
22             aerosol is expected to be  fully activated in convective clouds, but evidently only
23             partially in stratiform clouds and ground fogs, depending on particle dry size,
24             concentration,  and cooling rate.  In stratiform clouds under polluted  conditions,
25             activation efficiency decreases non-linearly with increasing particle concentration.
26
27
       April 1995                               3_171      DRAFT-DO NOT QUOTE OR CITE

-------
 1      REFERENCES
 2
 3      Ahr, M.; Flossman, A. I.; Pruppacher, H. R. (1989) A comparison between two formulations for nucleation
 4          scavenging. Contrib. Atmos. Phys. 62: 321-326.
 5
 6      Air Quality Group. (1993) Anual report: aerosol collection and compositional analysis for improvement. Davis,
 7          CA: University of California at Davis.
 8
 9      Akhter, M. S.; Chughtai, A. R.; Smith, D. M. (1984) J. Phys. Chem. 88: 5334-5342.
10
11      Akhter, M. S.; Chughtai, A. R.; Smith, D. M. (1985) Appl. Spectrosc. 39: 143-153.
12
13      Alheit, R. R.; Flossman, A.  I.;  Pruppacher, H.  R. (1990) A theoretical study of the wet removal of
14          atmospheric pollutants—Part IV: the uptake and redistribution of aerosol particles through nucleation and
15          impaction scavenging by growing cloud drops and ice particles. J. Atmos. Sci. 47: 870-887.
16
17      Alkezweeny, A.;  Powell, D. (1977) Estimation of transformation rate of SO2 to  SO4 from atmospheric
18          concentration data. Atmos. Environ.  11: 179-182.
19
20      Allen, D. T.; Palen, E. J.; Haimov, M. I.; Hering, S. V.; Young, J. R. (1994) Fourier transform infrared
21          spectroscopy  of aerosol collected in a low pressure impactor (LPI / FTIR): method development and field
22          calibration. Aerosol Sci. Technol. 21: 325-342.
23
24      Allwine, K. J. (1993) Atmospheric dispersion and tracer ventilation in a deep mountain valley. J. Appl.
25          Meteorol. 32: 1017-1037.
26
27      Alofs, D. J.; Hagen, D. E.;  Trueblood, M. B. (1989) Measured spectra of the hygroscopic fraction of
28          atmospheric aerosol particles. J. Appl.  Meteorol. 28: 126-136.
29
30      Altshuller, A. P.  (1983) Review: natural volatile organic substances and their effect on air quality in the United
31          States. Atmos. Environ.  17: 2131-2165.
32
33      Altshuller, A. P.  (1987) Potential contribution of sulfate production in cumulus cloud droplets to ground level
34          particle sulfur concentrations. Atmos. Environ. 21: 1097-1105.
35
36      American Conference of Governmental Industrial Hygienists. (1994) Appendix D: particle size-selective
37          sampling criteria for airborne paniculate matter. Cincinnati, OH: American  Conference of Governmental
38          Industrial Hygienists.  Threshold limit values for chemical substances and physical agents and biological
39          exposure indices; pp.  43-46.
40
41       Anderson, J. R.;  Aggett, F.  J.;  Buseck, P. R.; Germani, M. S.; Shattuck, T. W. (1988) Chemistry of
42          individual aerosol particles  from Chandler, Arizona, an  arid urban environment. Environ. Sci. Technol.
43          22: 811-818.
44
45       Andreae, M. O.; Charlson, R. J.; Bruynseels, F.; Storms, H.;  Van Grieken, R.; Maenhaut, W. (1986) Internal
46          mixture  of sea salt, silicates, and excess sulfate in marine aerosols. Science  (Washington, DC)
47          232: 1620-1623.
48
49       Andrews, E.; Larson, S. M. (1993) Effect of surfactant layers on the size changes of aerosol particles as a
50           function of relative humidity. Environ. Sci. Technol. 27: 857-865.
51
52       Appel, B.  R.; Tokiwa, Y.; Hsu, J.; Kothny, E. L.; Hahn, E. (1985) Visibility as related to atmospheric aerosol
53           constituents.  Atmos. Environ. 19: 1525-1534.
54

         April  1995                                    3-172       DRAFT-DO  NOT QUOTE  OR CITE

-------
 1       Appel, B. R.; Cheng, W.; Salaymeh, F. (1989) Sampling of carbonaceous particles in the atmosphere—II.
 2          Atmos. Environ. 23: 2167-2175.
 3
 4       Arimoto, R.; Duce, R. A.; Ray, B. J.; Unni, C. K. (1985)  Atmospheric trace elements at Enewetak Atoll.
 5          2. Transport to the ocean by wet and dry deposition. J.  Geophys. Res. 90: 2391-2408.
 6
 7       Arnts, R. R.; Gay,  B. W., Jr. (1979) Photochemistry of some naturally emitted hydrocarbons. Research
 8          Triangle Park, NC: U.S. Environmental Protection Agency, Environmental Sciences  Research Laboratory;
 9          EPA report no. EPA-600/3-79-081. Available from: NTIS, Springfield, VA;  PB80-131980.
10
11       Ashbaugh, L. L.; Myrup, L. O.; Flocchini, R.  G.  (1984) A principal component analysis of sulfur
12          concentrations in the western United States. Atmos. Environ. 18: 783-791.
13
14       Ayers, G.  P.; Larson, T. V. (1990)  Numerical study of droplet size dependent chemistry in oceanic, wintertime
15          stratus clouds at southern mid-latitudes. J. Atmos. Chem.  11: 143-167.
16
17       Bagnold, R.  A. (1941) The physics of blown  sand and desert dunes. London,  United Kingdom: Methuen & Co.
18
19       Baker, M. B.; Latham, J. (1979) The evolution of droplet spectra and the rate of production of embryonic
20          raindrops in small cumulus clouds. J. Atmos. Sci. 36: 1612-1615.
21
22       Baldwin, A.  C. (1982) Heterogeneous reactions of  sulfur dioxide with carbonaceous particles. Int. J. Chem.
23          Kinet. 14: 269-277.
24
25       Barnard, W. R.; Stensland, G. J.; Gatz, D. F. (1986)  Alkaline materials flux from unpaved roads: source
26          strength, chemistry and potential for acid rain neutralization. Water Air Soil Pollut. 30: 285-293.
27
28       Barnard, W. R.; Gatz, D. F.; Stensland, G. J. (1987)  Chemical characterization of aerosols emitted from
29          vehicle traffic on unpaved roads. Presented at:  the 80th annual meeting of the Air Pollution Control
30          Association; June; New  York, NY.
31
32       Barnard, W. R.; Stansland, G. J.; Gatz, D. F. (1988)  Flux  of alkaline materials from unpaved roads in the
33          southwestern United States. In: Particulate matter/fugitive dusts:  measurement and control in western arid
34          regions. Proceedings of an Air Pollution Control Association meeting. Pittsburgh, PA: Air Pollution
35          Control Association; pp. 27-39.
36
37       Barrie, L.  A. (1985) Scavenging ratios, wet deposition and in-cloud oxidation: an application to the oxides of
38          sulfur and nitrogen. J. Geophys.  Res. 90: 5789-5799.
39
40       Barrie, L.  A.; Georgii, H. W. (1976) An experimental investigation of the absorption of sulphur dioxide by
41          water drops containing heavy metal ions.  Atmos. Environ. 10: 743-749.
42
43       Barth, M.  C. (1994) Numerical modeling of sulfur and nitrogen chemistry in a narrow cold frontal rainband:
44          the impact of meteorological  and chemical parameters. J. Appl. Meteorol.  33: 855-868.
45
46       Barth, M.  C.; Hegg, D. A.; Hobbs, P. V.  (1992) Numerical modeling of cloud and precipitationchemistry
47          associated with two rainbands and some comparisons with observations. J. Geophys.  Res. 97: 5825-5845.
48
49       Bassett, M. E. (1990) Episodic PM10 model development and application for the South Coast air basin.
50          El Monte, CA: South Coast Air Quality Management District; draft technical report no.  V-E.
51
52       Bassett, M.; Seinfeld, J. H. (1983) Atmospheric equilibrium model of sulfate and nitrate  aerosols. Atmos.
53          Environ.  17: 2237-2252.
54


         April 1995                                    3473       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Bassett, M. E.;  Seinfeld, J. H. (1984) Atmospheric equilibrium model of sulfate and nitrate aerosols—II.
  2           particle size analysis. Atmos. Environ. 18: 1163-1170.
  3
  4       Batterman, S. A.; Dzubay, T. G.; Baumgardner, R. E. (1988) Development of crustal profiles for receptor
  5           modeling. Atmos. Environ. 22:  1821-1828.
  6
  7       Benkovitz, C. M.; Berkowitz, C.  M.; Easter,  R. C.; Nemesure, S.; Wagener, R.; Schwartz, S. E. (1994)
  8           Sulfate burdens over the North Atlantic and adjacent continental regions: evaluations for October and
  9           November,  1986 using a  three dimensional model driven by observation-derived meteorology. J. Geophys.
10           Res [Atmos.] 99: 20725-20756.
11
12       Bennett, R. L.;  Stockburger,  L.; Barnes, H. M. (1994) Comparison of sulfur measurements from a regional
13           fine particle network with concurrent acid modes network results. Atmos. Environ. 28: 409-419.
14
15       Berner, A. (1989) Haze and its relation to atmospheric accumulation aerosol. Sci. Total Environ. 86: 251-263.
16
17       Berner, A.; Liirzer, C.; Pohl, F.; Preining, O.; Wagner, P. (1979) The size distribution of the urban aerosol in
18           Vienna. Sci. Total Environ. 13: 245-261.
19
20       Bernstein, D. M.; Rahn, K. A. (1979) New York Summer Aerosol Study: trace element concentrations as a
21           function of particle size. In: Kneip, T.  J.; Lippmann, M., eds. The New York Summer  Aerosol Study,
22           1976. Ann.  N. Y. Acad.  Sci. 322: 87-97.
23
24       Bidleman, T. F. (1988) Atmospheric processes. Environ. Sci.  Technol. 22: 361-367.
25
26       Biggins, P. D. E.; Harrison, R. M.  (1980) Chemical speciation of lead compounds in street  dusts. Environ. Sci.
27           Technol. 14: 336-339.
28
29       Binkowski, F. S.; Shankar, U. (1994) The  regional paniculate model, part I: model description and preliminary
30           results. Research Triangle Park, NC: U.S. Environmental Protection Agency.
31
32       Biswas, P.; Jones, C. L.; Flagan, R. C. (1987) Distortion of size distributions by condensation and evaporation
33           in aerosol instruments. Aerosol Sci. Technol. 7: 231-246.
34
35       Bott, A. (1991)  On the influence of the physicochemical proerties of aerosol on the life cycle of radiation fogs.
36           Boundary Layer  Meteorol. 56: 1-31.
37
38       Bott, A.; Carmichael, G. R. (1993) Multiphase chemistry in a microphysical radiation fog model-A numerical
39           study. Atmos. Environ. 27: 503-522.
40
41       Bower, K. N.; Choularton, T. W. (1993) Cloud processing of the cloud condensation nucleus spectrum and its
42           climatological consequences.  Q. J. R. Meteorol. Soc. 119: 655-679.
43
44       Bower, K. N.; Hill, T. A.; Coe,  H.; Choularton, T. W.  (1991) SO2 oxidation in an entraining cloud model
45           with explicit microphysics. Atmos. Environ. Part A 25: 2401-2418.
46
47       Boyce,  S. D.; Hoffmann, M.  R. (1984) Kinetics and mechanism of the formation of hydroxymethanesulfonic
48           acid at low pH. J. Phys. Chem.  88: 4740-4746.
49
50       Braaten, D. A.;  Paw U, K. T. (1992) A stochastic  particle resuspension and deposition model. In: Schwartz,
51           S.; Slinn, W. G.  N., eds. Precipitation scavenging and atmosphere-surface exchange, v. 2. New York, NY:
52           Hemisphere Publishing Corp.; pp. 1143-1152.
53
         April 1995                                    3-174       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Braaten, D. A.; Paw U, K. T.; Shaw, R, H. (1989) Particles resuspension in a turbulent boundary layer:
  2          observed and modeled. J. Aerosol Sci. 12: 613-628.
  3
  4      Briggs, G. A.; Binkowski, F. S. (1985) Research on diffusion in atmospheric boundary layers: a position paper
  5          on status and needs. Research Triangle Park, NC: U.S. Environmental Protection Agency, Atmospheric
  6          Sciences Research  Laboratory; EPA report no. EPA/600/3-85/072. Available from: NTIS, Springfield, VA;
  7          PB86-122587.
  8
  9      Brown, N. J.; Dod, R. L.; Mowrer, F. W.; Novakov, T.; Williamson, R. B. (1989) Smoke emission factors
 10          from medium-scale fires: part 1. Aerosol Sci. Technol. 10: 2-19.
 11
 12      Buat-Menard, P.; Duce, R. A. (1986)  Precipitation scavenging of aerosol particles over remote marine regions.
 13          Nature (London) 321: 508-510.
 14
 15      Burkhard, E. G.; Dutkiewicz, V. A.; Husain, L. (1994) A study of SO2,  S02"4 and trace elements in clear air
 16          and clouds above the midwestern United States. Atmos. Environ. 28: 1521-1533.
 17
 18      Burnett, R. T.; Dales,  R.  E.; Raizenne, M.  E.; Krewski, D.; Summers, P. W.; Roberts, G. R.;  Raad-Young,
 19          M.; Dann, T.;  Brook, J. (1994) Effects of low ambient levels of ozone and sulfates on the frequency of
 20          respiratory admissions to Ontario hospitals. Environ. Res. 65: 172-194.
 21
 22      Burton, R. M.; Lundgren, D. A. (1987) Wide range aerosol classifier: a size selective  sampler for large
 23          particles. Aerosol Sci. Technol. 6: 289-301.
 24
 25      Burtscher, H. (1992) Measurement and characteristics  of combustion aerosols with special consideration of
 26          photoelectric charging and charging by flame ions. J. Aerosol Sci. 23: 549-595.
 27
 28      Burtscher, H.; Leonardi,  A.; Steiner, D.; Baltensperger, U.; Weber, A. (1993) Aging  of combustion particles
 29          in the atmosphere:  results from a field study in Zurich. Water Air Soil Pollut. 68:  137-147.
 30
 31      Byers, H. R. (1974) General meteorology. New York, NY: McGraw Hill.
 32
 33      Cadle, S.  H.; Dasch, J. M. (1988) Wintertime concentrations and sinks of atmospheric paniculate carbon at a
 34          rural  location in northern Michigan.  Atmos.  Environ. 22: 1373-1381.
 35
 36      Cahill, T. A.; Surovik, M.; Wittmeyer, I. (1990) Visibility and aerosols during the 1986 Carbonaceous Species
 37          Methods Comparison Study. Aerosol Sci. Technol. 12: 149-160.
 38
 39      Calvert, J. G.;  Su,  F.;  Bottenheim, J. W.; Strausz, O. P. (1978) Mechanism of the homogeneous oxidation of
40          sulfur dioxide in the troposphere. In: Sulfur in the atmosphere: proceedings of the international symposium;
41          September  1977; Dubrovnik, Yugoslavia. Atmos. Environ.  12: 197-226.
42
43      Calvert, J. G.;  Lazrus,  A.; Kok, G.; Heikes, B.; Walega, J.; Cantrell, C. A. (1985) Chemical mechanisms of
44          acid generation in the troposphere. Nature (London) 317: 27-38.
45
46      Cambray, R. S. (1989) Radioactive fallout in air and rain: results to the end of 1987. Harwell, United
47          Kingdom: Atomic Energy Establishment; report no. AERE-R 13226.
48
49      Cambray,  R. S.; Cawse, P. A.; Garland, J.  A.; Gibson, J. A. B.; Johnson, P.; Lewis, G. N. J.; Newton, D.;
50          Salmon, L.; Wade, B. O. (1987) Observations on  radioactivity from the  Chernobyl accident. Nucl. Energy
51          (Br. Nucl. Energy Soc.) 26: 77-101.
52
53      Carpenter, S. B.;  Montgomery,  T. L.; Leavitt, J. M.;  Colbaugh, W.  C.;  Thomas, F. W. (1971)  Principal
54          plume dispersion models: TVA power plants. J. Air Pollut. Control Assoc. 21: 491-495.


         April 1995                                    3.175      DRAFT-DO NOT QUOTE OR CITE

-------
  1       Carras, J. N.; Williams, D. J. (1981) The long-range dispersion of a plume from an isolated point source.
  2           Atmos. Environ. 15: 2205-2217.
  3
  4       Carras, J. N.; Williams, D. J. (1988) Measurements of relative a  up to 1800 km from a single source. Atmos.
  5           Environ. 22: 1061-1069.
  6
  7       Carter, E. J.; Borys, R. D. (1993) Aerosol cloud chemical fractionation: enrichment factor analysis of cloud
  8           water. J. Atmos. Chem. 17: 277-292.
  9
10       Cass,  G. R. (1977) Methods for sulfate air quality management with applications to Los Angeles
11           [Ph.D. dissertation]. Pasadena, CA: California Institute of Technology; EQL report no. 16-2.
12
13       Cass,  G. R. (1979) On the relationship between sulfate air quality and visibility with examples in Los Angeles.
14           Atmos. Environ. 13: 1069-1084.
15
16       Cass,  G. R.; Shair, F. H. (1984) Sulfate accumulation in a sea breeze/land breeze circulation  system.
17           J. Geophys. Res. 89: 1429-1438.
18
19       Cass,  G. R.; Boone, P. M.; Macias, E. S.  (1982) In: Wolff,  G.  T.; Klimisch, R. L., eds. Particulate carbon:
20           atmospheric life cycle. New York, NY: Plenum Press; pp. 207-240.
21
22       Cawse, P. A. (1981) Trace elements in the atmosphere of the United Kingdom. Presented at:  ESNA meeting;
23           September-October; Aberdeen, Scotland.
24
25       Chamberlain, A. C. (1983) Deposition and resuspension. In:  Pruppacher et al., eds. Precipitation scavenging,
26           dry deposition, and resuspension, v. 2. New York, NY:  Elsevier Science Publishing Co., Inc.;
27           pp. 731-751.
28
29       Chameides,  W. L. (1984) The photochemistry of a remote stratiform cloud. J. Geophys. Res. [Atmos.]
30           89: 4739-4755.
31
32       Chameides,  W. L.; Davis, D. D. (1982) The free radical chemistry of cloud droplets and its impact upon the
33           composition of rain. J.  Geophys. Res. C: Oceans Atmos. 87: 4863-4877.
34
35       Chan, W. H.; Tang Al, J. S.; Chung, D. H.  S.; Lusis, M. A. (1986) Concentration and deposition of trace
36           metals in Ontario, 1982. Water Air Soil Pollut. 29: 373-389.
37
38       Chang, S. G.; Brodzinsky, R.; Gundel, L.  A.; Novakov, T.  (1982) Chemical and catalytic properties of
39           elemental carbon. In: Wolff, G.  T.; Klimisch, R. L.,  eds. Particulate carbon: atmospheric life cycle.
40           New York, NY: Plenum Press; pp. 158-181.
41
42       Chang, J. S.; Middleton, P. B.; Stockwell, W. R.; Walcek, C. J.;  Pleim, J.  E.; Lansford, H. H.; Madronich,
43           S.; Binkowski, F. S.; Seaman, N.  L.; Stauffer, D. R.  (1991) The regional acid deposition model and
44           engineering model. In:  Irving, P. M., ed. Acidic deposition: state of science and technology, volume I:
45           emissions, atmospheric processes, and deposition. Washington, DC: The U.S. National Acid Precipitation
46           Assessment Program. (State of science and technology report no. 4).
47
48       Charlson, R. J.; Lovelock, J.  E.; Andreae, M. O.; Warren, S. G.  (1987) Oceanic phytoplankton, atmospheric
49           sulfphur cloud, albedo, and climate. Nature (London) 326: 655-661.
50
51       Cheng, Y.-S.; Carpenter, R. L.; Barr, E. B.; Hobbs, C. H. (1985) Size distribution of fine particle emissions
52           from a steam plant with a fluidized-bed coal combustor. Aerosol Sci. Technol. 4: 175-189.
53
         April 1995                                    3-176       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Cheng, M.-D.; Lioy, P. J.; Opperman, A. J. (1988) Resolving PM10 data collected in New Jersey by various
  2          multivariate analysis techniques. In: Mathai, C. V.; Stonefield, D. H., eds. Transactions: PM10
  3          implementation of standards. Pittsburgh, PA: Air Pollution Control Association; pp. 472-483.
  4
  5       Ching, J. K. S.; Alkezweeny, A. J. (1986) Tracer study of vertical exchange by cumulus clouds. J. Clim.
  6          Appl. Meteorol. 25: 1702-1711.
  7
  8       Choularton, T.  W.; Wicks, A. J.; Downer,  R. M.; Gallagher, M. W.; Penkett, S. A.; Bandy, B. J.; Dollard,
  9          G. J.; Jones, B. M. R.; Davies, T. D.; Gay, M. J.; Tyler, B. J.; Fowler,  D.; Cape, J. N.; Hargreaves,
 10          K. J. (1992) A field study of the generation  of nitrate in a hill cap cloud. Environ. Pollut. 75: 69-73.
 11
 12       Chow, J. C.;  Watson, J. G.; Lowenthal, D. H.; Solomon, P. A.; Magliano, K. L.; Ziman, S. D.;  Richards,
 13          L. W.  (1992) PM10 source apportionment in California's San Jaoquin Valley. Atmos. Environ.  Part A
 14          26: 3335-3354.
 15
 16       Chow, J. C.;  Watson, J. G.; Lowenthal, D. H.;  Solomon, P. A.; Magliano, K. L.; Ziman, S. D.;  Richards,
 17          L. W.  (1993) PM10 and PM2 s compositions in California's San Joaquin Valley. Aerosol Sci. Technol.
 18          18: 105-128.
 19
 20       Chow, J. C.;  Watson, J. G.; Fujita, E. M.; Lu,  Z.; Lawson, D. R.; Ashbaugh, L.  L. (1994) Temporal and
 21          spatial  variations  of PM2 5 and PM10 aerosol in the Southern California Air Quality Study. Atmos.  Environ.
 22          28: 2061-2080.
 23
 24       Chughtai, A. R.; Jassim, J. A.; Peterson, J. H.;  Stedman,  D. H.; Smith, D. M. (1991) Spectroscopic and
 25          solubility characteristics of oxidized soots. Aerosol  Sci. Technol. 15: 112-126.
 26
 27       Clark, T. L.;  Cohn, R. D. (1990) The Across North America Tracer Experiment (ANATEX) model evaluation
 28          study. Research Triangle Park, NC: U.S. Environmental Protection Agency, Atmospheric Research and
 29          Exposure Assessment Laboratory;  EPA  report no. EPA/600/3-90/051. Available from: NTIS, Springfield,
 30          VA; PB90-261454.
 31
 32      Clark, P.  A.;  Fletcher,  I.  S.; Kallend, A. S.; McElroy, W. J.; Marsh,  A.  R. W.; Webb, A. H. (1984)
 33          Observations of cloud chemistry during  long-range transport of power plant plumes.  Atmos. Environ.
 34          18: 1849-1858.
 35
 36      Clarke, A.  D. (1989)  Aerosol light absorption by soot in remote environments.  Aerosol Sci. Technol.
 37          10: 161-171.
 38
 39      Clarke, A. D. (1992)  Atmospheric nuclei in the remote  free-troposphere. J. Atmos.  Chem. 14: 479-488
 40
 41       Clarke, A. G.; Karani, G. N. (1992) Characterization of the carbon content of atmospheric aerosols. J. Atmos.
 42          Chem.  14: 119-128.
 43
 44       Clarke, A. G.; Radojevic, M. (1987) Oxidation of SO2 in rainwater and its role in acid rain chemistry.  Atmos.
 45           Environ. 21: 1115-1123.
 46
47       Clarke, J. F.;  Clark, T. L.; Ching, J. K. S.; Haagenson, P. L.; Husar, R. B.;  Patterson, D. E. (1983)
48           Assessment of model simulation of long-distance transport. Atmos.  Environ. 17: 2449-2462
49
50       Clarke, A. D.; Weiss, R. E.; Charlson, R. J. (1984) Elemental carbon  aerosols in the urban, rural and
51           remote-marine troposphere and in the stratosphere: interferences from light  absorption data and
52           consequences regarding radiative transfer. Sci. Total Environ. 36: 97-102.
53
         April 1995                                    3.177       DRAFT-DO NOT QUOTE OR CITE

-------
 1       Cobourn, W. G.; Husar,  R. B.; Husar, J. D. (1978) Continuous in situ monitoring of ambient paniculate sulfur
 2          using flame photometry and thermal analysis. In: Husar, R. B.; Lodge, J. P., Jr.; Moore, D. J., eds.
 3          Sulfur in the atmosphere:  proceedings of the international symposium; September 1977; Dubrovnik,
 4          Yugoslavia. Atmos. Environ.  12: 89-98.
 5
 6       Cocks, A. T.; McElroy, W. L.; Wallis, P. G. (1982) The oxidation of sodium sulphite solutions by hydrogen
 7          peroxide.  Central Electricity Research Laboratories; report no. RD/L/2215N81.
 8
 9       Cofer, W. R.; Stevens, R. K.; Winstead, E. L.; Pinto,  J. P.; Sebacher, D. L.; Abdulraheem, M. Y.; Alsahafi,
10          M.;  Mazurek, M. A.; Rasmussen, R. A. (1992) Kuwaiti oil fires: composition of source smoke. J.
11          Geophys.  Res. 97: 14,521-14,525.
12
13       Colbeck,  L; Appleby, L.; Hardman, E. J.; Harrison, R. M. (1990) The optical properties and morphology of
14          cloud-processed of Denver's ambient paniculate. Atmos. Environ. 21: 527-538.
15
16       Coles, D. G.; Ragaini, R. G.; Ondov, J. M.; Fisher, G. L.; Silberman, D.; Prentice, B. A. (1979) Chemical
17          studies of stack fly ash from a coal-fired power plant. Environ. Sci. Technol. 13: 455-459.
18
19       Collett, J. L.; Oberholzer, B.; Mosimann, L.; Staehelin, J.; Waldrogel, A. (1993a) Contributions of cloud
20          processes to precipitation chemistry in mixed-phase clouds. Water Air Soil Pollut. 68: 43-57.
21
22       Collett, J., Jr.; Oberholzer, B.; Staehelin, J. (1993b) Cloud chemistry in Mt. Rigi Switzerland: dependence on
23          drop size  and relationship to precipitation chemistry. Atmos. Environ. Part A 27: 33-42.
24
25       Colvile,  R. N.; Choularton, T. W.; Gallagher,  M. W.; Wicks, A. J.; Downer, R. M.; Tyler, B. J.;
26          Storeton-West, K. J.; Fowler, D.; Cape, J. N.; Dollard, G. J.; Davies, T. J.; Jones, B. M.  R.;  Penkett,
27          S. A.; Bandy, B. J.; Burgess, R. A. (1994) Observation on Great Dun Fell of the pathways  by which
28          oxides of nitrogen are converted to nitrate.  Atmos.  Environ. 28:  397-408.
29
30       Cooper,  J. A.; Watson, J. G., Jr.  (1980) Receptor oriented methods  of air paniculate source apportionment.
31          J. Air Pollut. Control Assoc.  30: 1116-1125.
32
33       Cotton, W. R.; Anthes, R. A. (1989) Storm and cloud  dynamics. New York, NY: Academic Press, Inc.
34          (Dmowska, R.; Holton, J. R., eds. International geophysics series: v. 44).
35
36       Countess, R. J.; Wolff, G. T.; Cadle, S. H. (1980) The Denver winter aerosol: a comprehensive chemical
37          characterization. J. Air Pollut. Control Assoc. 30:  1194-1200.
38
39       Countess, R. J.; Cadle, S. H.; Groblicki, P. J.; Wolff, G. T. (1981) Chemical analysis of size-segregated
40          samples of Denver's ambient particulate. J. Air Pollut. Control Assoc. 31: 247-252.
41
42       Coutant,  R. W.; Brown,  L.; Chuang, J. C.; Riggin, R. M.; Lewis, R. G. (1988) Phase distribution and artifact
43          formation in ambient air sampling for polynuclear aromatic hydrocarbons. Atmos. Environ.  22: 403-409.
44
45       Coutant,  R. W.; Callahan, P. J.; Chuang, J. C.; Lewis, R. G. (1992) Efficiency of silicone-grease-coated
46          denuders  for collection of polynuclear aromatic hydrocarbons. Atmos. Environ. Part A 26: 2831-2834.
47
48       Covert, D. S.; Heintzenberg, J. (1984) Measurement of the degree of internal/external mixing of hygroscopic
49          compounds and soot in atmospheric aerosols. Sci. Total Environ. 36:  347-352.
50
51       Covert, D. S.; Charlson, R. J.; Ahlquist, N. C. (1972) A study of the relationship of chemical composition and
52          humidity  to light scattering by aerosols. J.  Appl. Meteorol. 11: 968-976.
53
         April 1995                                    3-178      DRAFT-DO NOT QUOTE OR CITE

-------
  1       Covert, D. S.; Heintzenberg, J.; Hansson, H.-C. (1990) Electro-optical detection of external mixtures in
  2          aerosols. Aerosol Sci. Technol. 12: 446-456.
  3
  4       Covert, D. S.; Kapustin,  V. N.; Quinn, P. K.; Bates, T. S. (1992) New particle formation in the marine
  5          boundary layer. J. Geophys. Res. [Atmos.] 97: 20581-20589.
  6
  7       Cowherd, C., Jr.; Axtell, K., Jr.;  Guenther, C. M.; Jutze, G. A. (1974) Development of emission factors for
  8          fugitive dust sources. Research Triangle Park, NC: U.S. Environmental Protection Agency, Office of Air
  9          Quality Planning and Standards; EPA report no. EPA-450/3-74-037. Available from: NTIS, Springfield,
 10          VA; PB-238262.
 11
 12       Currie, L. A.; Stafford, T. W.; Sheffield, A. E.; Klouda, G. A.; Wise, S.  A.; Fletcher, R. A. (1989)
 13          Microchemical and molecular dating. Radiocarbon 31: 448-463.
 14
 15       Currie, L. A.; Sheffield,  A. E.; Riederer, C.; Gordon,  G. E. (1994) Improved atmospheric understanding
 16          through exploratory data analysis and complementary modeling: the urban K-Pb-C system. Atmos. Environ.
 17          28: 1359-1369.
 18
 19       d'Alameida, G. (1989) Desert aerosol: characteristics and effects on climate. In: Leinen,  M.; Sarnthein, M.
 20          Paleoclimatology and paleometeorology: modern and past patterns of global atmospheric transport.
 21          Dordrect, The Netherlands: Kluwer Academic Publishers; pp. 311-338.
 22
 23       Daisey, J. M. (1987) Chemical  composition of inhalable paniculate matter—seasonal and inters ite comparisons.
 24          In: Lioy, P. J.;  Daisey, J.  M., eds. Toxic air pollution: a comprehensive study of non-criteria air
 25          pollutants. Chelsea, MI: Lewis Publishers, Inc.; pp. 45-65.
 26
 27       Danielsen, E. F. (1961) Trajectories: isobaric, isentropic and actual. J. Meteorol. 18:  479-486.
 28
 29       Dasch, J. M.; Cadle, S. H. (1989) Atmospheric carbon  particles in the Detroit urban area: wintertime sources
 30          and sinks. Aerosol Sci. Technol. 10: 236-248.
 31
 32       Dattner, S. L.; Hopke, P. K. (1982) Receptor models applied to contemporary pollution problems.
 33          In: Proceedings: receptor models applied to contemporary pollution problems. Pittsburgh, PA: Air Pollution
 34          Control Association; pp.  1-5.
 35
 36       Daum, P. H. (1988) Processes  determining cloudwater composition: inferences  from field measurements.
 37          In: Unsworth, M. H.; Fowler, D., eds. Acid deposition at high elevation sites. Dordrecht,  The
 38          Netherlands: Kluwer; pp. 139-153.
 39
 40      Daum, P. H.; Kelly, T. J.; Schwartz, S. E.;  Newman, L. (1984) Measurements of the chemical composition of
 41           stratiform clouds. Atmos. Environ. 18: 2671-2684.
 42
 43       Daum, P. H.; Schwartz, S. E.;  Newman, L.  (1984) Acidic and  related constituents in liquid-water clouds.
 44          J.  Geophys. Res. 89:  1447-1458.
45
46       Daum, P. H.; Kelly, T. J.; Strappe, J. W.; Leaitch, W. R.;  Joe, P.; Schemenauer, R. S.; Isaac, G. A.;
47           Anlauf, K. G.; Wiebe, H. A.  (1987) Chemistry and physics of a winter stratus cloud layer: a case study.
48           J. Geophys. Res. 92:  8426-8436.
49
50       Daum, P. H.; Kleinman, L. I.;  Hill, A. J.; Lazrus, A.  L.; Leslie, A. C. D.; Bushess, K.; Boatman, J. (1990)
51           Measurement and interpretation of concentrations of H2O2 and related species in the upper midwest during
52           summer. J. Geophys.  Res.  D: Atmos. 95: 9857-9871.
53
         April 1995                                    3.179       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Daum, P. H.; Al-Sunaid, A.; Busness, K. M.; Hales, J. M.; Mazurek, M. (1993) Studies of the Kuwait oil fire
  2           plume during midsummer 1991. J. Geophys. Res. 98: 16,809-16,827.
  3
  4       Davidson, C. I.; Wu, Y.-L.  (1990) Dry deposition of panicles and vapors. In: Lindberg, S. E.; Page, A. L.;
  5           Norton, S. A., eds. Acidic precipitation: v. 3, sources, deposition, and canopy interactions. New York,
  6           NY: Springer-Verlag; pp. 103-216.
  7
  8       Davis, E. J.; Ray, A. K. (1978) Submicron droplet evaporation in the continuum and non-continuum regimes.
  9           J. Aerosol Sci. 9: 411-422.
10
11       Dayal, P.; Neuroth, G. R.; Lowe, T.P. (1992) Toxic metal characterization of coal-fired power plant
12           emissions. Transactions AWMA 1:  199-210.
13
14       De Bock, L. A.; Van Malderen, H.; Van Grieken, R. E. (1994) Individual aerosol particle composition
15           variations in air masses crossing the North sea. Environ. Sci. Technol. 28:  1513-1520.
16
17       De Valk, J. P. J.  M. M.  (1994) A model for cloud chemistry:  a comparison between model simulations and
18           observations in stratus and cumulus. Atmos. Environ.  28:  1665-1678.
19
20       DeSantis, F.; Allegrini, I. (1992) Heterogeneous reactions of SO2 and NO2 on carbonaceous surfaces. Atmos.
21           Environ. 26:  3061-3064.
22
23       Delumyea, R. D.; Kalivretenos, A.  (1987) Elemental carbon and lead content of fine particles from American
24           and French cities of comparable size and industry, 1985. Atmos. Environ. 21: 1643-1647.
25
26       Dennis, R. L.; McHenry, J.  N.; Barchet, W.  R.; Binkowski, F. S.; Byun, D. W. (1993) Correcting RADM's
27           sulfate underprediction: discovery and correction of model errors and testing the corrections through
28           comparisons against field data. Atmos. Environ. Part A 27: 975-997.
29
30       Dickerson, R. R.; Huffman,  G. J.;  Luke, W.  T.; Nunnermacker, L. J.; Pickering, K. E.; Leslie, A. C.  D.;
31           Lindsey, C. G.;  Slinn, W. G. N.; Kelly, T. J.; Daum, P.  H.; Delany, A. C.; Greenberg, J. P.;
32           Zimmerman,  P. R.; Boatman, J. F.; Ray, J. D.; Stedman, D. H. (1987)  Thunderstorms: an important
33           mechanism in the transport of air pollutants. Science (Washington, DC) 235: 460-465.
34
35       Dickson, R. J.; Wilkinson, J. G.; Bruckman,  L.; Tesche,  T. W. (1991) Conceptual formulation of the
36           SARMAP emissions modeling system. Presented at: the 84th annual meeting of the Air and Waste
37           Management Association; June; Vancouver, BC, Canada.  Pittsburgh, PA: Air and Waste Management
38           Association.
39
40       Dittenhoefer, A. C.;  de Pena, R. G. (1978) A study of production and growth of sulfate particles  in plumes
41           from a coal-fired power plant. Atmos. Environ.  12: 297-306.
42
43       Dlugi, R. (1989) Chemistry and deposition of soot particles in  moist air and fog. Aerosol Sci. Technol.
44           10: 93-105.
45
46       Dlugi, R.; Glisten, H. (1983) The catalytic and photocatalytic activity of coal  fly ashes. Atmos. Environ.
47           17: 1765-1771.
48
49       Dod, R.  L.; Brown, N. J.; Mowrer, F. W.; Novakov, T.; Williamson, R. B. (1989) Smoke emission factors
50           from medium-scale fires: part 2. Aerosol Sci. Technol. 10: 20-21.
51
52       Dodd, J. A.; Ondov, J. M.;  Tuncel, G.; Dzubay, T. G.; Stevens,  R. K.  (1991) Multimodal size spectra of
53           submicrometer particles bearing various elements in rural air. Environ. Sci. Technol. 25: 890-903.
54


         April 1995                                    3-180       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Doyle, G. J.; Tuazon, E. C.; Graham, R. A.; Mischke, T. M.; Winer, A. M.; Pitts, J. N., Jr. (1979)
  2          Simultaneous concentrations of ammonia and nitric acid in a polluted atmosphere and their equilibrium
  3          relationship to paniculate ammonium nitrate. Environ. Sci. Technol. 13: 1416-1419.
  4
  5      Draxler, R. R. (1982) Measuring and modeling the transport and dispersion of Krypton-85 1500 km from a
  6          point source. Atmos. Environ.  16: 2763-2776.
  7
  8      Draxler, R. R. (1984) Diffusion and transport experiments. In: Randerson, D., eds. Atmospheric science and
  9          power production. Washington, DC: U.S. Department of Energy, Office of Science and Technological
 10          Information; pp. 367-422; DOE report no. DOE/TIC-27601.
 11
 12      Draxler, R. R.; Dietz, R.; Lagomarsino, R. J.; Start, G. (1991) Across North America Tracer Experiment
 13          (ANATEX): sampling and analysis. Atmos. Environ. Part A 25: 2815-2836.
 14
 15      Duce, R. A.; Mohnen, V. A.; Zimmerman, P. R.; Grosjean, D.; Cautreels, W.;  Chatfield, R.; Jaenicke, R.;
 16          Ogren, J. A.;  Pellizzari, E. D.; Wallace, G. T. (1983) Organic material in the global troposphere. Rev.
 17          Geophys. Space Phys. 21: 921-952.
 18
 19      Dulac, F.; Buat-Menard, P.; Ezat,  U.; Melki, S.; Bengametti, G. (1989) Atmospheric input of trace metals to
 20          the western Mediterranean: uncertainties in modeling dry deposition from cascade impactor data. Tellus
 21          Ser. B41: 362-378.
 22
 23      Durham, J. L.; Wilson,  W. E.; Ellestad, T. G.; Willeke, K.; Whitby, K. T. (1975) Comparison of volume and
 24          mass distributions for Denver aerosols. Atmos. Environ.  9: 717-722.
 25
 26      Dzubay, T. G.; Mamane, Y. (1989) Use of electron microscopy  data in receptor models for PM-10. Atmos.
 27          Environ. 23: 467-476.
 28
 29      Dzubay, T. G.; Stevens, R. K.; Gordon, G. E.; Olmez, I.; Sheffield, A. E.; Courtney, W. J. (1988)
 30          A composite receptor method applied to Philadelphia aerosol. Environ. Sci. Technol. 22: 46-52.
 31
 32      Eatough, D. J.; Richter, B. E.; Eatough, N. L.; Hansen, L.  D. (1981) Sulfur chemistry in smelter and power
 33          plant plumes in the western U.S. In: White, W. H.; Moore,  D. J.; Lodge, J.  P., Jr. Plumes and visibility:
 34          measurements  and model components: proceedings  of the symposium; November 1980; Grand Canyon
 35          National Park, AZ. Atmos. Environ. 15: 2241-2253.
 36
 37      Eatough, D. J.; Arthur, R. J.;  Eatough, N. L.; Hill, M. W.; Mangelson, N. F.; Richter, B. E.; Hansen,
 38          L. D.; Cooper, J. A. (1984) Rapid conversion of SO2(g) to sulfate in a fog bank. Environ. Sci. Technol.
 39          18: 855-859.
 40
 41      Eatough, D. J.; Sedar, B.; Lewis, L.; Hansen, L. D.; Lewis, E. A.; Farber, R. J. (1989) Determination of
 42          semivolatile organic compounds in particles in the Grand Canyon area. Aerosol Sci. Technol. 10: 438-449.
 43
44      Eatough, D. J.; Wadsworth,  A.; Eatough, D. A.;  Crawford,  J. W.; Hansen, L. D.; Lewis, E. A. (1993) A
45          multiple-system, multi-channel diffusion denuder sampler for the determination of fine-particulate organic
46          material in the atmosphere. Atmos. Environ. Part A 27: 1213-1219.
47
48      Eisele, F. L.; Bradshaw, J. D. (1993) The elusive hydroxyl radical: measuring OH in the atmosphere. Anal.
49          Chem. 65: 927A-939A.
50
51      Eisele, F. L.; Tanner, D. J. (1993) Measurement of the gas phase concentration of H2SO4 and methane sulfonic
52          acid and estimates of H2SO4 production and loss in the atmosphere. J. Geophys. Res. [Atmos.] 98:
53          9001-9010.
54


         April  1995                                    3_18l        DRAFT-DO NOT QUOTE OR CITE

-------
 1       Eisele, F. L.; Mount, G. H.; Fehsenfeld, F. C.; Harder, J.; Marovich, E.; Parrish, D. D.; Roberts, J.;
 2           Trainer, M.; Tanner, D. (1994) Intercomparison of tropospheric OH and ancillary trace gas measurements
 3           at Fritz Peak Observatory, Colorado. J. Geophys. Res. [Atmos.] 99: 18605-18626.
 4
 5       Eisenreich,  S. (1980) Atmospheric input of trace metals to Lake Michigan. Water Air Soil Pollut. 13: 287-301.
 6
 7       Eldering, A.; Cass, G. R. (1994) A source-oriented model for air pollutant effects on visibility. J. Geophys.
 8           Res. [Atmos.]: submitted.
 9
10       Eldering, A.; Cass, G. R.;  Moon, K. C. (1994) An air monitoring network using continuous particle size
11           distribution monitors: connecting pollutant properties to visibility via Mie scattering calculations. Atmos.
12           Environ. 28: 2733-2749.
13
14       Eldred, R. A.; Cahill, T. A.; Flocchini, R. G.  (1994) Composition of PM10 and PM2 5 aerosols in the
15           IMPROVE network. Proceedings of the international specialty conference on aerosol and atmospheric
16           optics: radiative balance and visual air quality, volume A. Air Waste: submitted.
17
18       Er-El, J.; Peskin, R. L. (1981) Relative diffusion of constant-level balloons in the southern hemisphere.
19           J. Atmos. Sci. 38: 2264-2274.
20
21       Erickson, R. E.; Yates, L.  M.; Clark, R. L.; McEwen, D. (1977) The reaction of sulfur dioxide with ozone in
22           water and its possible atmospheric significance. Atmos. Environ.  11: 813-817.
23
24       Facchini, M. C.; Fuzzi, S.; Lind, J. A. (1992) Phase-partitioning and chemical  reactions of
25           low-molecular-weight organic compounds. Tellus Ser.  B 44B: 533-544.
26
27       Falerios, M.; Schild, K.; Sheehan, P.; Paustenbach, D. (1992) Hexavalent chromium: airborne concentrations.
28           J. Air Waste Manage.  Assoc. 42: 40-48.
29
30       Ferber, G. J.; Telegadas, K.; Heffter, J. L.; Dickson, C. R.; Dietz, R. N.; Krey, P. W. (1981) Demonstration
31           of a long-range atmospheric tracer system using perfluorocarbons. Silver Springs, MD: National Oceanic
32           and Atmospheric Administration; NOAA report no. NOAA TM ERL ARL-101.
33
34       Ferber, G. J.; Heffter, J. L.; Draxler, R. R.; Lagomarsino,  R. J.; Thomas, F. L.; Dietz, R. N.; Benkovitz,
35           C. M. (1986) Cross-Appalachian Tracer Experiment (CAPTEX-83) final report. Silver Springs,  MD:
36           National Oceanic and Atmospheric Administration; NOAA report no. NOAA TM ERL ARL-142.
37
38       Fernandez de la Mora, J. (1986) Inertia and interception in the deposition of particles from boundary layers.
39           Aerosol Sci. Technol. 5: 261-266.
40
41       Fernandez de la Mora, J.; Friedlander, S. K. (1982) Int. J. Heat Mass Transfer 25: 1725-1735.
42
43       Finlayson-Pitts, B. J.; Pitts, J. N. (1986) Paniculate matter in atmosphere: primary and secondary particles.
44           In: Finlayson-Pitts, B.  J.; Pitts, J. N., eds. Atmospheric chemistry: fundamentals and experimental
45           techniques. New York, NY: John Wiley and Sons; pp. 783-802.
46
47       Finlayson-Pitts, B. J.; Pitts, J. N., Jr. (1986) Atmospheric chemistry: fundamentals and experimental
48           techniques. New York, NY: John Wiley & Sons.
49
50       Fisher, G. L. (1980) Size-related chemical and physical properties of power plant fly ash.  In: Willeke, K., ed.
51           Generation of aerosols  and facilities for exposure experiments; April 1979;  Honolulu,  HI. Arm Arbor, MI:
52           Ann Arbor Science Publishers, Inc.; pp. 203-214.
53
         April 1995                                     3-182       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Fitzgerald, J. W. (1973) Dependence of the supersaturation spectrum of CCN on aerosol size distribution and
  2          composition. J. Atmos. Sci. 30: 628-634.
  3
  4      Fitzgerald, J. W.; Spyers-Duran, P. A. (1973) Changes in cloud nucleus concentration and cloud droplet size
  5          distribution associated with pollution from St. Louis. J. Appl. Meteorol. 12: 511-516.
  6
  7      Flossman, A. I.; Hall, W. D.; Pruppacher, H. R. (1985) A theoretical study of the wet removal of atmospheric
  8          pollutants—Part I. the redistribution of aerosol particles captured through nucleation and impaction
  9          scavenging by growing cloud drops. J. Atmos. Sci. 42: 583-606.
10
11      Foreman, W. T.; Bidleman, T. F. (1990) Semivolatile organic compounds in the ambient air of Denver,
12          Colorado. Atmos. Environ. Part A 2405-2416.
13
14      Forkel, R.;  Seidl, W.; Dlugi, R.; Deigle, E.  (1990) A one dimensioal numerical model to simulate formation
15          and balance of sulfate during radiation fog events. J. Geophys. Res. 95: 18501-18515.
16
17      Forrest, J.;  Newman, L. (1973) Sampling and analysis of atmospheric sulfur compounds for isotope ratio
18          studies. Atmos. Environ. 7: 561-573.
19
20      Forrest, J.;  Newman, L. (1977a) Further studies on the oxidation of sulfur dioxide in coal-fired power plant
21          plumes. Atmos. Environ. 11: 465-474.
22
23      Forrest, J.;  Newman, L. (1977b) Oxidation of sulfur dioxide in the Sudbury smelter plume.  Atmos.  Environ.
24          11:517-520.
25
26      Forrest, J.;  Garber, R.; Newman, L.  (1979) Formation of  sulfate, ammonium and nitrate in an oil-fired power
27          plant plume. Atmos. Environ. 13: 1287-1297.
28
29      Forrest, J.;  Garber, R. W.; Newman,  L. (1981) Conversion rates in power plant plumes based on filter pack
30          data: the coal-fired Cumberland plume. Atmos.  Environ. 15: 2273-2282.
31
32      Freeman, D. L.; Chow, J. C.; Egami, R. T.; Watson, J. G. (1989) A receptor and dispersion modeling
33          software package. In: Watson, J. G., ed. Transactions: receptor models in air resources management.
34          Pittsburgh, PA: Air and Waste Management Association; pp. 243-268.
35
36      Frick, G. M.; Hoppel, W. A. (1993) Airship measurements of aerosol size distributions, cloud droplet spectra,
37          and trace gas concentrations in the marine boundary layer. Bull. Am. Meteorol.  Soc. 74: 2195-2202.
38
39      Fried, A.; Henry, B.  E.; Calvert, J. G.; Mozurkewich, M. (1994) The reaction probability of N2O5 with
40          sulfuric acid aerosols at stratospheric temperatures and compositions. J. Geophys. Res.  [Atmos.]
41          99: 3517-3532.
42
43      Friedlander, S. K. (1970) The characterization of aerosols distributed with respect to size and chemical
44          composition. J. Aerosol Sci.  1: 295-307.
45
46      Friedlander, S. K. (1977) Smoke, dust and haze: fundamentals of aerosol behavior. New York,  NY: John
47          Wiley & Sons,  Inc.
48
49      Friedlander, S. K.;  Turner, J. R.; Hering, S. V. (1986) J.  Aerosol Sci.  17: 240-244.
50
51      Fuchs, N. A. (1964) The mechanics of aerosols. New York, NY: Pergamon Press.
52
53      Fuchs, N. A.; Stutugin, A. G. (1971) Highly dispersed aerosols. In: Hidy, G. M.; Brock, J. R., eds. Topics in
54          current  aerosol  research. New York, NY: Pergamon; pp. 4-60.


         April 1995                                    3483      DRAFT-DO NOT QUOTE OR CITE

-------
  1       Fung, C. S.; Misra, P. K.; Bloxam, R.; Wong, S. (1991) A numerical experiment on the relative importance of
  2           H2O2 and O3 in aqueous conversion of SO2 to SO2"/4. Atmos. Environ. Part A 25: 411-423.
  3
  4       Fuzzi, S.; Castillo, R. A.; Jiusto, J. E.; Lala, G.  G. (1984) Chemical composition of radiation fog water at
  5           Albany, New York, and its relationship to fog microphysics. J. Geophys. Res. [Atmos.] 89: 7159-7164.
  6
  7       Fuzzi, S.; Orsi, G.; Nardini, G.; Fachini, M. C.; Mariotti, M.; McLaren, S.; McLaren, E.  (1988)
  8           Heterogeneous processes in the Po Valley (Italy) radiation fog. J. Geophys. Res. 93: 11141-11151.
  9
10       Gagosian, R. B.; Peltzer, E. T.; Merrill, J. T. (1987) Long range transport of terrestrially derived lipids in
11           aerosols from South Pacific. Nature (London) 325:  800-803.
12
13       Galloway, J. N.; Thornton, J. D.; Norton, S. A.; Volchok,  H. L.; McLean, R. A.  N. (1982) Trace metals in
14           atmospheric deposition: a review and assessment. Atmos. Environ. 16:  1677-1700.
15
16       Garland, J. A.  (1979) Resuspension of paniculate matter from grass and soil. Harwell, United Kingdom:
17           UK Atomic Energy Authority;  report no. AERE-R9452.
18
19       Garland, J.; Nicholson, K. (1991) A review of methods for sampling large airborne particles and associated
20           radioactivity. J. Aerosol Sci. 22: 479-499.
21
22       Gatz, D. F. (1977) Scavenging ratio measurements in METROMEX. In: Semonin, R. G.;  Beadle, R. W., eds.
23           Precipitation scavenging (1974): proceedings of a symposium; October  1974;  Champaign, IL. Oak Ridge,
24           TN: Energy Research and Development Administration; pp. 71-87. Available from: NTIS, Springfield,
25           VA; CONF-741003.  (ERDA symposium series no.  41).
26
27       Gebhart, K. A.; Lattimer, D. A.; Sisler, J. F.  (1990) Empirical orthogonal function analysis of the particulate
28           sulfate concentrations measured during WHITEX. In: Mathai, C. V., ed. Transactions: visibility and fine
29           particles, Pittsburgh,  PA: Air and Waste Management Association; pp.  860.
30
31       Gebhart, K. A.; Malm, W. C.;  Day, D. (1994) Examination of the effects of sulfate acidity  and relative
32           humidity on light scattering at Shenandoah National Park.  Atmos. Environ. 28:  841-849.
33
34       Gelbard, F.; Seinfeld, J. H. (1978) Numerical solution of the dynamic equation for  particulate systems.
35           J. Comp. Phys. 28: 357-375.
36
37       Gelbard, F.; Seinfeld, J. H. (1980) Simulation of  multicomponent aerosol dynamics. J. Colloid Interface Sci.
38           78:485-501.
39
40       Gerde, P.; Scholander, P. (1989) Mass transfer rates of polycyclic aromatic hydrocarbons between micron-size
41           particles and their environment—theoretical estimates. Environ. Health Perspect. 79: 249-258.
42
43       Germani, M. S.; Buseck, P. R.  (1991) Automated scanning electron microscopy for atmospheric particle
44           analysis. Anal. Chem. 63: 2232-2237.
45
46       Gervat, G. P.;  Clark, P. A.; Marsh, A. R. W. (1988) Controlled chemical  kinetics  experiment in cloud:
47           a review of the CERL/UMIST  Great Dunn Fell Project. In: Unsworth,  M. H.;  Fowler, D., eds. Acid
48           deposition  at high elevation sites. Hingham, MA: Kluwer Academic; pp. 283-298.
49
50       Gery, M. W.; Fox, D. L.; Jeffries, H. E.; Stockburger, L.; Weathers, W.  S. (1985) A continuous stirred tank
51           reactor investigation of the gas-phase reaction of hydroxyl radicals and toluene.  Int. J.  Chem.  Kinet.
52           17: 931-955.
53
         April 1995                                    3-184       DRAFT-DO NOT QUOTE OR CITE

-------
   1       Gery, M. W.; Fox, D. L.; Kamens, R. M.; Stockburger, L. (1987) Investigation of hydroxyl radical reactions
   2           with o-xylene and m-xylene in a continuous stirred tank reactor. Environ. Sci. Technol. 21: 339-348.
   3
   4       Ghan, S. J.; Chuang, C. C.; Penner, J. E. (1994) A parameterization of cloud droplet nucleation, part I: single
   5           aerosol type. Atmos. Res.: submitted.
   6
   7       Gifford, F. (1961) Use of routine meteorological observations for estimating atmospheric dispersion. Nucl. Saf.
   8           2: 47-57.
   9
 10       Gill, P. S.; Graedel, T.  E.; Wechsler, C. J. (1983) Organic films on atmospheric aerosol particles,  fog
 11           droplets, raindrops and snowflakes. Rev. Geophys. Space Phys. 21: 903.
 12
 13       Gillani, N. V. (1985) Modeling of chemical transformations of SOx and NOX in the polluted atmosphere: an
 14           overview of approaches and current status. In: De Wispelaere,  C, ed. Air pollution modeling and  its
 15           application IV. New York, NY: Plenum Publishers.
 16
 17       Gillani, N. V. (1986) Ozone formation in pollutant plumes:  a reactive plume model with arbitrary crosswind
 18           resolution.  Research  Triangle Park, NC: U.S. Environmental Protection Agency, Atmospheric Sciences
 19           Research Laboratory; EPA report no. EPA/600/3-86/051. Available from: NTIS, Springfield, VA;
 20           PB86-236973.
 21
 22       Gillani, N. V.;  Husar, R. B. (1976) Synoptic-scale haziness over the eastern US and its long range transport.
 23           Presented at: 4th National AMS/SAF conference on fire and forest meteorology Boston  MA
 24
 25       Gillani, N. V.;  Pleim, J.  E. (1994) Subgrid-scale  features of anthropogenic emissions of VOC and NOX in the
 26           context of regional Eulerian models.  Atmos. Environ.: submitted.
 27
 28       Gillani, N. V.;  Wilson, W. E. (1980) Formation and transport of ozone and aerosols in power plant plumes.
 29           Ann. N. Y.  Acad. Sci. 338: 276-296.
 30
 31       Gillani, N. V.;  Wilson, W. E. (1983) Gas-to-particle conversion of sulfur in power plant plumes—II.
 32           observations of liquid-phase conversions. Atmos. Environ. 17:  1739-1752.
 33
 34       Gillani, N. V.; Husar, R. B.; Husar, J. D.; Patterson, D. E.; Wilson, W. E., Jr. (1978) Project MISTT:
 35           kinetics of paniculate sulfur formation in a power plant  plume out to 300 km.  Atmos. Environ.
 36           12: 589-598.
 37
 38       Gillani, N.  V.; Kohli, S.; Wilson, W. E. (1981) Gas-to-particle conversion of sulfur in power plant  plumes—I.
 39          parametrization of the conversion rate for dry, moderately polluted ambient conditions.  Atmos  Environ
 40           15: 2293-2313.
 41
 42      Gillani, N.  V.; Colby, J.  A.; Wilson, W. E. (1983) Gas-to-particle  conversion of sulfur in power plant
 43          plumes—III. parameterization of plume-cloud interactions. Atmos. Environ. 17: 1753-1763
 44
 45      Gillani, N.  V.; Shannon, J. D.; Patterson, D. E. (1984) Transport processes. In: Altshuller, A.  P.; Linthurst,
 46          R. A.,  eds.  The acidic deposition phenomenon and its effects: critical assessment review papers, v.  1.
 47          Research Triangle Park, NC: U.S. Environmental Protection Agency; EPA report no.
 48          EPA/600/8-83/016AF. Available from: NTIS, Springfield, VA;  PB85-100030
49
 50      Gillani, N. V.; Schwartz, S. E.;  Daum, P. H.; Leaitch, W.  R.; Strapp, J. W.; Isaac, G. A. (1992) Fractional
51          activation of accumulation-mode aerosols in continental stratiform clouds. Presented at:  The World
52          Meterological Organization workshop on cloud microphysics and global change; August; Toronto, ON,
53          Canada.
54


         April  1995                                     3_185       DRAFT-DO  NOT QUOTE  OR CITE

-------
  1       Gillani, N. V.; Leaitch, W. R.; Strapp, J. W.; Isaac, G. A. (1994) Field observations in continental stratiform
  2           clouds: partitioning of cloud particles between droplets and unactivated interstitial aerosols. J. Geophys.
  3           Res.:  submitted.
  4
  5       Gillette, D. A. (1974) On the production of soil wind erosion aerosols having the potential for long range
  6           transport. J. Rech. Atmos. 8: 735-744.
  7
  8       Gillette, D. (1980) Major contributions of natural primary continental aerosols: source mechanisms. In: Kniep,
  9           T. J.; Lioy, P. J., eds. Aerosols: anthropogenic and natural, sources and transport;  January  1979; New
10           York, NY. Ann N. Y. Acad. Sci. 338: 348-358.
11
12       Gillette, D. A.;  Hanson, K. J. (1989) Spatial and temporal variability of dust production caused by wind erosion
13           in the United  States. J. Geophys. Res. 94:  2197-2206.
14
15       Gillette, D.; Nagamoto, C. (1992) [Size distribution and single particle composition for two dust storms in
16           Soviet Central Asia in September 1989 and size distributions and chemical composition of local soil].
17           In: Golitsyn, G., ed. Soviet-American experiment for the investigation of arid aerosols. St. Petersburg,
18           USSR: Typhoon; pp.  130-140.
19
20       Gillette, D. A.;  Passi, R.  (1988) Modeling dust emission caused by wind erosion.  J. Geophys. Res.
21           93: 14,233-14,242.
22
23       Gillette, D. A.;  Sinclair, P. C. (1990) Estimation of suspension of alkaline material by dust devils in the United
24           States. Atmos. Environ. Part  A 24: 1135-1142.
25
26       Gillette, D.; Walker,  T. (1977) Characteristics  of airborne particles produced by wind erosion of sandy soil,
27           high plains of west Texas. Soil Sci.  123: 97-110.
28
29       Gillette, D. A.;  Clayton, R. N.; Mayeda, T. K.; Jackson, M. L.; Sridhar, K. (1978) Tropospheric aerosols
30           from some major dust storms of the Southwestern United States. J.  Appl. Meteorol. 17: 832-845.
31
32       Gillette, D. A.;  Stensland, G. J.;  Williams, A. L.;  Barnard, W.; Gatz,  D.; Sinclair, P. C.; Johnson, T. C.
33           (1992) Emissions of alkaline elements calcium,  magnesium, potassium, and sodium from open sources in
34           the contiguous United States.  In: Global biogeochemical cycles.
35
36       Gleason, J. F.; Sinha, A.;  Howard, C. J. (1987) J.  Phys.  Chem. 91: 719-724.
37
38       Godish, T. (1985) Air quality. Chelsea, MI: Lewis  Publishers, Inc.
39
40       Godowitch, J. M.  (1989) Evaluation and sensitivity  analyseis results of the MESOPUFF II model with
41           CAPTEX measurements. Research Triangle Park, NC: U.S. Environmental Protection Agency,
42           Atmospheric Research and Exposure Assessment Laboratory;  EPA report no. EPA/600/3-89/056. Available
43           from:  NTIS, Springfield, VA; PB89-198253.
44
45       Goldberg,  E. D. (1985) Black carbon in the environment:  properties and distribution. New York, NY: John
46           Wiley & Sons. (Metcalf, R. L.; Stumm, W., eds. Environmental science and technology  series).
47
48       Gordon, G. E. (1980) Receptor models.  Environ. Sci. Technol. 14: 792-800.
49
50       Gordon, G. E. (1988) Receptor models.  Environ. Sci. Technol. 22: 1132-1142.
51
52       Gordon, G. E. (1991) Airborne particles on global and regional scales. Environ. Sci. Technol. 25: 1822-1828.
53
         April  1995                                     3-186       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Gordon, R. J.; Trivedi, N. J.; Singh, B. P.; Ellis, E. C. (1988) Characterization of aerosol organics by diffuse
  2          reflectance Fourier transform infrared spectroscopy. Environ.  Sci. Technol.  22: 672-677.
  3
  4      Goschnich, J.; Fichtner, M.; Lipp, M.; Schuricht, J.; Ache, M. (1990) Depth resolved chemical analysis of
  5          environmental microparticles by secondary mass spectrometry. Appl. Surface Sci. 70: 63-67.
  6
  7      Graedel, T. E.; Goldberg, K. I. (1983) Kinetic studies of raindrop chemistry:  1. inorganic and organic
  8          processes. J.  Geophys. Res.  88: 10865-10882.
  9
 10      Graedel, T. E.; Weschler, C. J. (1981) Chemistry within aqueous atmospheric aerosols and raindrops. Rev.
 11          Geophys.  Space Phys. 19: 505-539.
 12
 13      Grams, G. W.; Blifford, I. H., Jr.; Gillette, D. A.; Russell, P. B. (1974) Complex index of refraction of
 14          airborne soil particles. J.  Appl. Meteorol. 13: 459-471.
 15
 16      Gray, H. A. (1986) Control of atmospheric fine primary carbon particle concentration [Ph.D. thesis]. Pasadena,
 17          CA: California Institute of Technology; EQL report no. 23.
 18
 19      Gray, H. A.; Cass, G. R.; Huntzicker, J. J.;  Heyerdahl, E. K.; Rau, J. A. (1984) Elemental and organic
 20          carbon particle concentrations: a long-term perspective. Sci. Total Environ. 36: 17-25.
 21
 22      Gray, H. A.; Cass, G. R.; Huntzicker, J. J.;  Heyerdahl, E. K.; Rau, J. A. (1986) Characteristics of
 23          atmospheric organic and elemental carbon particle concentrations in Los Angeles.  Environ. Sci. Technol.
 24          20: 580-589.
 25
 26      Greenberg, R. R.; Zoller, W. H.; Gordon, G. E. (1978) Composition and size distributions of particles released
 27          in refuse  incineration.  Environ. Sci. Technol. 12: 566-573.
 28
 29      Grosjean, D.  (1984a) Particulate  carbon in Los Angeles' air. Sci. Total Environ. 32: 133.
 30
 31      Grosjean, D.  (1984b) Photooxidation of 1-heptane. Sci. Total Environ. 37:  195-211.
 32
 33      Grosjean, D.  (1985) Reactions of o-cresol and nitrocresol with NOX in sunlight and with ozone-nitrogen dioxide
 34          mixtures in the dark. Environ. Sci. Technol.  19: 968-974.
 35
 36      Grosjean, D.  (1992) In situ organic aerosol formation during a smog episode: estimated production and
 37          chemical  functionality. Atmos. Environ. Part  A 26:  953-963.
 38
 39      Grosjean, D.; Friedlander, S.  K.  (1975) Gas-particle distribution factors for organic and other pollutants in the
 40          Los Angeles atmosphere.  J. Air Pollut. Control Assoc. 25: 1038-1044.
 41
 42      Grosjean, D.; Seinfeld, J. H. (1989) Parameterization of the formation potential of secondary organic aerosols.
 43          Atmos. Environ. 23: 1733-1747.
44
45      Gundel, L. A.; Guyot-Sionnest, N.  S.; Novakov,  T.  (1989) A study of the  interaction of NC^ with carbon
46          particles.  Aerosol  Sci.  Technol.  10: 343-351.
47
48      Gundel, L. A.; Lee, V. C.; Mahanama, K. R. R.; Stevens, R. K.; Daisey, J. M. (1994) Direct determination
49          of the phase distributions of semi-volatile polycyclic aromatic hydrocarbons using annular denuders. Atmos.
50          Environ.: in press.
 51
52      Gunz, D. W.;  Hoffmann, M. R.  (1990) Atmospheric chemistry of peroxides: a review. Atmos. Environ. Part A
53          24: 1601-1633.
54


         April  1995                                     3-187       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Hahn, J. (1980) Organic constituents of natural aerosols. Ann. N. Y. Acad. Sci. 338: 359-376.
  2
  3       Hall, F. P., Jr.; Duchon, C. E.; Lee, L. G.; Hagen, R. R. (1976) Long-range transport of air pollution: a case
  4           study, August 1970. Mon. Weather Rev. 101: 404.
  5
  6       Haltiner, G. J. (1971) Numerical weather prediction. New York, NY: John Wiley & Sons, Inc.
  7
  8       Hamilton, R. S.; Mansfield, T. A. (1991) Airborne  paniculate elemental carbon: its sources,  transport and
  9           contribution to dark smoke and soiling. Atmos.  Environ. Part A 25: 715-723.
10
11       Ha'nel,  G. (1976) The properties of atmospheric aerosol particles as functions of the relative humidity at
12           thermodynamic equilibrium with the surrounding moist air. In: Landsberg, H. E.; Van Mieghem, J., eds.
13           Advances in geophysics: v.  19. New York, NY: Academic Press; pp.  73-188.
14
15       Ha'nel,  H. (1987) The role of aerosol properties during the condensational growth of cloud: a reinvestigation of
16           numerics and metaphysics. Contrib. Atmos. Phys. 60: 321-339.
17
18       Hansen, A. D. A.; Novakov, T. (1988) Real time measurements of the size fractionation of ambient black
19           carbon aerosols at elevated humidities. Aerosol  Sci.  Technol. 106-110.
20
21       Hansen, A. D. A.; Rosen, H. (1990) Individual measurements of the emission factor of aerosol black carbon in
22           automobile plumes. J. Air Waste Manage. Assoc. 40: 1654-1657.
23
24       Hansen, A. D. A.; Benner, W. H.; Novakov, T. (1991) S02 oxidation in laboratory clouds. Atmos. Environ.
25           25: 2521-2530.
26
27       Hansson, H.-C.; Svenningsson, B. (1994) Aerosols and clouds. In: Angeletti,  G.; Restelli, G., eds.
28           Physico-chemical behaviour of atmospheric pollutants, v, 2.  Proceedings of the sixth European symposium;
29           October 1993; Varese, Italy. Luxembourg, Office of Official Publications of the European Commission;
30           pp. 837-846.
31
32       Harms, D. E.; Raman, S.; Madala, R. V. (1992) An examination of four-dimensional data-assimilation
33           techniques for numerical weather prediction. Bull. Am.  Meteorol. Soc. 73: 425-440.
34
35       Harrison, L. (1985) The segregation of aerosols by cloud nucleating activity: II. observation of urban aerosol.
36           J. Clim. Appl. Meteorol. 24: 312-321.
37
38       Harrison, R. M.; Msibi, I. M. (1994) Validation of techniques for fast response measurement of HNO3 and
39           NH3 and determination of the [NH3][HNO3] concentration product. Atmos. Environ. 28:  247-255.
40
41       Harrison, R. M.; Pio, C. A. (1983) Size-differentiatedl composition of inorganic atmospheric aerosols of both
42           marine and polluted continental origin. Atmos. Environ.  17: 1733-1738.
43
44       Harrison, R. M.; Pio, C. A. (1983) A comparative  study of the ionic composition of rainwater and atmospheric
45           aerosols: implications for the mechanism of acidification of rainwater.  Atmos. Environ. 17: 2539-2543.
46
47       Harrison, R. M.; Laxen,  D. P. H.; Wilson,  S. J. (1981) Chemical associations of lead, cadmium, copper,  and
48           zinc in street dusts and roadside soils. Environ.  Sci. Technol. 15: 1378-1383.
49
50       Hart, K. M.; Pankow, J.  F.  (1994) High-volume air sampler for particle and gas sampling. 2. Use of backup
51           filters to correct for the adsorption of gas-phase polycyclic aromatic hydrocarbons to the front filter.
52           Environ. Sci. Technol. 28: 655-661.
53
         April 1995                                    3-188       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Hasan, H.; Dzubay, T. G. (1987) Size distributions of species in fine particles in Denver using a microorifice
  2          impactor. Aerosol Sci. Technol. 6: 29-39.
  3
  4      Hatakeyama, S.; Tanonaka, T.; Weng, J.; Bandow, H.; Takagi, H.; Akimoto, H. (1985) Ozone-cyclohexene
  5          reaction in air: quantitative analysis of paniculate products and the reaction mechanism. Environ. Sci.
  6          Technol. 19: 935-942.
  7
  8      Hatakeyama, S.; Ohno, M.; Weng, J.; Takagi, H.; Akimoto, H. (1987) Mechanism for the formation of
  9          gaseous and paniculate products from ozone-cycloalkene reactions in air. Environ. Sci.  Technol. 21: 52-57.
 10
 11      Hatakeyama, S.; Izumi, K.; Fukuyama, T.; Akimoto, H. (1989) Reactions of ozone with a-pinene and /3-pinene
 12          in air: yields of gaseous and paniculate products. J. Geophys. Res. [Atmos.] 94: 13,013-13,024.
 13
 14      Hatakeyama, S.; Izumi, K.; Fukuyama, T.; Akimoto, H.; Washida, N. (1991) Reactions of OH with a-pinene
 15          and 0-pinene in air: estimate of global CO production from the atmospheric oxidation of terpenes.
 16          J. Geophys. Res. [Atmos.] 96: 947-958.
 17
 18      Hegg, D. A. (1985) The importance of liquid-phase oxidation of SO2 in the troposphere. J.  Geophys. Res.
 19          90:  3773-3779.
 20
 21      Hegg, D. A. (1991) Particle production in clouds. Geophys. Res. Lett. 18: 995-998.
 22
 23      Hegg, D. A.; Hobbs, P. V. (1979) Some observations of paniculate nitrate concentrations in coal-fired power
 24          plant plumes.  Atmos. Environ. 13: 1715-1716.
 25
 26      Hegg, D. A.; Hobbs, P. V. (1980) Measurements of gas-to-particle conversion in the plumes from five
 27          coal-fired electric power plants. Atmos. Environ. 14: 99-116.
 28
 29      Hegg, D. A.; Hobbs, P. V. (1982) Measurements of sulfate production in natural clouds. Atmos. Environ.
 30          16:  2663-2668.
 31
 32      Hegg, D. A.; Hobbs, P. V. (1983a) Errata [Hegg and Hobbs (1982)]. Atmos. Environ. 17:  1059.
 33
 34      Hegg, D. A.; Hobbs, P. V. (1983b) Author's reply: measurements  of sulfate production in natural clouds.
 35          Atmos. Environ. 17: 2632-2633.
 36
 37      Hegg, D. A.; Hobbs, P. V. (1986) Sulfate and nitrate chemistry in cumuliform clouds. Atmos.  Environ.
 38          20:  901-909.
 39
40      Hegg, D. A.; Hobbs, P. V. (1987) Comparisons  of sulfate production due to ozone oxidation in clouds with a
41          kinetic rate equation. Geophys. Res. Lett. 14: 719-721.
42
43      Hegg, D. A.; Hobbs, P. V. (1988) Comparisons  of sulfate and nitrate production in clouds on the mid-Atlantic
44          and  Pacific Northwest coast of the United States. J. Atmos. Chem. 7: 325-333.
45
46      Hegg, D. A.; Larson, T. V. (1990) The effects of microphysical parameterization on model predictions of
47          sulfate predictions in clouds. Tellus Ser. B 42:  272-284.
48
49      Hegg, D. A.; Hobbs, P. V.; Radke, L. F. (1984) Measurements of the scavenging of sulfate and nitrate in
50          clouds. Atmos. Environ. 18:  1939-1946.
51
52      Hegg, D. A.; Radke, L. F.; Hobbs, P. V. (1990) Particle production associated with marine clouds.
53          J. Geophys. Res. [Atmos.] 95: 13917-13926.
54


         April 1995                                    3.139       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Hegg, D. A.; Radke, L.  F.; Hobbs, P. V. (1991) Measurements of Aitken nuclei and cloud condensation nuclei
  2           in the marine atmosphere and their relationship to the DMS-cloud-climate hypothesis. J. Geophys. Res.
  3           96:  18727-18733.
  4
  5       Hegg, D. A.; Yuen, P. F.; Larson, T. V. (1992) Modeling the effects of heterogeneous cloud chemistry on the
  6           marine particle size distribution. J. Geophys. Res. 97:  12927-12933.
  7
  8       Heintzenberg, J. (1989) Fine particles  in the global troposphere: a review. Tellus Ser. B 41: 149-160.
  9
10       Heintzenberg, J.; Winkler, P. (1984) Elemental carbon in the urban aerosol: results of a seventeen month study
11           in Hamburg, FRG. Sci.  Total Environ. 36: 27-38.
12
13       Heintzenberg, J.; Ogren,  J. A.; Noone, K. J.; Gardneus, L. (1989) The size distribution of submicrometer
14           particles within and about stratocumulus cloud droplets on Mt. Areskutan, Sweden. Atmos. Res.
15           24:89-101.
16
17       Henry, R. C. (1991) Multivariate receptor models.  In: Hopke, P. K., ed.  Receptor modeling for air quality
18           management. Amsterdam,  The Netherlands: Elsevier.
19
20       Henry, R. C.; Hidy, G. M.  (1979) Multivariate analysis of paniculate sulfate and other air quality variables by
21           principal components—part I: annual data from Los Angeles and New York. Atmos. Environ.
22           13:  1581-1596.
23
24       Henry, R. C.; Hidy, G. M.  (1982) Multivariate analysis of particulate sulfate and other air quality variables by
25           principal components—part II: Salt Lake  City, UT and St. Louis, MO. Atmos. Environ. 16:  929-943.
26
27       Henry, R. C.; Kim, B. M. (1989) A factor analysis receptor model  with explicit physical constraints.
28           In: Watson, J. G., ed. Transactions: receptor models in air resources  management. Pittsburgh, PA: Air  and
29           Waste Management Association; pp. 214-225.
30
31       Henry, R. C.; Wang, Y.-J.;  Gebhart, K. A. (1990) The relationship between empirical orthogonal functions and
32           sources of air pollution.  Atmos.  Environ. Part A  24: 503-509.
33
34       Henry, R. C.; Lewis, C.  W.; Collins,  J. F. (1994) Vehicle-related source compositions from ambient data: the
35           GRACE/SAFER method. Environ. Sci. Technol.  28: 823-832.
36
37       Hering, S. V.; Friedlander, S.  K. (1982) Origins of aerosol sulfur size distributions in the Los Angeles basin.
38           Atmos. Environ. 16:  2647-2656.
39
40       Hering, S. V.; McMurry, P. H. (1991) Optical counter response to  monodisperse atmospheric aerosols. Atmos.
41           Environ. Part A 25: 463-468.
42
43       Hicks, B. B. (1984) Dry deposition processes. In: Altshuller, A. P.; Linthurst, R. A., eds. Acidic deposition
44           phenomenon and its effects: critical assessment review papers, volume I, atmospheric sciences. Washington,
45           DC: U.S. Environmental Protection Agency, Office of Research and Development; pp. 7-1 to 7-70; EPA
46           report no. EPA/600/8-83/016AF. Available from: NTIS, Springfield,  VA; PB85-100030.
47
48       Hicks, B. B.; Meyers, T. P.  (1989) Atmosphere-surface exchange processes. Presented at: 82nd annual meeting
49           of the Air & Waste Management Association; June; Anaheim, CA. Pittsburgh, PA: Air & Waste
50           Management Association; paper no. 89-113.4.
51
52       Hicks, B. B.; Wesely, M. L.; Lindberg, S. E.; Bromberg,  S. M., eds.  (1986) Proceedings of the NAPAP
53           workshop on dry deposition; March; Harpers Ferry, WV. Oak Ridge, TN: National Oceanic and
54           Atmospheric Administration.


         April 1995                                   3.190       DRAFT-DO NOT QUOTE OR  CITE

-------
  1      Hicks, B. B.; Draxler, R. R.; Dodge, M.; Hales, J. M.; Albritton, D. L.; Schwartz. S. E.; Meyers, T. P.;
  2          Fehsenfeld, F. C.; Tanner, R. L.; Vong, R. J. (1991) Atmospheric processes research and process model
  3          development.  In: Irving,  P. M., ed. Acidic deposition: state of science and technology, volume I:
  4          emissions, atmospheric processes, and deposition. Washington, DC: The U.S. National Acid Precipitation
  5          Assessment Program. (State of science and technology report no. 2).
  6
  7      Hidy, G. M.; Appel, B. R.; Charlson, R. J.; Clark, W. E.; Friedlander, S.  K.; Hutchison, D. H.;  Smith,
  8          T. B.; Suder,  J.; Wesolowski, J. J.; Whitby, K. T. (1975) Summary of the California Aerosol
  9          Characterization Experiment.  J. Air Pollut. Control Assoc. 25: 1106-1114.
 10
 11      Hildemann, L. M. (1990) A study of the origin of atmospheric organic aerosols [Ph.D. thesis]. Pasadena, CA:
 12          California Institute of Technology.
 13
 14      Hildemann, L. M.; Russell, A.  G.; Cass, G. R. (1984) Ammonia and nitric acid concentrations in equilibrium
 15          with atmospheric aerosols: experiment vs theory, Atmos. Environ. 18:  1737-1750.
 16
 17      Hildemann, L. M.; Cass,  G.  R.; Mazurek, M. A.; Simoneit, B. R. T. (1993) Mathematical modeling of urban
 18          organic aerosol: properties measured by high-resolution gas chromatography. Environ. Sci. Technol.
 19          27:  2045-2055.
 20
 21      Hildemann, L. M.; Klinedinst, D. B.; Klouda, G.  A.; Currie, L. A.; Cass,  G. R. (1994) Sources of urban
 22          contemporary  carbon  aerosol. Environ. Sci. Technol. 28: 1565-1576.
 23
 24      Hildemann, L. M.; Mazurek, M.  A.; Cass, G. R.; Simoneit, B. R. T. (1994) Seasonal trends in Los Angeles
 25          ambient organic aerosol observed by high-resolution gas chromatography. Aerosol Sci. Technol.
 26          20:  303-317.
 27
 28      Hill, T.  A.; Choularton, T. W.; Penkett, S. A. (1986)  A model of sulphate production in a cap cloud and
 29          subsequent turbulent deposition onto the hill surface. Atmos.  Environ. 20: 1763-1771.
 30
 31      Hinds, W. C. (1982) Aerosol technology. New York, NY: John Wiley and Sons.
 32
 33      Hoffmann, M. R.; Boyce, S.  D. (1983) Catalytic autooxidation of aqueous sulfur dioxide  in relationship to
 34          atmospheric systems.  Adv. Environ. Sci. Technol.  12:  148-149.
 35
 36      Hoffmann, M. R.; Calvert, J. G. (1985) Chemical transformation modules for Eulerian acid deposition models.
 37          In: Lee,  Y.-N., ed. Chemical transformations in acid rain: volume II. investigation of kinetics and
 38          mechanism of aqueous-phase peroxide formation. Research Triangle Park, NC: U.S.  Environmental
 39          Protection Agency, Atmospheric Sciences Research Laboratory;  EPA report no. EPA/600/3-85/017.
40          Available from: NTIS, Springfield, VA;  PB85-173433.
41
42      Hoffmann, M. R.; Edwards, J. O. (1975) Kinetics of the oxidation of sulfite by hydrogen peroxide  in acidic
43          solution. J. Phys. Chem.  79: 2096-2098.
44
45      Hoffmann, M. R.; Jacob,  D. J.  (1984) Kinetics and mechanisms of the catalytic oxidation of dissolved sulfur
46          dioxide in aqueous solution: an application to nighttime fog water chemistry. In: Calvert, J. G., ed.  SO2,
47          NO  and NO2 oxidation mechanisms: atmospheric considerations. Boston, MA: Butterworth Publishers;
48          pp.  101-172. (Teasley, J. I., ed. Acid precipitation series: v.  3).
49
50      Hoffmann, H.  E.;  Kuehnemann, W. (1979) Comparison of the results of two measuring methods determining
51          the horizontal  standard visibility with the visual visibility range. Atmos.  Environ. 13:  1629-1634
52
53      Holsen,  T. M.; Noll, K. E. (1992) Dry deposition of atmospheric particles: application of current models to
54          ambient data.  Environ. Sci. Technol. 26: 1807-1815.


         April 1995                                    3_191       DRAFT-DO NOT  QUOTE OR CITE

-------
 1       Holsen, T. M.; Noll, K. E.; Fang, G.-C.; Lee, W.-J.; Lin, J.-M.; Keeler, G. J. (1993) Dry deposition and
 2           particle size distributions measured during the Lake Michigan Urban Air Toxics Study. Environ. Sci.
 3           Techno!. 27: 1327-1333.
 4
 5       Hong, M.-S.; Carmichael, G. R. (1986) An investigation of sulfate production in an orographic storm using a
 6           detailed transport/chemistry model coupled with a detailed cloud scavenging model. Atmos. Environ.
 7           20: 1989-1997.
 8
 9       Hopke, P. K. (1985) Receptor modeling in environmental chemistry.  New York, NY: John Wiley and Sons.
10
11       Hopke, P. K., ed. (1991) Receptor modeling for air quality management. Amsterdam,  The Netherlands:
12           Elsevier Publishers.
13
14       Hoppel, W. A. (1988) The role of non-precipitating cloud cycles and gas-to-particle conversion in the
15           maintainence of the submicron aerosol size distribution over the tropical oceans. In: Hobbs, P. V.;
16           McCormick, M. P., eds. Aerosols and climate.  Hampton, VA: A. Deepak Publishers; pp. 9-19.
17
18       Hoppel, W. A.; Frick, G. M. (1990) Submicron aerosol size distributions measured over the tropical and South
19           Pacific. Atmos. Environ. Part A 24:  645-659.
20
21       Hoppel, W. A.; Frick, G. M.; Larson, R. E. (1986) Effect of nonprecipitating clouds on the aerosol size
22           distribution in the marine boundary layer. Geophys.  Res. Lett. 13: 125-128.
23
24       Horvath,  H.; Noll, K. E. (1969) The relationship between atmospheric light scattering  coefficient and visibility.
25           Atmos. Environ. 3: 543-550.
26
27       Hosker, R. P., Jr.; Lindberg, S. E. (1982) Review:  atmospheric deposition and  plant assimilation of gases and
28           particles. Atmos. Environ. 16: 889-910.
29
30       Hov, O.;  Isaksen, I. S. A. (1981) Generation of secondary pollutants in a power plant  plume: a model study.
31           Atmos. Environ. 15:  2367-2376.
32
33       Huang, P.-F.; Turpin, B. J.; Pipho, M. J.; Kittelson, D. B.; McMurry,  P. H. (1994) Effects of water
34           condensation and evaporation  on diesel chain-agglomerate morphology. J. Aerosol  Sci. 25: 447-459.
35
36       Huie, R.  E.; Neta, P. (1987) Rate constants for some oxidations of S(IV) by radicals in aqueous solutions.
37           Atmos. Environ. 21:  1743-1747.
38
39       Huntzicker, J. J.; Hoffman, R. S.; Ling, C.-S. (1978) Continuous measurement and speciation of
40           sulfur-containing aerosols by flame photometry.  Atmos. Environ. 12: 83-88.
41
42       Huntzicker, J. J.; Johnson, R. L.; Shah, J. J.; Gary, R.  A. (1982) Analysis of organic and elemental carbon in
43           ambient  aerosols by a thermal-optical method. In: Wolff, G. T.; Klimisch, R. L.,  eds. Particulate
44           carbon—atmospheric  life cycle: proceedings of an international symposium;  October 1980; Warren, MI.
45           New  York, NY: Plenum Press; pp. 79-88.
46
47       Husain, L.; Dutkiewicz, V. A.; Hussain, M. M.; Khwaja, H. A.;  Burkhard, E. G.; Mehmood, G.; Parekh,
48           P. P.; Canelli, E. (1991) A study of heterogeneous oxidation of SO2 in summer clouds. J. Geophys. Res.
49           [Atmos.] 96: 18789-18805.
50
51       Husar, R. B.; Gillani, N. V.; Husar, J. D. (1976) Particulate sulfur formation in power plant: urban and
52           regional plumes. Presented at: Symposium on aerosol science and technology at the 82nd national meeting
53           of the AIChE; August-September; Atlantic City, MD.
54


         April 1995                                    3-192       DRAFT-DO  NOT  QUOTE OR CITE

-------
   1       Husar, R. B.; Patterson, D. E.; Husar, J. D.; Gillani, N. V.; Wilson, W. E., Jr. (1978) Sulfur budget of a
   2          power plant plume. In: Husar, R. B.; Lodge, J.  P., Jr.; Moore, D.  J., eds. Sulfur in the atmosphere:
   3          proceedings of the international symposium; September 1977; Dubrovnik, Yugoslavia. Atmos. Environ.
   4          12: 549-568.
   5
   6       Husar, R. B.; Holloway, J. M.; Patterson, D. E.; Wilson, W. E. (1981) Spatial and temporal pattern of eastern
   7          U.S. haziness: a summary. Atmos. Environ. 15: 1919-1928.
   8
   9       Huss, A., Jr.; Lim, P. K.; Eckert, C. A. (1982a) Oxidation of aqueous  sulfur dioxide. 1. Homogeneous
 10          manganese(II) and iron(II) catalysis at low pH. J. Phys. Chem.  86: 4224-4228.
 11
 12       Huss, A., Jr.; Lim, P. K.; Eckert, C. A. (1982b) Oxidation of aqueous  sulfur dioxide. 2. High-pressure studies
 13          and proposed reaction mechanisms. J. Phys. Chem.  86: 4229-4233.
 14
 15       Ibusuki, T.;  Takeuchi, K. (1987) Sulfur dioxide oxidation by oxygen catalyzed by mixtures of manganese(II)
 16          and iron(III) in aqueous solutions at environmental reaction conditions. Atmos. Environ. 21:  1555-1560.
 17
 18       Ip, W. M.; Gordon, R. J.; Ellis, E. C. (1984) Characterization of organics in aerosol samples from a
 19          Los angeles receptor site using extraction and liquid chromatography methodology. Sci. Total Environ.
 20          36: 203-208.
 21
 22       Irwin, J. S. (1983) Estimating plume dispersion: a comparison of several sigma schemes. J. Clim. Appl.
 23          Meteorol. 22: 92-114.
 24
 25       Irving, P. M., ed. (1991) Acidic deposition: state of  science and technology, volume 1: emissions, atmospheric
 26          processes, and deposition. Washington, DC: The U.S.  National Acid Precipitation Asssessment Program.
 27
 28       Izumi,  K.; Fukuyama,  T. (1990) Photochemical aerosol formation from aromatic hydrocarbons in the presence
 29          of NOX. Atmos. Environ. Part A 24: 1433-1441.
 30
 31       Izumi,  K.; Murano,  K.; Mizuochi, M.; Fukuyama, T. (1988) Aerosol formation by the photooxidation of
 32          cyclohexene in the presence of nitrogen oxides. Environ. Sci. Technol. 22: 1207-1215.
 33
 34      Jackson, M.  L.; Gillette, D. A.; Danielsen, E. F.; Blifford, I. H., Jr.; Bryson, R.; Syers, T.  K.  (1973) Global
 35          dustfall during the  quaternary as related to environments. Soil Sci. INQUA Issue 166: 135-145.
 36
 37      Jacob, D. J.  (1986) Chemistry of OH in remote clouds and its role in the production of formic acid and
 38          peroxymonosulfate. J. Geophys. Res.  [Atmos.] 91: 9807-9826.
 39
 40      Jacob, D. J.; Hoffmann, M. R. (1983) A dynamic model for the production of H+,  NO3 , and SO42" in urban
 41          fog. J. Geophys. Res. 88: 6611-6621.
 42
 43      Jacob, D. J.; Waldman, J. M.; Munger, J. W.;  Hoffmann, M. R. (1985) Chemical composition of fogwater
 44          collected along the California coast. Environ. Sci. Technol. 19:  730-736.
 45
 46      Jacob, D. J.; Waldman, J. M.; Munger, J. M.; Hoffmann,  M. R. (1986a) The H2SO4-HNO3-NH3 system at
 47          high humidities and in fogs, I: spatial and temporal patterns in the San Joaquin Valley of California.
48          J.  Geophys. Res. [Atmos.] 91: 1073-1088.
49
 50      Jacob, D. J.; Munger, J. W.;  Waldman, J. M.; Hoffman, M. R. (1986b) The H2SO4-HN03-NH3 system at
51          high humidities and in fogs: II. comparison of field data with the thermodynamic calculations. J. Geophys
52          Res. [Atmos.]  91: 1089-1096.
53
         April 1995                                   3.193       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Jaffrezo, J.-L.; Colin, J.-L. (1988) Rain-aerosol coupling in urban area: scavenging ratio measurement and
  2           identification of some transfer processes. Atmos. Environ. 22: 929-935.
  3
  4       Japar, S. M.; Brachaczek, W. W.; Gorse, R. A., Jr.; Norbeck, J. M.; Pierson, W. R. (1986) The contribution
  5           of elemental carbon to the optical properties of rural atmospheric aerosols. Atmos. Environ. 20: 1281-1289.
  6
  7       Javitz, H.  S.; Watson, J. G.; Guertin, J. P.; Mueller, P. K. (1988) Results of a receptor modeling feasibility
  8           study. JAPCA38: 661.
  9
10       Jensen, J.  B.; Charlson, R. J. (1984) On the efficiency of nucleation scavenging. Tellus Ser. B 36:  367-375.
11
12       Jessium, J. A.; Lu, H. P.; Chughtai, A. R.; Novakov, T. (1986) Appl. Spectrosc. 40: 113-119.
13
14       John,  W.;  Sethi, V.  (1993) Threshold for resuspension by particle impaction. Aerosol Sci. Technol.  19: 69-79.
15
16       John,  W.;  Wang, H.-C. (1991) Laboratory testing method for PM-10 samplers: lowered effectiveness from
17           particle loading. Aerosol Sci. Technol.  14: 93-101.
18
19       John,  W.;  Wall, S. M.; Ondo, J. L.; Winklmayr, W. (1990) Modes in the size distributions of atmospheric
20           inorganic aerosol. Atmos. Environ. Part A 24: 2349-2359.
21
22       John,  W.;  Winklmayr, W.; Wang, H.-C.  (1991) Particle deagglomeration and reentrainment in a PM-10
23           sampler. Aerosol Sci. Technol.  14: 165-176.
24
25       Johnson, C. A.; Sigg, L.; Zobrist, J. (1987) Case studies on the chemical composition of fogwater: the
26           influence of local gaseous emissions. Atmos. Environ. 21: 2365-2374.
27
28       Joos,  F.; Baltensperger, U. (1991) A field study on chemistry, S(IV) oxidation rates and vertical transport
29           during fog conditions. Atmos. Environ. Part A 25: 217-230.
30
31       Junge, C.  E. (1963) Air chemistry and radioactivity. New York, NY: Academic Press. (Van Mieghem, J.;
32           Hales, A. L.,  eds. International geophysics series:  v. 4).
33
34       Junge, C.  E. (1977) Basic considerations about trace constituents in the atmosphere as related to the fate of
35           global pollutants. In: Suffet, I. H., ed. Fate of pollutants in the air and water environments: part I,
36           mechanism of interaction between environments and mathematical modeling and the physical fate of
37           pollutants, papers from the 165th national American Chemical Society meeting; April 1975; Philadelphia,
38           PA. New York, NY: John Wiley & Sons; pp.  7-25. (Advances in environmental science and technology:
39           v. 8).
40
41       Junge, C.; McLaren, E. (1971) Relationship of cloud nuclei spectra to aerosol size distribution and composition.
42           J. Atmos. Sci. 28: 382-390.
43
44       Kadowaki, S. (1977) Size distribution and chemical composition of atmospheric paniculate nitrate in the Nagoya
45           area.  Atmos. Environ. 11: 671-675.
46
47       Kadowaki, S. (1990) Characterization of carbonaceous  aerosols  in the Nagoya urban area.  1. Elemental and
48           organic carbon concentrations and the origin of organic aerosols. Environ. Sci. Technol. 24: 741-744.
49
50       Kadowaki, S. (1994) Characterization of carbonaceous  aerosols  in the Nagoya urban area. 2. Behavior and
51           origin of paniculate «-alkanes. Environ. Sci. Technol. 28: 129-135.
52
53       Kahl,  J. D.;  Samson, P. J. (1986) Uncertainty in trajectory  calculations due to low-resolution meteorological
54           data. J. Clim.  Appl.  Meteorol. 25:  1816-1831.


         April 1995                                     3-194      DRAFT-DO  NOT QUOTE OR CITE

-------
  1      Kahl, J. D.; Samson, P. J. (1988) Trajectory sensitivity to rawinsonde data resolution. Atmos. Environ.
  2          22: 1291-1299.
  3
  4      Kahl, J. D.; Schnell, R. C.; Sheridan, P. J.; Zak, B.  D.; Church, H. W.; Mason, A. S.; Heffter, J. L.;
  5          Harris, J. M. (1991) Predicting atmospheric debris transport in real-time using a trajectory forecast model.
  6          Atmos. Environ. Part A 25: 1705-1713.
  7
  8      Kamens, R. M.; Jeffries, H. E.; Gery, M. W.; Wiener, R. W.; Sexton, K. G.; Howe,  G. B. (1981) The
  9          impact of *-pinene  on urban smog formation: an outdoor smog chamber study. Atmos. Environ.
 10          15: 969-981.
 11
 12      Kamens, R. M.; Gery,  M. W.; Jeffries, H. E.; Jackson, M.; Cole, E. I. (1982) Ozone-isoprene reactions:
 13          product formation and aerosol potential. Int. J. Chem. Kinet. 14: 955-975.
 14
 15      Kamens, R.; Odum, J.; Fan, Z.-H.  (1995) Some observations on times to equilibrium for semivolatile
 16          polycyclic aromatic hydrocarbons. Environ. Sci. Technol. 29: 43-50.
 17
 18      Kaplan, C. R.; Gentry,  J. W. (1988) Agglomeration of chain-like combustion aerosols due to Brownian motion.
 19          Aerosol Sci. Technol. 8: 11-28.
 20
 21      Kaplan, I. R.; Gordon,  R. J. (1994) Non-fossil-fuel fine-particle organic carbon aerosols in Southern California
 22          determined during the Los Angeles Aerosol Characterization and Source Apportionment Study. Aerosol Sci.
 23          Technol. 21: 343-359.
 24
 25      Karamchandani, P.; Venkatram, A.  (1992) The role of non-precipitating clouds in producing ambient sulfate
 26          during summer: results from simulations with the  Acid Deposition and Oxidant Model (ADOM) Atmos.
 27          Environ. Part A 26: 1041-1052.
 28
 29      Katrinak, K. A.; Rez, P.; Buseck, P. R. (1992) Structural variations in individual carbonaceous particles from
 30          an urban aerosol. Environ. Sci.  Technol. 26: 1967-1976.
 31
 32      Katrinak, K. A.; Rez, P.; Perkes, P. R.; Buseck, P. R. (1993) Fractal geometry of carbonaceous aggregates
 33          from an urban aerosol. Environ. Sci. Technol. 27: 539-547.
 34
 35      Kawamura, K.;  Ng, L.-L.; Kaplan,  I. R. (1985) Determination of organic acids (CrC10) in the atmosphere,
 36          motor exhausts, and engine oils. Environ. Sci. Technol.  19: 1082-1086.
 37
 38      Keeler, G. J.; Japar, S.  M.; Brachaczek, W. W.; Gorse, R.  A., Jr.; Norbeck, J. M.; Pierson, W. R. (1990)
 39          The sources of aerosol elemental carbon at Allegheny Mountain. Atmos. Environ. Part A 24: 2795-2805.
 40
 41       Kelly, N. A. (1987) The photochemical formation and fate of nitric acid in the metropolitan Detroit area:
42          ambient, captive-air irradiation and modeling results. Atmos. Environ. 21: 2163-2177.
43
44      Kelly, T. J.; Schwartz, S. E.; Daum, P. H. (1989) Detectability of acid producing reactions in natural clouds.
45           Atmos. Environ. 23: 569-583.
46
47       Kerminen, V. M.; Wexler, A. S. (1994) Particle formation due to SO2 oxidation and high relative humidity in
48           the remote  marine boundary layer. In press.
49
50       Kerminen, V.; Wexler, A. S. (1994) Post-fog nucleation of H2SO4-HO2 particles in smog. Atmos. Environ.
51           28: 2399-2406.
52
         April 1995                                    3_195       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Khlystov, A.; ten Brink, H. M.; Wyers, G. P. (1993) Hygroscopic growth rates of aerosols at high relative
  2           humidity. Petten, The Netherlands: Netherlands Energy Research Foundation ECN; report no.
  3           ECN-C-93-011.
  4
  5       Kim, Y. J. (1995) Response of the Active Scattering Aerosol Spectrometer Probe (ASASP-IOOX) to particles of
  6           different chemical composition. Aerosol Sci. Technol. 22: 33-42.
  7
  8       Kim, D.; Hopke, P. K.  (1988) Classification of individual particles based on computer-controlled scanning
  9           electron microscopy data. Aerosol Sci. Technol. 9: 133-151.
10
11       Kim, M. G.; Yagawa, K.;  Inoue, H.;  Lee, Y. K.; Shirai, T. (1990) Measurement of tire  tread in urban air by
12           pyrolysis-gas chromatography with flame photometric detection. Atmos. Environ. Part A 24: 1417-1422.
13
14       Kim, Y. P.; Seinfeld, J. H.; Saxena, P. (1993a) Atmospheric gas-aerosol, equilibrium: I.  thermodynamic
15           model. Aerosol Sci. Technol. 19:  157-181.
16
17       Kim, Y. P.; Seinfeld, J. H.; Saxena, P. (1993b) Atmospheric gas-aerosol equilibrium: II.  analysis of common
18           approximations and activity coefficient calculation methods.  Aerosol  Sci. Technol. 19: 182-198.
19
20       Kim, Y. J.; Boatman, J. F.; Gunter, R. L.; Wellman, D.  L.; Wilkison, S. W. (1993c) Vertical distribution of
21           atmospheric aerosol size distribution over south-central New Mexico. Atmos. Environ. Part A 27:
22           1351-1362.
23
24       Kirchner, W.; Welter, F.; Bongartz, A.; Kames, J.; Schweighoefer, S.; Schurath, U. (1990) Trace gas
25           exchange at the air/water interface: measurements of mass accomodation coefficients.  J. Atmos. Chem.
26           10: 427-449.
27
28       Kitsa, V.; Lioy, P. J. (1992) Near field dispersion of mechanically resuspended dust from an unpaved road.
29           Transactions AWMA 1: 199-210.
30
31       Kitsa, V.; Lioy, P. J.; Chow, J.  C.; Watson, J. G.;  Shupack, S.;  Howell, T.; Sanders, P. (1992) Particle-size
32           distribution of chromium: total and hexavalent chromium in inspirable, thoracic, and respirable soil particles
33           from contaminated sites in New Jersey.  Aerosol  Sci.  Technol. 17: 213-229.
34
35       Klee, A. J. (1984) Source control:  municipal solid waste incinerators. In:  Calvert, J. G.; Englund, eds.
36           Handbook of air pollution technology. New York, NY: John Wiley and Sons; pp. 513-550.
37
38       Kleinman, M. T.;  Pasternack, B. S.; Eisenbud, M.; Kneip, T. J. (1980)  Identifying and estimating the relative
39           importance of sources of airborne  particulates. Environ. Sci. Technol. 14: 62-65.
40
41       Klouda, G. A.; Currie,  L. A.; Verkouteren, R. M.;  Eifield, W.; Zak, B. D. (1988) Advances in
42           microradiocarbon dating and the direct tracing of environmental carbon. J. Radioanal. Nucl. Chem.
43           123: 191-197.
44
45       Knapp, K. T.; Bennett,  R.  L. (1990) Procedures for chemical characterization of sized particles in stationary
46           source emissions. Aerosol Sci. Technol.  12: 1067-1074.
47
48       Koenig, L. R.; Murray, F.  W. (1976)  Ice-bearing cumulus cloud evolution, numerical simulation and general
49           comparison against  observations. J. Appl. Meteorol.  15: 747-766.
50
51       Kohler, H. (1936) The nucleus in,  and the growth of, hygroscopic droplets. Trans. Faraday Soc.
52           32: 1152-1161.
53
         April 1995                                    3-196       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Kolb, C. E.; Jayne, J. T.; Worsnop, D. R.; Molina, M. J.; Meads, R. F.; Viggiano, A. A.  (1994) The gas
  2          phase reaction of sulfur trioxide with water vapor. J. Am. Chem.  Soc. 116: 10314-10321.
  3
  4       Kopcewicz, B.; Nagamoto, C.; Parungo, F.; Harris, J.; Miller, J.; Sievering, H.; Rosinski, J. (1991)
  5          Morphological studies of sulfate and nitrate particles on the east coast of North America and over the North
  6          Atlantic Ocean. Atmos. Res. 26: 245-271.
  7
  8       Korzun, E. A.; Heck, H. H. (1990) Sources and fates  of lead and cadmium in municipal solid waste. J. Air
  9          Waste  Manage. Assoc. 40: 1220-1226.
 10
 11       Koutrakis,  P.; Wolfson,  J. M.; Slater, J. L.; Brauer, M.; Spengler, J. D.;  Stevens, R. K.; Stone,  C. L. (1988)
 12          Evaluation of an annular denuder/filter pack system to collect acidic aerosols and gases. Environ. Sci.
 13          Technol. 22:  1463-1468.
 14
 15       Koutrakis,  P.; Wolfson,  J. M.; Spengler, J. D.; Stern, B.; Franklin, C. A. (1989) Equilibrium size of
 16          atmospheric aerosol  sulfates as  a function of the relative humidity. J. Geophys. Res. [Atmos.]
 17          94: 6442-6448.
 18
 19       Kowalczyk, G. S.; Gratt, L. B.; Ricci, P. F. (1987) An air emission risk assessment  for benzo(a)pyrene and
 20          arsenic from the Mt. Tom Power Plant. JAPCA 37: 361-369.
 21
 22       Krieger, M. S.; Kites, R. A. (1992) Diffusion denuder for the collection of semivolatile organic compounds.
 23          Environ. Sci.  Technol. 26: 1551-1555.
 24
 25       Krieger, M. S.; Kites, R. A. (1994) Measurement of polychlorinated biphenyls and polycyclic aromatic
 26          hydrocarbons  in air with a diffusion denuder. Environ.  Sci. Technol. 28: 1129-1133.
 27
 28       Kunen, S. M.; Lazrus, A.  L.; Kok, G. L.; Heikes, B.  G. (1983) Aqueous oxidation of SC^ by hydrogen
 29          peroxide. J. Geophys.  Res. C: Oceans Atmos. 88:  3671-3674.
 30
 31       Kuo, Y.-H.; Skumanich, M.; Haagenson, P. L.; Chang, J.  S. (1985) The accuracy of trajectory models as
 32          revealed by the observing system simulation experiments. Mon. Weather Rev. 113: 1852-1867.
 33
 34       Lamb, B.; Guenther, A.; Gay, D.; Westberg, H. (1987) A national inventory of biogenic hydrocarbon
 35          emissions. Atmos. Environ. 21: 1695-1705.
 36
 37       Lane, D. A.; Johnson, N. D.; Barton, S. C.; Thomas,  G. H. S.; Schroeder,  W. H. (1988) Development and
 38          evaluation of a novel gas and particle sampler for semivolatile chlorinated organic compounds in ambient
 39          air.  Environ. Sci. Technol. 22:  941-947.
 40
 41       Langner, J.; Rodhe, H. (1991) A  global three-dimensional model of the tropospheric sulfur cycle. J.  Atmos.
 42          Sci. 13: 225-263.
 43
 44      Larson, S. M.; Cass, G.  R. (1989) Characteristics of summer midday low-visibility events in the Los Angeles
 45           area. Environ. Sci. Technol. 23: 281-289.
46
47       Larson, S. M.; Cass, G.  R.; Gray, H.  A. (1989) Atmospheric carbon particles and the Los Angeles visibility
48           problem.  Aerosol Sci. Technol.  10:  118-130.
49
50       Lazrus, A. L.; Haagenson, P. L.; Kok, G. L.; Huebert, B.  J.; Kreitzberg, C. W.; Likens,  G. E.; Mohnen,
51           V. A.;  Wilson, W. E.; Winchester, J.  W. (1983) Acidity in air and water in a case of warm frontal
52           precipitation. Atmos. Environ. 17: 581-591.
53
         AP»1 1995                                    3.!97       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Leaitch, W. R.; Strapp, J. W.; Wiebe, H. A.; Isaac, G. A. (1983) Measurements of scavenging and
  2           transformation of aerosol inside cumulus. In: Pruppacher, H. R.; Seronin, R. G.; Slinn, W. G. N., eds.
  3           Precipitation scavenging, dry deposition and resuspension, v. 1. New York, NY: Elsevier; pp. 53-69.
  4
  5       Leaitch, W. R.; Bottenheim, J. W.; Strapp, J. W. (1988) Possible contribution of N2O5 scavenging to HNO3
  6           observed in winter stratiform cloud. J.  Geophys. Res. 93: 12569-12584.
  7
  8       Leaitch, W. R.; Strapp, J. W.; Wiebe, H. A.; Anlauf, K. G.; Isaac, G. A. (1986a) Chemical and
  9           microphysical studies of non-precipitating summer cloud. J.  Geophys.  Res. [Atmos.] 91: 11821-11831.
10
11       Leaitch, W. R.; Strapp, J. W.; Hudson, J.  G.; Isaac, G. A. (1986b) Cloud droplet nucleation and cloud
12           scavenging of aerosol sulfate in polluted atmospheres. Tellus Ser. B 38: 328-344.
13
14       Leaitch, W. R.; Isaac, G. A.; Strapp, J. W.; Banic, C. M.; Wiebe, H. A.  (1992) The relationship between
15           cloud droplet number concentrations and anthropogenic pollution: observations and climatic implications.
16           J. Geophys. Res. [Atmos.] 97: 2463-2474.
17
18       Lee, Y. N.; Schwartz, S. E. (1983) Kinetics of oxidation of aqueous sulfur(IV) by nitrogen dioxide.
19           In: Pruppacher, H. R.; Semonin, R. G.; Slinn,  W. G. N., eds. Precipitation scavenging, dry deposition and
20           resuspension: v. 1. New York, NY: Elsevier.
21
22       Lee, Y.-N.; Shen, J.; Klotz, P. J.; Schwartz, S. E.; Newman, L. (1986) Kinetics of the hydrogen
23           peroxide-sulfur(IV) reaction in rainwater collected at a northeastern U.S.  site. J. Geophys. Res. [Atmos.]
24           91: 13264-13274.
25
26       Lelieveld, J.; Crutzen, P. J. (1990) Influences of cloud photochemical processes on tropospheric ozone.  Nature
27           (London) 343: 227-233.
28
29       Lelieveld, J.; Crutzen, P. J. (1991) The role of clouds in tropospheric photochemistry. J. Atmos. Chem.
30           12: 229-267.
31
32       Lelieveld, J.; Heintzenberg, J. (1992) Sulfate cooling effect on climate through in-cloud oxidation of
33           anthropogenic SO2. Science (Washington, DC) 258: 117-120.
34
35       Leone, J. A.; Flagan, R. C.;  Grosjean, D.;  Seinfeld, J. H. (1985) An outdoor smog chamber and modeling
36           study of toluene-NOx photooxidation. Int. J. Chem. Kinet. 17: 177-216.
37
38       Lesaffre, F. (1989) Characterization of aerosol aggregates through fractal parameters: effects due to humidity.
39           J. Aerosol.  Sci. 20: 857-860.
40
41       Lewis, R. S.; Deen, W. M. (1994) Kinetics of the reaction of nitric oxide with oxygen in aqueous solutions.
42           Chem. Res. Toxicol. 7: 568-574.
43
44       Lewis, C. W.;  Stevens, R.  K. (1987) Hybrid receptor models. In: Johnson, R. W.; Gordon, G.,  eds. The
45           chemistry of acid rain:  sources and atmospheric processes. Washington, DC: American Chemical Society;
46           pp. 58-65.  (ACS symposium series 349).
47
48       Lewis, C. W.;  Baumgardner, R. E.; Stevens, R, K.; Claxton, L. D.; Lewtas, J.  (1988) Contribution of
49           woodsmoke and motor  vehicle emissions to ambient aerosol  mutagenicity. Environ. Sci. Technol.
50           22: 968-971.
51
52       Li, W.; Montassier, N.; Hopke, P. K. (1992) A system to measure the hygroscopicity of aerosol particles.
53           Aerosol Sci. Technol. 17: 25-35.
54


         April 1995                                    3-198       DRAFT-DO NOT  QUOTE OR CITE

-------
  1      Li, S.; Anlauf, K. G.; Wiebe, H. A. (1993) Heterogeneous nighttime production and deposition of paniculate
  2          nitrate at a rural site in North America during summer 1988. J.  Geophys. Res. 98: 5139-5157.
  3
  4      Liebsch,  E. J.; de Pena, R. G. (1982) Sulfate aerosol production in coal-fired power plant plumes. Atmos.
  5          Environ. 16: 1323-1331.
  6
  7      Ligocki,  M. P.; Pankow, J. F. (1989) Measurements of the gas/particle distributions of atmospheric organic
  8          compounds. Environ. Sci. Technol. 23: 75-83.
  9
 10      Ligocki,  M. P.; Salmon, L. G.; Fall, T.; Jones, M. C.; Nazaroff, W. W.; Cass, G. R. (1993) Characteristics
 11          of airborne particles inside southern California museums. Atmos. Environ. Part A 27: 697-711.
 12
 13      Lin, X.;  Chameides, W. L. (1991) Model studies of the impact of chemical inhomogeneity on SC«2 oxidation in
 14          warm stratiform clouds. J. Atmos.  Chem. 13: 109-129.
 15
 16      Lin, J.-M.; Fang, G.-C.; Holsen, T. M.; Noll, K. E. (1993) A comparison of dry  deposition modeled from
 17          size distribution data and measured with a smooth surface for total particle mass, lead and calcium in
 18          Chicago. Atmos. Environ. Part A 27: 1131-1138.
 19
 20      Lin, J. J.; Noll, K. E.; Holsen, T. M.  (1994) Dry deposition velocities as a function of particle size in the
 21          ambient atmosphere. Aerosol Sci. Technol. 20: 239-252.
 22
 23      Lind, J. A.; Kok, G. L. (1986) Henry's law determinations for aqueous solutions of hydrogen peroxide,
 24          methylhydroperoxide, and peroxyacetic acid.  J. Geophys. Res. [Atmos.] 91: 7889-7895.
 25
 26      Lind, J. A.; Kok, G. L. (1994) Correction to "Henry's law determinations for aqueous solutions of hydrogen
 27          peroxide, methylhydroperoxide, and peroxyacetic acid" by John  A. Lind and Gregory L. Kok. J.  Geophys.
 28          Res.  [Atmos.] 99: 21119.
 29
 30      Lind, J. A.; Lazrus, A. L.  (1983) Aqueous-phase oxidation of sulfur(IV) by some organic peroxides.  EOS
 31          Trans. 64: 670.
 32
 33      Lindberg, S. E. (1982) Factors influencing trace metal, sulfate and hydrogen ion concentrations in rain. Atmos.
 34          Environ.  16: 1701-1709.
 35
 36      Linton, R. W.; Loh, A.; Natusch, D. F. S.;  Evans, C. A.; Williams, P. (1976)  Surface predominance of trace
 37          elements in airborne particles. Science (Washington, DC) 191: 852.
 38
 39      Lioy, P. J.; Daisey, J. M. (1986) Airborne toxic elements and organic substances. Environ. Sci. Technol.
 40          20: 8-14.
 41
 42      Lioy, P. J.; Watson, J. G.; Spengler, J. D. (1980) APCA specialty conference workshop  on baseline data for
 43          inhalable paniculate matter. J. Air Pollut. Control Assoc. 30: 1126-1130.
 44
45      Lioy, P. J.; Mallon, R. P.; Lippmann, M.; Kneip, T.  J.; Samson, P. J. (1982) Factors affecting the variability
46          of summertime sulfate in a rural area using principal component  analysis. J.  Air Pollut. Control Assoc.
47          32: 1043-1047.
48
49      Lipfert, F. W. (1994) Filter artifacts associated with paniculate measurements: recent evidence and effects on
50          statistical relationships. Atmos. Environ. 28: 3233-3249.
51
52      Lippmann, M. (1980) Size distributions in urban aerosols.  In: Kneip, T. J.; Lioy, P. J., eds.  Aerosols:
53           anthropogenic and  natural, sources and transport. Ann. N.  Y. Acad. Sci. 338: 1-12
54


         April 1995                                    3499       DRAFT-DO NOT  QUOTE OR CITE

-------
 1       Liu, P. S. K.; Leaitch, W. R.; Strapp, J. W.; Wasey, M. A. (1992) Response of particle measuring systems
 2          airborne to ASASP and PCASP to NaCl and latex particles. Aerosol Sci. Technol. 16: 83-95.
 3
 4       Liu, P. S. K.; Leaitch, W. R.; McDonald,  A. M.; Isaac, G. A.; Strapp, J. W.;  Wiebe, H. A. (1994) Sulfate
 5          production in a summer cloud over Ontario, Canada. Tellus 45: 368-389.
 6
 7       Lowenthal, D. H.; Rogers, C.  F.; Saxena,  P.; Watson, J. G.; Chow, J. C. (1994) Sensitivity of estimated light
 8          extinction coefficients to model assumptions and measurement errors. Atmos. Environ: accepted.
 9
10       Lundgren, D.  A.; Hausknecht, B. J. (1982) Ambient aerosol mass distribution of 1 to 100 /*m particles in five
11          cities. Presented at: 75th annual meeting, Air Pollution Control Association; June; New Orleans, LA.
12          Pittsburgh, PA: Air Pollution Control Association; paper no. 82-45.4.
13
14       Lundgren, D.  A.; Paulus, J.  J. (1975) The  mass distribution of large atmospheric particles. J. Air Pollut.
15          Control Assoc.  25: 1227-1231.
16
17       Lundgren, D.; Burton, R. M. (1995) The effect of particle size distribution on the cut point between fine and
18          coarse ambient  mass  fractions. Inhalation Toxicol.: accepted.
19
20       Lundgren, D.; Haasknecht, B.; Burton, R.  (1984) Large particle size distribution in five U.S. cities and the
21          effect on the new ambient paniculate matter standard (PM10). Aerosol Sci. Technol. 7: 467-473.
22
23       Lurmann, F. W.; Collins, J.; Coyner,  L. (1988) Development of a chemical transformation submodel for
24          annual PM10 dispersion modeling. El Monte,  CA: South Coast Air Quality Management District; final
25          report no. P-6190-003.
26
27       Lusis, M. A.; Anlauf, K. G.; Barrie, L.  A.; Wiebe, H. A. (1978) Plume chemistry studies at a northern
28          Alberta power plant. Atmos. Environ.  12: 2429-2437.
29
30       Lyons, W. A.; Tremback, C. J.; Tesche, T. W.  (1991) Lake Michigan ozone study prognostic modeling: model
31          performance evaluation and sensitivity  testing. Crested Butte/ Fort Collins, CO: Alpine Geophysics,
32          ASTeR, Inc.
33
34       Maahs, H. G. (1983) Kinetics and mechanism of the oxidation of S(IV) by ozone in aqueous solution with
35          particular reference to SO2 conversion  in nonurban tropospheric clouds. J. Geophys. Res. 88: 10721-10723.
36
37       Macias, E. S.; Hopke, P. K., eds. (1981) Atmospheric aerosol: source/air quality relationships. Washington,
38          DC: American  Chemical Society. (ACS symposium series 167).
39
40       Macias, E. S.; Zwicker, J. O.; White, W.  H. (1981) Regional haze case studies in the southwestern U.S.—II.
41          source contributions. In: White, W. H.; Moore, D. J.; Lodge, J. P., Jr., eds. Plumes and visibility:
42          measurements and model components:  proceedings of the symposium; November 1980; Grand Canyon
43          National Park,  AZ. Atmos. Environ. 15: 1987-1997.
44
45       Makhon'ko, K.  P. (1986) Vetrovoy zakhvat pweel c podstilaioushey poberkhnosty, pokreetoy travot. Meteorol.
46          Gidrol. 10: 61-65.
47
48       Malm, W. C.; Gebhart, K. A.; Molenar, J.; Cahill, T.; Eldred, R.; Huffman,  D. (1994) Examining the
49          relationship between atmospheric aerosols and light extinction at Mount Rainier and North Cascades
50          National Parks. Atmos.  Environ. 28: 347-360.
51
52       Mamane, Y.;  Noll, K. E. (1985) Characterization of large particles at a rural site in the eastern United States:
53          mass distribution and individual particle analysis.  Atmos.  Environ. 19: 611-622.
54


         April 1995                                     3-200        DRAFT-DO NOT  QUOTE OR CITE

-------
  1      Mamane, Y.; Miller, J. L.; Dzubay, T. G. (1986) Characterization of individual fly ash particles emitted from
  2          coal- and oil-fired power plants. Atmos. Environ. 20: 2125-2135.
  3
  4      Mamane, Y.; Stevens, R. K.; Dzubay, T. G. (1990) On the sources of fine and coarse carbon particles in the
  5          Great Lakes. J. Aerosol Sci.  21: S353-S356.
  6
  7      Mamane, Y.; Dzubay, T. G.; Ward, R. (1992) Sulfur enrichment of atmospheric minerals and spokes. Atmos.
  8          Environ. Part A 26: 1113-1120.
  9
 10      Mansoori, B. A.; Johnston, M. V.; Wexler, A. S. (1994) Quantitation of ionic species in single microdroplets
 11          by on-line laser desorption/ionization. Anal. Chem. 66: 3681-3687.
 12
 13      Marlia, J. C.; Kim, B. M.; Wu, R. (1990) Annual PM10 dispersion model application to the South Coast air
 14          basin. El Monte, CA:  South Coast Air Quality Management District; technical report V-D.
 15
 16      Marple, V. A.; Rubow, K. L.; Behm,  S. M. (1991) A microorifice uniform deposit impactor (MOUDI):
 17          description, calibration, and use. Aerosol Sci. Technol. 14: 434-446.
 18
 19      Martin, L. R. (1984) Kinetic studies of sulfite oxidation in aqueous solution. In: Calvert, J. G., ed. SO2,
 20          NO and NO2 oxidation mechanisms:  atmospheric considerations. Boston, MA: Butterworth Publishers; pp.
 21          63-100.  (Teasley, J.  I., ed. Acid precipitation series: v. 3).
 22
 23      Martin, L. R.; Damschen,  D.  E.  (1981) Aqueous oxidation of sulfur dioxide by hydrogen peroxide at low pH.
 24          Atmos. Environ. 15: 1615-1621.
 25
 26      Martin, L. R.; Hill, M. W. (1987a) The iron catalyzed oxidation of sulfur: reconciliation of the literature rates.
 27          Atmos. Environ. 21: 1487-1490.
 28
 29      Martin, L. R.; Hill, M. W. (1987b) The effect of ionic strength on the manganese catalyzed oxidation of
 30          sulfur(IV). Atmos. Environ. 21: 2267-2270.
 31
 32      Martin, L. R.; Hill, M. W.; Tai, A. F.; Good,  T. W. (1991) The iron catalyzed oxidation of sulfur(IV) in
 33          aqueous  solution: differing effects of organics at high and low pH. J. Geophys. Res. 96: 3085-3097.
 34
 35      Martonen, T. B.; Zhang, Z. (1993) Deposition of sulfate acid aerosols in the developing human lung. Inhalation
 36          Toxicol. 5:  165-187.
 37
 38      Martonen, T. B.; Barnett, A. E.; Miller, F. J. (1985) Ambient sulfate aerosol  deposition in man: modeling the
 39          influence of hygroscopicity. Environ. Health Perspect.  63: 11-24.
 40
 41      Mason, A. S.; Gifford, F.  A. (1992) Atmospheric tracer  dispersion over a 24-h time span. Atmos. Environ.
 42          Part A 26: 3203-3205.
 43
 44      Matsoukas, T.; Friedlander, S. K. (1991) Dynamics of aerosol agglomerate formation. J. Colloid Interface Sci.
 45          146: 495-506.
46
 47      Mazurek, M.  A.; Simoneit, B. R. T. (1984) Characterization of biogenic and petroleum-derived organic matter
48          in aerosols over remote, rural and urban areas.  In: Identification and analysis of organic pollutants in air:
49          an ACS symposium.  Woburn, MA: Ann Arbor Science/Butterworth  Publishers; pp. 353-370.
 50
51       Mazurek, M. A.; Cass, G.  R.; Simoneit, B. R.  T. (1989) Interpretation of high-resolution gas chromatography
52          / mass spectrometry data aquired from atmospheric organic aerosol samples. Aerosol Sci.  Technol. 10:
53          408-420.
54


         April 1995                                    3-201       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Mazurek, M. A.; Cass, G. R.; Simoneit, B. R. T. (1991) Biological input to visibility-reducing aerosol particles
  2          in the remote arid southwestern United States. Environ. Sci. Technol.  25: 684-694.
  3
  4       McArdle, J. V.; Hoffmann, M. R. (1983) Kinetics and mechanism of the oxidation of aquated sulfur dioxide by
  5          hydrogen peroxide at low pH. J. Phys. Chem. 87: 5425-5429.
  6
  7       McHenry, J. N.; Dennis, R.  L. (1991) Partitioning of the sulfate budget into gas- and aqueous-phase
  8          components in the regional acid deposition model (RADM). Presented at: Proceedings of the 7th joint
  9          AMS-AWMA conference: applications of air pollution meteorology; New Orleans, LA. pp. 143-147.
10
11       McHenry, J. N.; Dennis, R.  L. (1994) The relative importance of oxidation pathways and clouds to atmospheric
12          ambient sulfate production as predicted by the regional and deposition model. J. Appl. Meteorol.
13          33: 890-905.
14
15       McHenry, J. N.; Binkowski, F. S.;  Dennis, R. L.; Chang, J. S.; Hopkins, D. (1992) The tagged species
16          engineering model (TSEM). Atmos.  Environ. Part A 26: 1427-1443.
17
18       McMahon, T.  A.; Denison, P. J. (1979) Empirical atmospheric deposition parameters—a survey.  Atmos.
19          Environ. 13: 571-585.
20
21       McMurry, P. H.; Friedlander, S. K. (1979) New particle formation in the presence of an aerosol. Atmos.
22          Environ. 13:  1635-1651.
23
24       McMurry, P. H.; Grosjean, D. (1985) Photochemical formation of organic aerosols: growth laws and
25          mechanisms. Atmos. Environ. 19: 1445-1451.
26
27       McMurry, P. H.; Stolzenburg, M. R. (1989) On the sensitivity of particle size to relative humidity for Los
28          Angeles aerosols.  Atmos. Environ. 23: 497-507.
29
30       McMurry, P. H.; Wilson, J.  C. (1982) Growth laws for the formation of secondary ambient aerosols:
31          implications for chemical conversion mechanisms. Atmos. Environ.  16: 121-134.
32
33       McMurry, P. H.; Wilson, J.  C. (1983) Droplet phase (heterogeneous) and gas phase (homogeneous)
34          contributions to secondary ambient aerosol formation as functions of relative humidity. J. Geophys. Res.
35          88: 5101-5108.
36
37       McMurry, P. H.; Zhang, X.  Q.  (1989) Size distributions of ambient organic and elemental carbon. Aerosol Sci.
38          Technol. 10: 430-437.
39
40       McMurry, P. H.; Rader, D. J.; Stith, J.  L. (1981) Studies of aerosol formation in power plant plumes—I.
41          growth laws for secondary aerosols in power plant plumes: implications for chemical conversion
42          mechanisms. In: White, W. H.; Moore, D. J.; Lodge, J. P., Jr., eds.  Plumes and visibility: measurements
43          and model components: proceedings  of the symposium; November 1980; Grand Canyon National Park, AZ.
44          Atmos. Environ. 15: 2315-2327.
45
46       Meagher, J. F.; Stockburger, L.; Bailey, E. M.; Huff,  O.  (1978) The oxidation of sulfur dioxide to sulfate
47          aerosols in the plume of a coal-fired power plant. Atmos. Environ.  12: 2197-2203.
48
49       Meagher, J. F.; Stockburger, L., Ill; Bonanno, R. J.; Bailey, E. M.; Luria, M.  (1981) Atmospheric oxidation
50          of flue gases from coal-fired power plants—a comparison between conventional and scrubbed plumes.
51          Atmos. Environ. 15: 749-762.
52
53       Meagher, J. F.; Olszyna, K.  J.; Weatherford, F. P.; Mohnen, V. A. (1990) The availability of H2O2 and O3
54          for aqueous phase oxidation of SO2.  The question of linearity. Atmos. Environ. Part A 24: 1825-1829.


         April 1995                                   3-202       DRAFT-DO NOT QUOTE  OR CITE

-------
  1       Meng, Z.;  Seinfeld, J. H. (1994) On the source of the submicrometer droplet mode of urban and regional
  2          aerosols. Aerosol Sci. Technol. 20: 253-265.
  3
  4       Meng, Z.;  Seinfeld, J. H.; Saxena,  P.; Kim, Y. P. (1995) Contribution of water to paniculate mass in the
  5          South Coast Air Basin.  Aerosol Sci. Technol. 22:  111-123.
  6
  7       Michaud, D.; Baril, M.; Perrault, G. (1993) Characterization of airborne dust from cast iron foundries by
  8          physico-chemical methods and multivariate statistical analyses. Air Waste 43: 729-735.
  9
 10       Middleton, P.; Brock, J. (1976) Simulation of aerosol kinetics. J. Colloid Interface Sci. 54:  249-264.
 11
 12       Midwest Research Institute. (1976) Section 11.2, fugitive dust sources: an AP-42 update of open source fugitive
 13          dust emissions.  Research Triangle Park, NC: U.S. Environmental Protection Agency; report no.
 14          EPA-450/2-76-029.
 15
 16       Midwest Research Institute. (1983) Section 11.2, fugitive dust sources: an AP-42 update of open source fugitive
 17          dust emissions.  Research Triangle Park, NC: U.S. Environmental Protection Agency; report no.
 18          EPA-450/4-836-010.
 19
 20       Milford, J.  B.; Davidson, C. I. (1985) The sizes of paniculate trace elements in the atmosphere—a review.
 21          J. Air Pollut. Control Assoc. 35: 1249-1260.
 22
 23       Milford, J.  B.; Davidson, C. I. (1987) The sizes of paniculate sulfate and nitrate in the atmosphere—a review.
 24          JAPCA 37: 125-134.
 25
 26       Miller, F. J.; Gardner, D. E.; Graham, J.  A.; Lee, R. E., Jr.; Wilson, W. E.; Bachmann, J. D.  (1979) Size
 27          considerations for establishing a standard for inhalable particles.  J. Air Pollut. Control Assoc. 29:  610-615.
 28
 29       Mohnen, V. A.;  Kedlacek, J. A. (1989) Cloud chemistry research at Whiteface Mountain. Tellus Ser.
 30          B 41: 79-91.
 31
 32       Moran,  M.  D. (1992) Numerical modeling of mesoscale atmospheric dispersion [Ph.D. dissertation].
 33          Ft. Collins, CO: Colorado State University,  Department of Atmospheric Science.
 34
 35       Morandi, M. (1985) Development of source apportionment models for inhalable particluate matter and its
 36          extractable organic fractions in urban areas of New Jersey [Ph.D. dissertation]. Syracuse, NY: New York
 37          University.
 38
 39      Morandi, M. T.; Lioy, P. J.; Daisey, J. M. (1991) Comparison of two multivariate modeling approaches for
 40          the source apportionment of inhalable particulate matter in Newark,  NJ. Atmos. Environ. Part A
 41           25:  927-937.
 42
 43       Mountain, R.  D.; Mulholland, G. W.; Baum, H. (1986) Simulation of aerosol agglomeration in the free
 44          molecular and continuum flow regimes. J. Colloid Interface Sci.  114: 67-81.
45
46       Moyers, J. L.; Ranweiler, L. E.; Hopf, S.  B.; Korte, N. E.  (1977) Evaluation of particulate trace species in
47           southwest desert atmosphere. Environ.  Sci. Technol.  11: 789-795.
48
49       Mozurkewich, M.; Calvert, J. G. (1988) Reaction probability of N2O5 on aqueous aerosols. J. Geophys. Res.
50           [Atmos.] 93: 15889-15896.
51
52       Mozurkewich, M.; McMurry, P. H.; Gupta, A.; Calvert, P.  G.  (1987) Mass accomodation coefficient for HO2
53           radicals on aqueous particles. J.  Geophys. Res. [Atmos.] 92: 4163-4170.
54


         April 1995                                    3_203        DRAFT-DO NOT QUOTE OR CITE

-------
 1       Mueller, P. K.  (1982) Atmospheric participate carbon observations in urban and rural areas of the United
 2           States. In: Wolff, G. T.; Klimisch, R. L., eds. Paniculate carbon—atmospheric life cycle: proceedings of
 3           an international symposium; October 1980; Warren, MI. New York, NY: Plenum Press.
 4
 5       Mulhbaier, J. L.; Williams, R. L. (1982) Fireplaces, furnaces, and vehicles as emission sources of paniculate
 6           carbons. In: Wolff, G. T.; Klimisch, R. L., eds. Paniculate carbon:  atmospheric life cycle.  New York,
 7           NY: Plenum Press; pp. 185-205.
 8
 9       Munger, J. W.; Jacob, D. J.; Waldman, J.  M.; Hoffmann, M. R. (1983) Fogwater chemistry in an urban
10           atmosphere. J. Geophys. Res. C:  Oceans Atmos. 88: 5109-5121.
11
12       Munger, J. W.; Jacob, D. J.; Hoffmann, M. R. (1984) The occurence of bisulfite-aldehyde addition products in
13           fog- and cloudwater.. J.  Atmos. Chem. 1: 335.
14
15       Munger, J. W.; Tiller, C.; Hoffman,  M. R. (1986) Identification of hydroxymethanesulfonate in fog water.
16           Science (Washington, DC) 231: 247-249.
17
18       Munger, J. W.; Collett,  J., Jr.; Daube, B.,  Jr.; Hoffmann, M. R. (1990) Fogwater chemistry at Riverside,
19           California.  Atmos. Environ. Part  B 24: 185-205.
20
21       Mylonas, D. T.; Allen, D. T.; Ehrman, S.  H.; Pratsinis, S. E.  (1991) The sources and size distributions of
22           organonitrates in Los Angeles aerosol. Atmos. Environ. Part A 25: 2855-2861.
23
24       National Research  Council. (1979) Airborne particles. Baltimore, MD: University Park Press.
25
26       National Research  Council. (1991) Rethinking the ozone problem in urban and regional air pollution.
27           Washington, DC: National Academy Press.
28
29       Natusch, D. F. S.; Wallace,  J. R.;  Evans, C. A., Jr. (1974) Toxic trace  elements:  preferential concentration in
30           respirable particles.  Science (Washington, DC) 183: 202-204.
31
32       Newman, L. (1981) Atmospheric oxidation of sulfur dioxide: a review as viewed from power plant and smelter
33           plume studies. Atmos. Environ. 15: 2231-2239.
34
35       Newman, L.; Forrest, J.; Manowitz, B.  (1975a) The application of an isotopic ratio technique to a study of the
36           atmospheric oxidation of sulfur dioxide in the plume from an oil-fired power plant. Atmos. Environ.
37           9: 959-968.
38
39       Newman, L.; Forrest, J.; Manowitz, B.  (1975b) The application of an isotopic ratio technique to a study of the
40           atmospheric oxidation of sulfur dioxide in the plume from a coal fired power plant. Atmos. Environ.
41           9: 969-977.
42
43       Nicholson, K. W.  (1988) The dry deposition of small particles:  a review of experimental measurements. Atmos.
44           Environ. 22: 2653-2666.
45
46       Noll, K. E.; Pilat, M. J. (1971) Size distribution of atmospheric giant particles. Atmos. Environ. 5: 527-540.
47
48       Noll, K. E.; Pontius, A.; Frey, R.; Gould, M. (1985) Comparison of atmospheric  coarse particles at an urban
49           and non-urban site. Atmos. Environ. 19:  1931-1943.
50
51       Noll, K. E.; Yuen, P.-F.; Fang, K. Y.-P. (1990) Atmospheric coarse paniculate concentrations and dry
52           deposition fluxes for ten metals in two urban environments. Atmos. Environ. Part A 24: 903-908.
53
         April 1995                                    3-204       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Noone, K. J. et al. (1992) Changes in aerosol size and phase distributions due to physical and chemical
  2          processes in fog. Tellus Ser. B 44: 489-504.
  3
  4      Noone, K. J.; Ogren, J. A.; Hallberg, A.; Hansson, H. C.; Wiedensohler, A.; Swietlicki, E.  (1992)
  5          A statistical examination of the chemical differences between interstitial and scavenged aerosol. Tellus Ser.
  6          B 44: 581-592.
  7
  8      Novakov, T. (1984) The role of soot and primary oxidants in the atmospheric chemistry. Sci.  Total Environ.
  9          36:  1-10.
 10
 11      Noxon, J. F. (1983) NO3 and NO2 in the mid-Pacific troposphere. J. Geophys. Res. C: Oceans Atmos.
 12          88:  11017-11021.
 13
 14      Nriagu, J. O. (1990) Global metal  pollution: poisoning the biosphere? Environment 32(7): 6-11, 28-33.
 15
 16      Nunes, T. V.; Pio, C. A. (1993) Carbonaceous aerosols in industrial and coastal atmospheres. Atmos. Environ.
 17          Part  A 27: 1339-1346.
 18
 19      O'Brien,  R. J.; Crabtree, J. H.; Holmes, J. R.; Hoggan, M. C.; Bockian, A. H. (1975) Formation of
 20          photochemical aerosol from hydrocarbons: atmospheric analysis. Environ. Sci. Technol. 9: 577-582.
 21
 22      O'Dell, R. A.; Taheri, M.; Kabel, R. L. (1977) A model for uptake of pollutants by vegetation. J.  Air Pollut.
 23          Control Assoc. 27: 1104-1109.
 24
 25      Ogren, J. A.; Charlson, R. J. (1992) Implications for models and measurements of chemical inhomogeneities
 26          among cloud droplets. Tellus 44: 208-225.
 27
 28      Olmez, I.; Sheffield, A. E.; Gordon, G. E.;  Houck, J. E.; Pritchett, L. C.; Cooper, J. A.; Dzubay, T. G.;
 29          Bennett, R.  L. (1988) Composition of particles  from selected sources in Philadelphia for receptor modeling
 30          applications. JAPCA 38: 1392-1402.
 31
 32      Olson, T. M.; Hoffman, M. R. (1989) Hydroxyalkylsulfonate formation: its role as a S(IV) reservoir in
 33          atmospheric water droplets. Atmos.  Environ. 23: 985-997.
 34
 35      Oppelt, E. T. (1987) Incineration of hazardous waste: a critical review.  JAPCA 37: 558-586.
 36
 37      Otani, Y.; Wang,  C. S. (1984) Growth and deposition of saline droplets covered with a monolayer of
 38          surfactant. Aerosol Sci.  Technol. 3: 155-166.
 39
 40      Pace, T. G., ed. (1986) Transactions, receptor methods for source apportionment: real world issues and
 41          applications. Pittsburgh, PA: Air Pollution Control Association.
 42
 43      Pack, D.  H.; Ferber, G. J.;  Heffter, J.  L.; Telegadas, K.; Angel 1, J. K.; Hoecker, W. H.; Machta, L. (1978)
44          Meteorology of  long-range transport. Atmos. Environ. 12: 425-444.
45
46      Pandis, S. N.; Seinfeld, J. H. (1989a) Sensitivity analysis of a chemical mechanism for aqueous-phase
47          atmospheric chemistry. J. Geophys.  Res. [Atmos.] 94: 1105-1126.
48
49      Pandis, S. N.; Seinfeld, J. H. (1989b) Mathmatical modeling of acid deposition due to radiation fog.
50          J. Geophys. Res. 94: 12911-12923.
51
52      Pandis, S. N.; Seinfeld, J. H.;  Pilinis, C. (1990a) The smog-fog-smog cycle and acid deposition. J.  Geophys.
53          Res. 95: 18489-18500.
54


         April 1995                                    3-205       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Pandis, S. N.; Seinfeld, J. H.; Pilinis, C. (1990b) Chemical composition differences in fog and cloud droplets
  2          of different sizes. Atmos. Environ. Part A 24: 1957-1969.
  3
  4      Pandis, S. N.; Paulson, S. E.; Seinfeld, J. H.; Flagan, R. C.  (1991) Aerosol formation in the photooxidation of
  5          isoprene and /?-pinene. Atmos. Environ. Part A 25: 997-1008.
  6
  7      Pandis, S. N.; Harley, R. A.; Cass, G. R.; Seinfeld, J. H. (1992a) Secondary organic aerosol formation and
  8          transport. Atmos. Environ.  Part A  26: 2269-2282.
  9
10      Pandis, S. N.; Seinfeld, J. H.; Pilinis, C. (1992b) Heterogeneous sulfate production in an urban fog. Atmos.
11          Environ.  Part A 26:  2509-2522.
12
13      Pandis, S. N.; Wexler, A. S.; Seinfeld, J. H. (1993) Secondary organic aerosol formation and transport—II.
14          predicting the ambient secondary organic aerosol size distribution. Atmos. Environ. Part A 27: 2403-2416.
15
16      Pankow, J. F. (1987) Review and comparative analysis of the theories on partitioning between the gas and
17          aerosol paniculate phases in the atmosphere.  Atmos. Environ. 21: 2275-2283.
18
19      Pankow, J. F. (1988) Gas phase retention volume behavior of organic compounds on the sorbent
20          poly(oxy-/n-terphenyl-2'-5'-ylene).  Anal.  Chem. 60: 950-958.
21
22      Pankow, J. F. (1989) Overview of the gas phase retention volume behavior of organic compounds on
23          polyurethane foam. Atmos.  Environ.  23:  1107-1111.
24
25      Pankow, J. F. (1991) Common y-intercept and single compound regressions  of gas-particle partitioning data
26          vs. l/T. Atmos. Environ. Part A: 25: 2229-2239.
27
28      Pankow, J. F. (1994a) An absorption model of gas/particle partitioning of organic compounds in the
29          atmosphere. Atmos.  Environ. 28: 185-188.
30
31      Pankow, J. F. (1994b) An absorption model of the gas/aerosol partitioning involved in the formation of
32          secondary organic aerosol. Atmos.  Environ. 28: 189-193.
33
34      Pankow, J. F.; Bidleman, T. F. (1991)  Effects of temperature, TSP and per  cent non-exchangeable material in
35          determining the gas-particle partitioning of organic compounds. Atmos. Environ. Part A 25: 2241-2249.
36
37      Pankow, J. F.; Bidleman, T. F. (1992)  Interdependence of the slopes and intercepts from log-log correlations of
38          measured  gas-particle partitioning and vapor pressure-I. Theory and analysis of available data. Atmos
39          Environ.  Part A 26:  1071-1080.
40
41      Pankow, J. F.; Storey, J. M. E.; Yamasaki, H. (1993) Effects of relative humidity on gas/particle partitioning
42          of semivolatile organic compounds  to urban paniculate matter. Environ.  Sci. Technol. 27: 2220-2226.
43
44      Parrish, D. D.; Fahey, D. W.; Williams, E. J.; Liu, S. C.; Trainer, M.; Murphy,  P. C.; Albritton, D. L.;
45          Fehsenfeld, F. C. (1986) Background ozone and anthropogenic ozone enhancement at Niwot Ridge,
46          Colorado. J. Atmos.  Chem. 4: 63-80.
47
48      Pastuszka, J. S.; Kwapulinski, J. (1988) The change in mass size distribution of aerosol near dumps as  a result
49          of resuspension of dust. Presented at: the 81st annual meeting of Air Pollution Control Association; June;
50          Dallas, TX.
51
52      Patterson,  E. M.; Gillette, D. A. (1977a) Commonalities in measured size distributions for aerosols having a
53          soil-derived component. J. Geophys. Res. 82: 2074-2082.
54


         April 1995                                    3-206       DRAFT-DO  NOT QUOTE OR CITE

-------
  1      Patterson, E. M.; Gillette, D. A. (1977b) Measurements of visibility vs. mass-concentration for airborne soil
  2          particles. Atmos. Environ. 11:  193-196.
  3
  4      Patterson, R. K.; Wagman, J. (1977) Mass and composition of an urban aerosol as a function of particle size
  5          for several visibility levels. J. Aerosol Sci.  8: 269-279.
  6
  7      Patterson, E. M.; Gillette, D. A.; Grams, G. W. (1976) On the relation between visibility and concentration for
  8          airborne soil particles. J. Appl. Meteorol. 15: 470-478.
  9
 10      Patterson, E. M.; Gillette, D. A.; Stockton, B.  H. (1977) Complex index of refraction between 300 and 700 nm
 11          for Saharan aerosols. J. Geophys. Res. 82:  3153-3159.
 12
 13      Peirson, D.  H.; Cawse, P. A.; Salmon, L.; Cambray, R. S. (1973) Trace elements in the atmospheric
 14          environment. Nature (London) 241: 252-256.
 15
 16      Penkett, S. A. (1972) Oxidation of SO2 and other atmospheric gases by ozone in aqueous solution. Nature
 17          (London) 240: 105-106.
 18
 19      Penkett, S. A.; Jones,  B.  M. R.; Brice, K.  A.;  Eggleton, A. E. J.  (1979) The importance of atmospheric ozone
 20          and hydrogen peroxide in oxidising sulphur dioxide in cloud and rainwater. Atmos. Environ. 13: 123-137.
 21
 22      Penner, J. E.; Eddleman, H.; Novakov, T. (1993) Towards the development of a global inventory for black
 23          carbon emissions.  Atmos. Environ.  Part A 27: 1277-1295.
 24
 25      Perry,  G.  B.; Chai, H.; Dickey, D. W.; Jones,  R. H.; Kinsman, R. A.; Morrill, C. G.; Spector, S. L.;
 26          Weiser, P. C.  (1983) Effects of paniculate air pollution on asthmatics. Am. J. Public Health 73: 50-56.
 27
 28      Peters, K.; Eiden, R. (1992) Modelling the dry  deposition velocity of aerosol particles to a  spruce forest.
 29          Atmos.  Environ. Part A 26: 2555-2564.
 30
 31      Peterson, J.  T. (1970)  Distribution of sulfur dioxide over metropolitan St. Louis, as described by empirical
 32          eigenvectors, and its relation to meteorological parameters. Atmos. Environ. 4: 501-518.
 33
 34      Peterson, J.; Junge, C. (1971) Sources  of paniculate matter in the atmosphere.  In: Matthews, W.; Kellogg, W.;
 35          Robinson, G. D., eds. Man's impact on the climate. Cambridge, MA: MIT Press.
 36
 37      Phalen, R., ed. (1985) Particle size-selective sampling in the workplace:  report  of the ACGIH technical
 38          committee on air sampling procedures. Cincinnati, OH: American Conference of Governmental Industrial
 39          Hygienists.
 40
 41      Pickle, T.; Allen, D. T.; Pratsinis, S. E. (1990) The sources and size distributions of aliphatic and carbonyl
 42          carbon in Los Angeles aerosol.  Atmos.  Environ. Part A 24: 2221-2228.
 43
 44      Pielke, R.  A. (1984) Mesoscale meteorological modeling. Orlando, FL: Academic Press.
45
46      Pielke, R.  A. (1989) Status of subregional and mesoscale models, v. 2: mesoscale meteorological models in the
47          United States. Fort Collins, CO: Electric Power Research  Institute; report no. EN-6649, v. 2.
48
49      Pierson, W.  R.; Brachaczek, W. W. (1983) Paniculate matter  associated with vehicles on the road. II. Aerosol
50          Sci. Technol. 2: 1-40.
 51
52      Pierson, W.  R.; Gertler, A. W.; Bradow, R. L. (1990) Comparison of the SCAQS tunnel study with other
53          on-road  vehicle emission data. J. Air Waste Manage. Assoc. 40:  1495-1504.
 54


         April  1995                                   3_207      DRAFT-DO NOT QUOTE OR CITE

-------
  1       Pilinis, C. (1990) Derivation and numerical solution of the species mass distribution equations for
  2           multicomponent paniculate systems. Atmos. Environ. Part A 24: 1923-1928.
  3
  4       Pilinis, C.; Seinfeld, J. H. (1987) Continued development of a general equilibrium model for inorganic
  5           multicomponent atmospheric aerosols. Atmos. Environ. 21: 2453-2466.
  6
  7       Pilinis, C.; Seinfeld, J. H. (1988) Development and evaluation of an Eulerian photochemical gas-aerosol model.
  8           Atmos. Environ. 22:  1985-2001.
  9
10       Pilinis, C.; Seinfeld, J. H.; Grosjean, D. (1989) Water content of atmospheric aerosols. Atmos. Environ.
11           23:  1601-1606.
12
13       Pinnick, R. G.; Fernandez, G.; Hinds, B. D.;  Bruce, C. W.; Schaefer, R. W.; Pendleton, J. D. (1985) Dust
14           generated by vehicular traffic on unpaved roadways: sizes and infrared extinction characteristics. Aerosol
15           Sci. Technol. 4: 99-121.
16
17       Pinnick, R. G.; Fernandez, G.; Martinez-Andazola, E.; Hinds,  B. D.; Hansen, A. D.  A.; Fuller, J.  (1993)
18           Aerosol in the arid southwestern United States: measurements of mass loading, volatility, size distribution,
19           absorption characteristics, black carbon content, and vertical structure to 7 km above sea level. J. Geophys.
20           Res. 98: 2651-2666.
21
22       Pires, M.; Rossi, M. J.; Ross, D. S. (1994) Kinetic and mechanistic aspects  of the NO oxidation by  O2 in
23           aqueous phase. Int. J. Chem. Kinet. 26: 1207-1227.
24
25       Pitchford, M. L.; McMurry,  P.  H. (1994) Relationship between measured water vapor growth and chemistry of
26           atmospheric aerosol for Grand Canyon, Arizona, in winter 1990. Atmos. Environ. 28: 827-839.
27
28       Pitts, J.  N., Jr.; Atkinson, R.; Sweetman, J. A.; Zielinska, B. (1985) The gas-phase reaction of naphthalene
29           with N2O5 to form nitronaphthalenes. Atmos. Environ. 19: 701-705.
30
31       Placet, M.; Battye, R. E.;  Fehsenfeld, F. C.; Bassett,  G. W. (1991) Emissions involved  in acidic deposition
32           processes.  In: Irving, P.  M., ed. Acidic depostion: state of science and technology, volume I: emissions,
33           atmospheric processes, and deposition. Washington,  DC: The U.S. National Acid Precipitation Assessment
34           Program. (State of science and technology  report no. 1).
35
36       Pleim, J.; Venkatram, A.;  Yamartino, R. J. (1984) The dry deposition model, v. 4. Rexdale, ON, Canada:
37           Ontario Ministry of the Environment, ADOM/TADAP Model Development Program.
38
39       Post, J.  E.; Buseck, P. R.  (1985) Quantitative energy-dispersive analysis of lead halide particles from the
40           Phoenix urban aerosol. Environ. Sci. Technol.  19: 682-685.
41
42       Pratsinis, S.; Ellis, E. C.; Novakov, T.; Friedlander, S.  K. (1984) The carbon containing component of the  Los
43           Angeles aerosol: source apportionment and contributions to the visibility  budget. J. Air Pollut. Control
44           Assoc. 34: 643-650.
45
46       Pratsinis, S. E.; Zeldin, M. D.;  Ellis, E. C. (1988) Source resolution of the  fine carbonaceous aerosol by
47           principal component-stepwise regression analysis. Environ. Sci. Technol. 22: 212-216.
48
49       Pruppacher, H.  R.; Klett, J. D. (1980) Microphysics of clouds and precipitation. Boston, MA: D. Reidel.
50
51       Pueschel, R. F.; Van Valin, C. C.;  Castillo, R. C.; Kadlecek, J. A.; Ganor, E. (1986) Aerosols in polluted
52           versus nonpolluted airmasses: long-range transport and effects on clouds. J. Clim.  Appl. Meteorol.
53           25:  1908-1917.
54


         April 1995                                     3-208       DRAFT-DO NOT QUOTE  OR CITE

-------
  1      Rabano, E. S.; Castillo, N. T.; Torre, K. J.; Solomon, P. A. (1989) Speciation of arsenic in ambient aerosols
  2          collected in Los Angeles. JAPCA 39: 76-80.
  3
  4      Rader, D. J.; McMurry, P. H.  (1986) Application of the tandem differential mobility analyzer to studies of
  5          droplet growth or evaporation. J. Aerosol Sci. 17: 771-787.
  6
  7      Radke, L. F.; Hobbs, P. V. (1991) Humidity and particle fields around some small cumulus clouds. J. Atmos.
  8          Sci. 48: 1190-1193.
  Q
 10      Raga, G. B.; Jonas, P. R. (1993) On the link between cloud radiative properties and subcloud aerosol
 11          concentrations. Q. J. R. Meteorol. Soc. 119: 1419-1425.
 12
 13      Ranade, M.  B. (Arun); Woods, M. C.; Chen, F.-L.; Purdue, L. J.; Rehme, K. A. (1990) Wind tunnel
 14          evaluation of PM10 samplers. Aerosol Sci. Technol. 13:  54-71.
 15
 16      Rao, A. K.; Whitby, K. T. (1978) Non-ideal collection characteristics of inertial impactors—I. single-stage
 17          impactors and solid particles. J. Aerosol Sci. 9: 77-86.
 18
 19      Rau, J. A. (1989) Composition  and size distribution of residential wood smoke particles. Aerosol Sci. Technol.
 20          10:  181-192.
 21
 22      Reeks, M. W.; Reed, J.; Hall, D. (1988) On the resuspension of small particles by a turbulent flow. J. Phys.
 23          D: Appl. Phys. 21: 574-589.
 24
 25      Richards, L. W.  (1983) Comments on the oxidation of NO2 to nitrate—day and night. Atmos. Environ.
 26          17: 397-402.
 27
 28      Richards, L. W.; Anderson, J. A.; Blumenthal, D. L.; Brandt, A. A.;  McDonald, J. A.; Waters,  N.; Macias,
 29          E. S.; Bhardwaja, P. S. (1981) The chemistry, aerosol physics, and optical properties of a western
 30          coal-fired power plant plume. Atmos. Environ.  15: 2111-2134.
 31
 32      Richards, L. W.; Blanchard, C. L.; Blumenthal, D. L. (1991) Navajo generating station visibility study,  draft
 33          #2. Santa Rosa,  CA: Sonoma Technology, Inc.; final report STI-90200-1124-FRD2.
 34
 35      Richardson, C. B.; Spann, J. F. (1984) Measurement of the water cycle in a levitated ammonium sulfate
 36          particle. J. Aerosol Sci.  15: 563-571.
 37
 38      Roberts, D.; Williams, D. (1979) The kinetics of oxidation of sulphur dioxide within the plume from a sulphide
 39          smelter in a remote region.  Atmos. Environ. 13: 1485-1499.
 40
 41       Rodes, C.; Holland, D.; Purdue, L.; Rehme, K. (1985) A field comparison of PM10 inlets at four  locations.
 42          J. Air Pollut. Control Assoc. 35: 345-354.
 43
 44      Roelofs, G. J. (1992a) Drop size dependent sulfate distribution in a growing cloud. J. Atmos. Chem.
45           14: 109-118.
 46
47       Roelofs, G. J. H. (1992b) On the drop and aerosol size dependence of aqueous sulfate formation in a
48           continental cumulus cloud. Atmos. Environ.  Part A 26: 2309-2321.
49
 50       Roelofs, G. J. H. (1993) A cloud chemistry sensitivity study  and comparison of explicit and bulk cloud model
51           performance. Atmos. Environ. Part A 27: 2255-2264.
52
53       Rogge, W. F. (1993) Molecular tracers for sources of atmospheric carbon particles: measurements and model
54           predictions [Ph.D. thesis]. Pasadena, California: California Institute of Technology.


         April 1995                                    3-209        DRAFT-DO  NOT QUOTE OR CITE

-------
  1       Rogge, W. F.; Hildemann, L. M.; Mazurek, M. A.; Cass, G. R.; Simoneit, B. R. T. (1991) Sources of fine
  2           organic aerosol. 1. Charbroilers and meat cooking operations. Environ. Sci. Technol. 25: 1112-1125
  3
  4       Rogge, W. F.; Hildemann, L. M.; Mazurek, M. A.; Cass, G. R.; Simoneit, B. R. T. (1993a) Sources of fine
  5           organic aerosol. 2. Noncatalyst and catalyst-equipped automobiles and heavy-duty diesel trucks.  Environ.
  6           Sci. Technol. 27: 636-651.
  7
  8       Rogge, W. F.; Hildemann, L. M.; Mazurek, M. A.; Cass, G. R.; Simoneit, B. R. T. (1993b) Sources of fine
  9           organic aerosol. 3. Road dust, tire debris, and organometallic brake lining dust: roads as sources and sinks.
10           Environ. Sci. Technol. 27: 1892-1904.
11
12       Rogge, W. F.; Hildemann, L. M.; Mazurek, M. A.; Cass, G. R.; Simoneit, B. R. T. (1993c) Sources of fine
13           organic aerosol. 4. Paniculate abrasion products from leaf surfaces of urban plants. Environ. Sci. Technol.
14           27:2700-2711.
15
16       Rogge, W. F.; Mazurek, M. A.; Hildemann, L. M.; Cass, G. R.; Simoneit, B. R. T. (1993d) Quantification of
17           urban organic aerosols at a molecular level: identification, abundance and seasonal variation. Atmos.
18           Environ. Part A 27: 1309-1330.
19
20       Rogge, W. F.; Hildemann, L. M.; Mazurek, M. A.; Cass, G. R. (1994) Sources of fine organic aerosol:
21           6. cigarette smoke in the urban atmosphere. Environ. Sci. Technol. 28: 1375-1388.
22
23       Rojas, C. M.; Injuk, J.; Van Grieken, R. E.; Laane, R.  W. (1993) Dry and wet deposition fluxes of Cd, Cu,
24           Pb and Zn into the Southern Bight of the North Sea. Atmos. Environ. Part A 27: 251-259.
25
26       Rolph, G. D.; Draxler, R. R. (1990) Sensitivity of three-dimensional trajectories to the spatial and temporal
27           densities of the wind field. J.  Appl. Meteorol. 29: 1043-1054.
28
29       Rood, M. J.; Larson, T. V.; Covert, D. S.; Ahlquist, N. C.  (1985) Measurement of laboratory and  ambient
30           aerosols with temperature and humidity controlled nephelometry. Atmos. Environ. 19: 1181-1190.
31
32       Rood, M. J.; Shaw,  M. A.; Larson, T. V.; Covert, D. S. (1989) Ubiquitous nature of ambient metastable
33           aerosol. Nature (London) 337: 537-539.
34
35       Roth,  P. M. (1992) [Personal  ommunication].
36
37       Rounds, S. A.; Pankow, J. F. (1990) Application of a radial diffusion model to describe gas/particle sorption
38           kinetics. Environ. Sci. Technol. 24: 1378-1386.
39
40       Rubel, G. O. (1991) Partitioning of partially soluble volatiles between the vapor and liquid aerosol phase.
41           Atmos.  Environ. Part  A 25: 1009-1012.
42
43       Rubel, G. 0.; Gentry, J. W. (1984) Measurement of the kinetics of solution droplets in the  presence of
44           adsorbed monolayers:  determination of water accommodation coefficients. J. Phys. Chem. 88: 3142-3148.
45
46       Russell,  A. G.; Cass, G. R. (1986) Verification of a mathematical model for aerosol nitrate and nitric acid
47           formation and its use for control measure evaluation. Atmos. Environ. 20: 2011-2025.
48
49       Samson, P. J.; Ragland, K. W. (1977) Ozone and visibility reduction in the Midwest: evidence for large-scale
50           transport. J. Appl. Meteorol. 16: 1101-1106.
51
52       Samson, R.  J.; Mulholland, G. W.; Gentry, J. W. (1987) Structural analysis of soot agglomerates. Langmuir
53           3: 272-281.
54


         April 1995                                    3.210       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Satsumakayashi, H.; Kurita, H.; Yokouchi, Y.; Ueda, H, (1989) Mono and dicarboxylic acids under long-range
  2          transport of air pollution in central Japan. Tellus Ser. B. 41: 219-229.
  3
  4       Satsumakayashi, H.; Kurita, H.; Yokouchi, Y.; Ueda, H. (1990) Photochemical formation of paniculate
  5          dicarboxylic acids under long-range transport in central Japan.  Atmos. Environ. Part A 24: 1443-1450.
  6
  7       Savoie, D. L.; Prospero, J. M.; Nees, R. T. (1987) Washout ratios of nitrate, non-sea-salt sulfate and sea-salt
  8          on Virginia Key, Florida and on American Samoa. Atmos. Environ. 21: 103-112.
  9
 10       Saxena, V. K.; Hendler, A. H. (1983) In-cloud scavenging and resuspension of cloud active aerosols during
 11          winter storms over Lake Michigan. In: Pruppacher, H. R.; Semonin, R. G.; Slinn,  W. G. N., eds.
 12          Precipitation scavenging, dry deposition and  resuspension. New York, NY: Elsevier; pp. 91-102.
 13
 14       Saxena, P.; Peterson, T. W. (1981) Thermodynamics of multicomponent electrolytic aerosols. J. Colloid
 15          Interface Sci. 79: 496-510.
 16
 17       Saxena, P.; Hudischewskyj, A. B.; Seigneur, C.; Seinfeld, J. H. (1986) A comparative  study of equilibrium
 18          approaches to the chemical characterization of secondary aerosols. Atmos. Environ. 20:  1471-1483.
 19
 20       Saxena, P.; Mueller, P. K.; Kim, Y. P.; Seinfeld, J. H.; Koutrakis, P. (1993) Coupling thermodynamic theory
 21          with measurements to characterize acidity of atmospheric particles. Aerosol Sci. Technol. 19: 279-293.
 22
 23       Schaefer, D. W.; Hurd, A. J. (1990) Growth and structure of combustion aerosols: fumed silica. Aerosol Sci.
 24          Technol. 12: 876-890.
 25
 26       Schroeder,  W. H.; Dobson, M.; Kane, D. M.; Johnson,  N. D. (1987) Toxic trace elements associated with
 27          airborne paniculate matter: a review. JAPCA 37: 1267-1285.
 28
 29       Schuetzle, D.; Cronn, D.; Crittenden, A. L.; Charlson, R. J. (1975) Molecular composition of secondary
 30          aerosol and its possible origin. Environ. Sci.  Technol. 9: 838-845.
 31
 32       Schwartz, S. E. (1979)  Residence times in reservoirs under non-steady state conditions: application to
 33          atmospheric S02 and aerosol sulfate. Tellus 31:  530-547.
 34
 35       Schwartz, S. E. (1984)  Gas-aqueous reactions of  sulfur and nitrogen oxides in liquid-water clouds. In: Calvert,
 36          J.  G., ed.  SO2, NO and NO2 oxidation mechanisms:  atmospheric considerations. Boston, MA: Butterworth
 37          Publishers; pp.  173-208. (Teasley, J. I., ed.  Acid precipitation series: v. 3).
 38
 39       Schwartz, S. E. (1984)  Gas- and aqueous-phase chemistry of HO2 in liquid water clouds. J. Geophys. Res.
 40          89: 11589-11598.
 41
 42      Schwartz, S. E. (1986) Mass-transport considerations pertinent to aqueous-phase reactions of gases in
 43           liquid-water clouds. In: Jaeschke, W., ed. Chemistry of multiphase atmospheric systems. Heidelberg,
 44          Federal Republic of Germany: Springer; pp.  415-471.
 45
 46       Schwartz, S. E. (1987) Both sides now: the chemistry of clouds. Ann.  N.  Y. Acad. Sci.  502- 83-144
47
48       Schwartz, S. E. (1988) Mass-transport limitation to the rate of in-cloud oxidation of SO2: re-examination in the
49           light of new data. Atmos. Environ. 22: 2491-2499.
50
51       Schwartz, S. E. (1989) Acid deposition: unraveling a regional phenomenon.  Science (Washington, DC)
52           243:  753-763.
53
         April 1995                                    3_2H       DRAFT-DO NOT QUOTE OR CITE

-------
 1       Schwartz, S. E. (1991) Aqueous phase reactions. In: Irving, P. M., ed. Acidic deposition: state of science and
 2           technology, volume I: emissions, atmospheric processes, and deposition. Washington, DC: The U.S.
 3           National Acid Precipitation Assessment Program; pp. 2-93—2-104. (Hicks,  B. B.; Draxler, R. R.; Dodge,
 4           M.; Hales, J. M.; Albritton, D. L.; Schwartz, S. E.; Meyers, T. P.; Fehsenfeld, F. C.; Tanner, R. L.;
 5           Vong, R. J. Atmospheric processes research and process model development: state of science and
 6           technology report no. 2).
 7
 8       Schwartz, S. E. (1994) Cloud studies seen from a physical chemist's perspective. In: Angeletti, G.; Restelli,
 9           G., eds. Physico-chemical behaviour of atmospheric pollutants, v. 2. Proceedings of the sixth European
10           symposium; October 1992; Varese, Italy. Luxembourg: Office of Official Publications of the European
11           Commission; pp. 891-900.
12
13       Schwartz, S. E.; Freiberg, J. E. (1981) Mass-transport  limitation to the rate of reaction of gases in liquid
14           droplets: application to oxidation of SO2 in aqueous solutions. Atmos. Environ. 15: 1129-1144.
15
16       Schwartz, S. E.; Newman,  L.  (1978) Processes limiting oxidation of sulfur dioxide in stack plumes. Environ.
17           Sci. Technol. 12: 67-73.
18
19       Schwartz, S. E.; Newman,  L.  (1983) Measurements of sulfate production in natural clouds [discussion of Hegg
20           and Hobbs (1982)]. Atmos. Environ. 17: 2629-2632.
21
22       Schwartz, S. E.; White, W. H. (1983) Kinetics of reactive dissolution of nitrogen oxides into aqueous solution.
23           In: Advances in environmental science and technology, v. 12. New York, NY: John Wiley and Sons;
24           pp. 1-116.
25
26       Scire, J. S.; Lurmann, F. W.; Bass, A.; Hanna,  S. R.  (1983) Development of the MESOPUFF II  dispersion
27           model. Research Triangle  Park, NC: U.S. Environmental Protection Agency, Environmental Sciences
28           Research Laboratory; EPA report no. EPA-600/3-84-057. Available from: NTIS, Springfield,  VA;
29           PB84-184753.
30
31       Scott, B. C.; Laulainen, N. S. (1979) On the concentration of sulfate in precipitation. J. Appl. Meteorol.
32           18: 138-147.
33
34       Seaman, N. L. (1990) Meteorological modeling applied to regional air-quality studies using four-dimensional
35           data assimilation. In: Proceedings of the IBM summer institute on environmental modeling; July; Oberlech,
36           Austria.
37
38       Sega,  K.; Fugas, M. (1984) Seasonal and spatial differences in mass concentration levels and particle size
39           distribution of aerosols over an urban area. Atmos. Environ. 18: 2433-2437.
40
41       Sehmel, G.  A. (1973) Particle resuspension from an asphalt road caused by car  and truck traffic. Atmos.
42           Environ. 7: 291-309.
43
44       Sehmel, G.  A. (1980) Particle and gas dry deposition: a review. Atmos. Environ. 14: 983-1011.
45
46       Sehmel, G.  (1984) Deposition and resuspension.  In: Atmospheric science and power production. Washington,
47           DC: U.S. Department of Energy, Office  of Science and Technical Information; pp. 533-583; report no.
48           DOE/TIC-27601.
49
50       Seidl, W. (1989) Ionic concentrations and initial S(IV) oxidation rates in droplets during the condensation stage
51           of a cloud. Tellus Ser.  B 41:  32-50.
52
53       Seigneur, C.;  Saxena, P. (1984) A study of atmospheric acid formation in different environments. Atmos.
54           Environ. 18: 2109-2124.


         April 1995                                    3-212       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Seinfeld, J. H. (1986) Atmospheric chemistry and physics of air pollution. New York, NY: John Wiley and
  2          Sons.
  3
  4      Seinfeld, J. H. (1988) Ozone air quality models: a critical review. JAPCA 38: 616-645.
  5
  6      Seinfeld, J. H.; Bassett, M. (1982) Effect of the mechanism of gas-to-particle conversion on the evolution of
  7          aerosol size distributions. In: Heterogeneous atmospheric chemistry. Washington, DC: American
  8          Geophysical Union; pp. 6-12.
  9
 10      Seinfeld, J. H.; Flagan, R. C.; Petti, T. B.;  Stern, J. E.; Grosjean, D. (1987) Aerosol  formation in aromatic
 11          hydrocarbon-NOx systems. Pasadena, CA: California Institute of Technology; project no. AP-6.
 12
 13      Sergides, C. A.; Jassim, J. A.; Chughtai, A. R.; Smith, D.  M. (1987) Appl. Spectrosc. 41: 482-492.
 14
 15      Shah, J. J.; Johnson, R. L.; Heyerdahl, E. K.; Huntzicker, J.  (1986) Carbonaceous aerosol at urban and rural
 16          sites in the United States. J. Air Pollut. Control Assoc. 36: 254-257.
 17
 18      Shao, Y.; Raupach, M.; Findlater, P.  (1993) The effect of saltation bombardment on the entrainment of dust by
 19          wind. J. Geophys. Res.: 12719-12726.
 20
 21      Shaw, G. E. (1989) Production of condensation nuclei in clean air by nucleation of H2SO4. Atmos. Environ.
 22          23:  2841-2846.
 23
 24      Shaw, M. A.;  Rood, M. J. (1990) Measurement of the crystallization humidities of ambient aerosol particles.
 25          Atmos. Environ. Part A 24: 1837-1841.
 26
 27      Sheridan, P. J.; Schnell, R. C.; Kahl, J. D.; Boatman, J. F.; Garvey, D. M. (1993) Microanalysis of the
 28          aerosol collected over south-central New Mexico during the ALIVE field experiment, May-December 1989.
 29          Atmos. Environ. Part A 27: 1169-1183.
 30
 31      Shi, B.;  Seinfeld, J. H. (1991) On mass transport limitation to  the rate of reaction of gases in liquid droplets.
 32          Atmos. Environ. Part A 25: 2371-2383.
 33
 34      Shinn, J.; Homan,  D.; Gay, D. (1983) Plutonium aerosol fluxes and pulmonary exposure rates during
 35          resuspension from base soils near  a chemical separation facility. In: Pruppacher, H.; Sermonin, R.; Slinn,
 36          W., eds. Precipitation scavenging, dry deposition and resuspension. Amsterdam,  The Netherlands: Elsevier;
 37          pp.  1131-1143.
 38
 39      Showman, R. E.; Hendricks, J. C. (1989) Trace element content of Flavoparmelia caperata (L.) hale due to
 40          industrial emissions. JAPCA 39: 317-320.
 41
 42      Sievering, H.;  Van Valin, C. C.;  Barrett, E.  W.; Pueschel, R. F. (1984) Cloud scavenging of aerosol sulfur:
 43           two  case studies. Atmos. Environ. 18: 2685-2690.
44
45       Sievering, H.;  Boatman, J.; Galloway, J.; Keene, W.;  Kim,  Y.; Luria, M.; Ray, J. (1991) Heterogeneous
46           sulfur conversion in sea-salt aerosol particles: the role of aerosol water content and  size distribution. Atmos.
47           Environ. Part A 25: 1479-1487.
48
49       Sievering et al. (1994) dry deposition.
50
51       Sievering, H.;  Gorman, E.; Kim, Y.; Ley, T.; Seidl, W.; Boatman, J. (1994) Heterogeneous conversion
52           contribution to  the sulfate observed over Lake Michigan. Atmos. Environ. 28: 367-370
53
         April 1995                                    3_213       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Simoneit, B. R. T. (1984) Organic matter of the troposphere—III. characterization and sources of petroleum and
  2           pyrogenic residues in aerosols over the western United States. Atmos. Environ. 18: 51-67.
  3
  4       Simoneit, B. R. T. (1986) Characterization of organic constituents in aerosols in relation to their origin and
  5           transport: a review. Presented at: the 2nd workshop on chemistry analysis of hydrocarbons in the
  6           environment;  November 1985; Barcelona, Spain. Int.  J. Environ. Anal. Chem. 23:  207-237.
  7
  8       Simoneit, B. R. T. (1989) Organic matter in the troposphere: V. application of molecular marker analysis to
  9           biogenic emissions into  the troposphere for  sources reconciliations.. J. Atmos. Chem.: 251-278.
10
11       Simoneit, B. R. T.; Mazurek, M. A. (1982) Organic matter of the troposphere—II. natural background of
12           biogenic lipid matter in  aerosols over the rural western United States. Atmos. Environ. 16: 2139-2159.
13
14       Sisler, J. F.; Malm, W. C.  (1994) The relative  importance of soluble aerosols to spatial and seasonal trends of
15           impaired visibility in the United States. Atmos. Environ. 28: 851-862.
16
17       Sisterson, D. L.; Bowersox, V. C.; Olsen,  A. R.; Meyers, T. P.; Olsen, A. R.; Vong, R.  L. (1991)
18           Deposition monitoring:  methods and results. In: Irving, P. M., ed. Acidic deposition: state of science and
19           technology, volume I: emissions, atmospheric processes, and depostition. Washington, DC: The U.S.
20           National Acid Precipitation Assessment Program. (State of science and technology report no. 6).
21
22       Slinn, W. G. N. (1983) Air-to-sea transfer of particles. In: Liss, P.  S.; Slirm, W. G.  N., eds. Air-to-sea
23           exchange of gases and particles. Hingham, MA: D. Reidel; pp.  299-405.
24
25       Sloane, C. S.  (1982) Visibility trends—II. mideastern United States 1948-1978. Atmos. Environ.
26           16:2309-2321.
27
28       Sloane, C. S.  (1983) Summertime visibility declines: meteorological influences. Atmos.  Environ. 17: 763-774.
29
30       Sloane, C. S.; Groblicki, P. J. (1981) Denver's  visibility history. Atmos. Environ. 15: 2631-2638.
31
32       Sloane, C. S.; Watson, J.; Chow, J.; Pritchett,  L.;  Richards, L. W. (1991) Size-segregated fine particle
33           measurements by chemical species and their impact on visibility impairment in Denver. Atmos. Environ.
34           Part A 25: 1013-1024.
35
36       Smith, T. B. (1981) Some observations of pollutant transport associated with elevated plumes. Atmos. Environ.
37           15: 2197-2203.
38
39       Smith, D. M.; Akhter, M. S.; Jassim, J. A.; Sergides, C. A.; Welch, W. F.; Chughtai, A. R. (1989) Studies
40           of the structure and reactivity of soot. Aerosol Sci. Technol. 10: 311-325.
41
42       Snider, J. R.; Vali, G. (1994) Sulfur dioxide oxidation in winter orographic clouds. J. Geophys. Res. [Atmos.]
43           99:  18713-18733.
44
45       Solomon, P. A.; Larson, S.  M.; Fall, T.; Cass,  G. R. (1988) Basinwide nitric acid and related species
46           concentrations observed during the Claremont Nitrogen Species Comparison Study.  Atmos. Environ.
47           22:  1587-1594.
48
49       Solomon, P. A.; Fall,  T.; Salmon, L.; Cass, G. R.; Gray, H. A.; Davidson, A. (1989) Chemical
50           characteristics of PM10 aerosols collected in the Los Angeles area. JAPCA 39: 154-163.
51
52       Solomon, P. A.; Salmon,  L. G.; Fall, T.; Cass, G. R. (1992) Spatial and temporal distribution of atmospheric
53           nitric acid and paniculate  nitrate concentrations in the Los Angeles area. Environ. Sci. Technol.
54           26:  1596-1601.


         April 1995                                     3-214      DRAFT-DO NOT QUOTE OR CITE

-------
  1      Spann, J. F.; Richardson, C. B. (1985) Measurement of the water cycle in mixed ammonium acid sulfate
  2          particles. Atmos. Environ. 19: 819-825.
  3
  4      Squires, P. (1958a) The microstructure and colloidal stability of warm clouds: I. the relation between structure
  5          and stability. Tellus 10: 256-261.
  6
  7      Squires, P. (1958b) The microstructure and colloidal stability of warm clouds: II. the causes of the variations in
  8          microstructure. Tellus  10: 262-271.
  9
 10      Standley, L. J.; Simoneit, B. R. T. (1987) Characterization of extractable plant wax, resin, and thermally
 11          matured components in smoke particles from prescribed burns. Environ. Sci.  Technol. 21:  163-169.
 12
 13      Stein, S. W.; Turpin, B. J.; Cai, X.; Huang, P.-F.; McMurry, P. H. (1994) Measurements of relative
 14          humidity-dependent bounce and density for atmospheric particles using the DMA-impactor technique.
 15          Atmos. Environ.  10: 1739-1746.
 16
 17      Steiner, D.;  Burtscher, H.; Gross, H.  (1992) Structure and disposition of particles from a spark-ignition engine.
 18          Atmos. Environ.  Part A 26: 997-1003.
 19
 20      Stelson, A. W.; Seinfeld, J. H. (1981) Chemical mass accounting of urban aerosol. Environ. Sci. Technol.
 21          15: 671-679.
 22
 23      Stelson, A. W.; Seinfeld, J. H. (1982a) Relative humidity and temperature dependence of the ammonium nitrate
 24          dissociation constant. Atmos.  Environ.  16: 983-992.
 25
 26      Stelson, A. W.; Seinfeld, J. H. (1982b) Relative humidity and pH dependence of the vapor pressure  of
 27          ammonium nitrate-nitric acid solutions at 25 °C. Atmos. Environ. 16: 993-1000.
 28
 29      Stern, J. E.; Flagan, R. C.; Grosjean, D.; Seinfeld, J.  H. (1987) Aerosol  formation and growth in atmospheric
 30          aromatic hydrocarbon photooxidation. Environ. Sci. Technol. 21: 1224-1231.
 31
 32      Stevens, R. K.; Pace, T. G. (1984) Overview of the mathematical and empirical receptor models  workshop
 33          (Quail Roost II).  Atmos. Environ. 18: 1499-1506.
 34
 35      Stevens, R. K.; Dzubay, T. G. ; Russwurm, G.;  Rickel,  D. (1978) Sampling and analysis of atmospheric
 36          sulfates and related species. Atmos. Environ. 12: 55-68.
 37
 38      Stevens, R. K.; Dzubay,  T. G.; Shaw, R.  W., Jr.; McClenny, W. A.; Lewis, C. W.; Wilson,  W. E. (1980)
 39          Characterization of the aerosol in the Great Smoky Mountains. Environ. Sci.  Technol. 14:  1491-1498.
40
41      Stevens, R. K.; Dzubay,  T. G.; Lewis, C. W.; Shaw, R. W., Jr.  (1984) Source apportionment methods applied
42          to the  determination of the origin of ambient aerosols that affect visibility in forested areas. Atmos.
43          Environ. 18: 261-272.
44
45      Stewart, D. A.; Liu, M.-K. (1981) Development and application of a  reactive plume model. Atmos.  Environ.
46          15: 2377-2393.
47
48      Strapp, J. W. et al. (1988)  Winter cloud water and air composition in central Ontario. J. Geophys. Res.
49          93: 3760-3772.
50
51      Stumm, W.; Morgan, J. J.  (1970) Aquatic chemistry. New York,  NY: Wiley-Interscience.
52
53      Suh, H. H.;  Allen, G. A.; Aurian-Blajeni, B.; Koutrakis, P.; Burton,  R. M. (1994) Field method comparison
54          for the characterization of acid aerosols and gases.  Atmos.  Environ. 28: 2981-2989.


         April 1995                                    3-215       DRAFT-DO NOT QUOTE  OR CITE

-------
 1       Sutherland, A. J. (1967) Proposed mechanism for sediment entrainment by turbulent flows. J. Geophys. Res.
 2           72: 6183-6194.
 3
 4       Sverdrup, G. M.; Whitby, K. T. (1980) The effect of changing relative humidity on aerosol size distribution
 5           measurements. In: Hidy, G. M.; Mueller, P. K.; Grosjean, D.; Appel, B. R.; Wesolowski, J. J., eds. The
 6           character and origins of smog aerosols: a digest of results from the California Aerosol Characterization
 7           Experiment  (ACHEX). New York, NY: John Wiley & Sons; pp. 527-558. (Advances in environmental
 8           science and  technology: v. 10).
 9
10       Sviridenkov, M. A.;  Gillette, D. A.; Isakov,  A. A.; Sokolik, I. N.; Smirnov, V. V.; Belan, B. D.;  Pachenko,
11           M. V.; Andronova, A. V.; Kolomiets, S. M.;  Zhukov, V. M.; Zhukovsky, D. A. (1993) Size distributions
12           of dust aerosol measured during the Soviet-American experiment in Tadzhikistan, 1989. Atmos. Environ.
13           Part A 27: 2481-2486.
14
15       Sweet, C.; Vermette, S. (1993) Sources of toxic trace elements in urban air in Illinois. Environ. Sci. Technol.
16           27: 2502-2510.
17
18       Swozdziak, J. W.; Swozdziak, A. B. (1990) Sulfate aerosol production in the Sudely Range, Poland. J. Aerosol
19           Sci.: S369-S372.
20
21       ten Brink, H. M.; Schwartz, S. E.; Daum, P. H. (1987) Efficient scavenging of aerosol sulfate by liquid-water
22
clouds. Atmos. Environ. 21: 2035-2052.
23
24       Tang, I. N. (1980) Deliquescence properties and particle size change of hygroscopic aerosols. In: Willeke, K.,
25           ed. Generation of aerosols and  facilities for exposure experiments.  Ann Arbor,  MI: Ann Arbor Science;
26           pp. 153-167.
27
28       Tang, I. N.; Lee, J. H.  (1987) Accomodation coefficients of ozone and sulfur dioxide: their implications on
29           S02 oxidation in cloud water. In: Johnson, R. W.; Gordon, G. E., eds. The chemistry of acid rain: sources
30           and atmospheric processes. Washington, DC: American Chemical Society; pp.  109-117.
31
32       Tang, I. N.; Munkelwitz, H. R. (1977) Aerosol growth studies—III.  ammonium bisulfate aerosols in a moist
33           atmosphere. J. Aerosol Sci. 8:  321-330.
34
35       Tang, I. N.; Munkelwitz, H. R.; Davis, J. G. (1978) Aerosol growth studies—IV. phase transformation of
36           mixed salt aerosols  in a moist atmosphere. J. Aerosol Sci. 9:  505-511.
37
38       Tang, I. N.; Munkelwitz, H. R. (1993) Composition and temperature dependence of the deliquescence
39           properties of hygroscopic aerosols.  Atmos. Environ. Part A 27: 467-473.
40
41       Tang, I. N.; Munkelwitz, H. R. (1994) Water activities, densities, and refractive indices of aqueous sulfates and
42           sodium nitrate droplets of atmospheric importance. J.  Geophys. Res. [Atmos.] 99: 18801-18808.
43
44       Tang, I. N.; Wong, W.  T.; Munkelwitz, H. R.  (1981) The relative importance of atmospheric sulfates and
45           nitrates in visibility reduction. Atmos.  Environ. 15: 2463-2471.
46
47       Tani, B.; Siegel, S.; Johnson, S. A.; Kumar, R. (1983) X-ray diffraction investigation of atmospheric aerosols
48           in the 0.3-1.0 pan aerodynamic size range. Atmos. Environ. 17:  2277-2283.
49
50       Tao, Y.;  McMurry, P. H. (1989) Vapor pressures  and surface free energies of C14-C18 monocarboxylic  acids
51           and C5 and C6 dicarboxylic acids.  Environ.  Sci. Technol. 23: 1519-1523.
52
53       Taylor, G. I. (1921) Diffusion by continuous movements. Proc. London Math. Soc. 20: 196-212.
54


         April  1995                                    3-216       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Taylor, G. R. (1989) Sulfate production and deposition in midlatitude continental cumulus clouds: II. chemistry
  2          model formulation and sensitivity analysis. J.  Atmos. Sci. 46:  1991-2007.
  3
  4       Taylor, D. D.; Flagan, R. C. (1982) The influence of combustor operation on fine particles from coal
  5          combustion. Aerosol Sci. Technol.  1: 103-117.
  6
  7       Tesche, T. W.; Haney, J. L.; Morris, R.  E. (1987) Performance evaluation of four grid-based dispersion
  8          models in complex terrain. Atmos.  Environ. 21: 233-256.
  9
 10       Tesche, T. W.; Balentine, H.; Fosdick, E. (1990) Ozone modeling protocol for the San Diego air quality study.
 11          San Diego, CA: San Diego County  Air Pollution Control District.
 12
 13       Tesche, T. W.; Roth, P.  M.; Reynolds, S. D.;  Lurmann, F. W. (1992) Scientific assessment of the urban
 14          airshed model (UAM-IV). Washington, DC: American Petroleum Institute;  report no. API-300-92.
 15
 16       Thibodeaux, L. J.; Nadler, K. C.; Valsaraj, K.  T.; Reible, D. D.  (1991) The effect of moisture on volatile
 17          organic chemical gas-to-particle partitioning with atmospheric aerosols—competitive adsorption theory
 18          predictions. Atmos.  Environ. Part A 25: 1649-1656.
 19
 20       Thomas, P.; Vogt, S. (1990) Mesoscale atmospheric dispersion experiments using tracer and tetroons. Atmos.
 21          Environ.  Part A 24: 1271-1284.
 22
 23       TSI. (1993) Model 390045 DISTFIT ® aerosol data fitting software. St. Paul and Golden Valley, MN: TSI, Inc.
 24          and Chimera Software.
 25
 26       Turpin, B. J.; Huntzicker, J. J. (1991) Secondary  formation of organic aerosol  in the Los Angeles basin: a
 27          descriptive analysis of organic and elemental carbon concentrations. Atmos. Environ. Part A 25: 207-215.
 28
 29       Turpin, B. J.; Huntzicker, J. J. (1994) Identification of secondary organic aerosol  episodes  and quantitation of
 30          primary and secondary organic aerosol concentrations during SCAQS. Atmos. Environ.: accepted.
 31
 32       Turpin, B. J.; Cary, R. A.; Huntzicker, J. J. (1990) An in situ,  time-resolved analyzer for  aerosol organic and
 33          elemental carbon. Aerosol Sci. Technol. 12: 161-171.
 34
 35       Turpin, B. J.; Huntzicker, J. J.; Larson, S. M.; Cass, G. R.  (1991) Los Angeles summer midday paniculate
 36          carbon: primary and secondary aerosol. Environ. Sci. Technol. 25: 1788-1793.
 37
 38      Turpin, B. J.; Liu, S.-P.; Podolske, K.  S.; Gomes, M. S. P.; Eisenreich, S. J.;  McMurry,  P. H.  (1993)
 39          Design and evaluation of a novel diffusion separator for measuring gas/particle distributions of semivolatile
 40          organic compounds.  Environ. Sci. Technol.  27: 2441-2449.
 41
 42      Twohy, C. H.; Austin, P. H.; Charlson, R. J. (1989) Chemical consequences of the initial diffusional growth of
 43          cloud droplets: a clean marine case.  Tellus Ser. B 41: 51-60.
 44
 45       Twomey, S. (1959) The nuclei of natural cloud formation. Part II.  The supersaturation in natural clouds and the
 46          variation of cloud droplet concentration. Geofis. Pura Appl. 43: 243-249.
 47
48       Twomey, S.;  Warner, J.  (1967) Comparison of measurements of cloud droplets and cloud nuclei. J. Atmos.
49           Sci. 24: 702-703.
50
51       U.S. Environmental Protection Agency.  (1982) Air quality criteria for paniculate matter and sulfur oxides.
52           Research  Triangle Park, NC:  Office of Health and Environmental Assessment, Environmental Criteria and
53           Assessment Office; EPA  report no. EPA-600/8-82-029aF-cF. 3v. Available from: NTIS, Springfield VA-
54           PB84-156777.


         April  1995                                    3.217        DRAFT-DO NOT  QUOTE OR CITE

-------
 1       U.S.  Environmental Protection Agency. (1986a) Health assessment document for nickel and nickel compounds.
 2          Research Triangle Park, NC: Environmental Criteria and Assessment Office; report no.
 3          EPA/600/8-83/012FF. Available from:  NTIS,  Springfield, VA; PB86-232212.
 4
 5       U.S.  Environmental Protection Agency. (1986b) Air quality criteria for lead.  Research Triangle Park, NC:
 6          Office of Health and Environmental Assessment, Environmental Criteria and Assessment Office; EPA
 7          report no. EPA-600/8-83/028aF-dF. 4v. Available from: NTIS, Springfield, VA; PB87-142378.
 8
 9       U.S.  Environmental Protection Agency. (1987a) PM10 SIP development guideline. Research Triangle Park, NC:
10          U.S.  Environmental Protection Agency, Office of Air Quality Planning and Standards; EPA report no.
11          EPA-450/2-86-001. Available from: NTIS, Springfield, VA; PB87-206488.
12
13       U.S.  Environmental Protection Agency. (1987b) Protocol for reconciling differences among receptor and
14          dispersion models. Research Triangle Park, NC: U.S. Environmental Protection Agency, Office of Air
15          Quality Planning and Standards; EPA report no. EPA-450/4-87-008. Available from: NTIS, Springfield,
16          VA; PB87-206504.
17
18       U.S.  Environmental Protection Agency. (1994) Guidelines for PM-10 sampling and analysis applicable to
19          receptor modeling. Research  Triangle Park, NC: Office  of Air Quality Planning and Standards; report no.
20          EPA/452/R-94/009. Available from: NTIS, Springfield,  VA; PB94-177441.
21
22       U.S.  Environmental Protection Agency. (1995) Air quality criteria for ozone and related photochemical
23          oxidants. Washington, DC: Office of Research and Development; EPA report nos. EPA/600/AP-93/004a-c.
24          3v.
25
26       Van Borm, W.  A.; Adams, F. C.; Maenhaut, W.  (1989) Characterization of individual particles in the Antwerp
27          aerosol. Atmos. Environ. 23:  1139-1151.
28
29       Van Doren, J. M.; Watson, L. R.; Davidovits, P.; Worsnop, D. R.; Zahniser,  M.  S.; Kolb, C.  E.  (1990)
30          Termperature dependence of  the uptake coefficients  of HNO3, HC1, and N2O5 by water droplets. J. Phys.
31          Chem. 94:  3265-3269.
32
33       Van Grieken, R.; Xhoffer, C. (1992) Microanalysis of individual environmental particles. J. Anal.
34          At. Spectrom. 7: 81-88.
35
36       Van Malderen,  H.; Rojas, C.; Van Grieken, R. (1992) Characterization of individual giant aerosol particles
37          above the North Sea. Environ. Sci. Technol. 26: 750-756.
38
39       Van Valin, C. C.; Luria, M.; Ray, J. D.; Boatman, J. F. (1990) Hydrogen peroxide and ozone over
40          northeastern United States in  1987. J. Geophys. Res. 95: 5689-5695.
41
42       Vanderpool, R. W.; Lundgren, D. A.; Marple, V. A.; Rubow, K. L.  (1987) Cocalibration of four
43          large-particle impactors. Aerosol Sci. Technol. 7: 177-185.
44
45       Venkataraman,  C.; Friedlander, S. K. (1994) Size distributions of polycyclic  aromatic hydrocarbons and
46          elemental carbon. 2. Ambient measurements and effects  of atmospheric processes. Environ.  Sci. Technol.
47          28: 563-572.
48
49       Venkataraman,  C.; Hildemann, L. M.  (1994) factors affecting the distribution of primary semivolatile organi
50          compounds of urban aerosol particles. Environ. Sci. Technol.: submitted.
51
52       Venkataraman,  C.; Lyons, J.  M.; Friedlander, S.  K. (1994)  Size distributions of polycyclic aromatic
53          hydrocarbons and elemental carbon. 1.  Sampling, measurement methods, and source characterization.
54          Environ. Sci. Technol.  28: 555-562.


         April 1995                                    3-218       DRAFT-DO  NOT QUOTE OR CITE

-------
  1      Wagman, D. D.; Evans, W. H.; Parker, V. B.; Schumm, R. H.; Halow, I.; Bailey, S. M.; Churney, K. L.;
  2          Nuttall, R. L. (1982) The NBS tables of chemical thermodynamic properties, selected values for inorganic
  3          and Cl and C2 organic substances in SI units. J. Phys. Chem. Ref. Data II(suppl. 2): 1-392.
  4
  5      Walcek, C. J.; Taylor, G. R. (1986) A theoretical method for computing vertical distributions of acidity and
  6          sulfate  production within cumulus clouds. J. Atmos. Sci. 43: 339-355.
  7
  8      Walcek, C. J.; Stockwell, W. R.; Chang, J. S. (1990) Theoretical estimates of the dynamic radiative and
  9          chemical effects of clouds on tropospheric gases. Atmos. Res. 25: 53-69.
 10
 11      Waldman, J. M.; Lioy,  P. J.; Thurston, G. D.; Lippmann, M. (1990) Spatial and temporal patterns in
 12          summertime sulfate aerosol acidity and neutralization within a metropolitan area. Atmos. Environ. Part B
 13          24: 115-126.
 14
 15      Wall, S. M.; John, W.;  Ondo, J. L. (1988) Measurement of aerosol size distributions for nitrate and major
 16          ionic species.  Atmos. Environ. 22:  1649-1656.
 17
 18      Wang, C.; Chang, J.  S. (1993)  A three-dimensional numerical study of cloud dynamics, micrphysics, and
 19          chemistry. 4. Cloud chemistry and precipitation chemistry. J.  Geophys. Res.  98: 16799-16808.
 20
 21      Wang, H.-C.; John, W. (1987)  Comparative bounce properties of particle materials. Aerosol Sci. Technol.
 22          7: 285-299.
 23
 24      Wang, H.-C.; John, W. (1988)  Characteristics of the Berner impactor for sampling inorganic ions. Aerosol Sci.
 25          Technol. 8: 157-172.
 26
 27      Wang, S.-C.;  Paulson, S. E.; Grosjean, D.; Flagan, R. C.; Seinfeld, J. H. (1992a) Aerosol formation and
 28          growth  in atmospheric organic/NOx systems-I. Outdoor smog  chamber studies of C7- and C8- hydrocarbons.
 29          Atmos.  Environ.  Part A 26: 403-420.
 30
 31      Wang, S.-C.;  Flagan, R. C.; Seinfeld, J. H. (1992b) Aerosol formation and growth in atmospheric organic/NOx
 32          systems—II. aerosol dynamics. Atmos. Environ. Part A  26: 421-434.
 33
 34      Warneck, P. (1991) Chemical reactions  in clouds. Fresenius' J. Anal. Chem. 340: 585-590.
 35
 36      Warneck, P. (1992) Chemistry and photochemistry in atmospheric water drops. Ber. Bunsen Ges. Phys. Chem.
 37          96: 454-460.
 38
 39      Warner, T.  T. (1981) Verification of a three-dimensional transport model using tetroon data from Project
 40          STATE. Atmos. Environ. 15: 2219-2222.
 41
 42      Warner, J.;  Twomey, S. (1967) The production of cloud nuclei by can fires and the effect on cloud droplet
 43          concentration.  J. Atmos. Sci. 24: 704-706.
 44
 45      Warren, R.  S.; Birch, P. (1987) Heavy metal levels  in atmospheric particulates, roadside dust and soil along  a
46          major urban highway. Sci. Total Environ. 59: 253-256.
47
48      Watson, J. G.; Chow, J. C. (1992) Data bases for PM10 and PM2 5 chemical compositions and source profiles.
49          Transactions AWMA 1: 61-91.
50
 51       Watson, J. G.; Henry, R. C.; Cooper, J. A.; Macias, E. S. (1981) The state of the art receptor models relating
52          ambient suspended paniculate matter to sources.  In: Macias, E. S.; Hopke, P. K., eds. Atmospheric
53           aerosol: source/air quality relationships. Washington, DC:  American Chemical Society. (ACS symposium
54          series no.  167).


         April  1995                                    3-219       DRAFT-DO NOT  QUOTE OR CITE

-------
  1      Watson, J. G.; Cooper, J. A.; Huntzicker, J. J. (1984) The effective variance weighting for least squares
  2          calculations applied to the mass balance receptor model. Atmos. Environ. 18: 1347-1355.
  3
  4      Watson, J. G.; Chow, J. C.; Mathai, C. V. (1989) Receptor models in air resources management: a summary
  5          of the APCA international specialty conference. JAPCA 39: 419-426.
  6
  7      Watson, J. G.; Chow, J. C.; Pace, T. G. (1991) Chemical mass balance. In: Hopke, P. K., ed. Data handling
  8          in science and technology: v. 7, receptor modeling for air quality management.  New York, NY: Elsevier
  9          Press; pp. 83-116.
 10
 11      Watson, J. G.; Chow, J. C.; Blumenthal, D. L.; Lurmann, F. W.; Hackney, R. J.; Magliano, K. A.;
 12          Pederson, J. R.; Neff, W.; Roth, P. M.; Solomon, P. A.; Thuillier, R. H.; Ziman, S. D. (1994a) Planning
 13          for data analysis. In: Solomon, P. A.; Silver, T., eds. Planning and managing air quality modeling and
 14          measurement studies: a perspective through SJVAQS/AUSPEX. Pittsburgh, PA: Air and Waste
 15          Management Association; pp. 335-349.
 16
 17      Watson, J. G.; Chow, J. C.; Lurmann, F. W.; Musarra,  S. P.  (1994b) Ammonium nitrate, nitric acid, and
 18          ammonia equilibrium in wintertime Phoenix, Arizona. Air Waste 44: 405-412.
 19
 20      Watson, J. G.; Chow, J. C.; Lu, Z.; Fujita, E. M.; Lowenthal, D. H.; Lawson, D. R.; Ashbaugh, L. L.
 21          (1994c) Chemical mass balance source apportionment of PM10 during the Southern California Air Quality
 22          Study. Aerosol Sci. Technol. 21: 1-36.
 23
 24      Weathers, K. C.; Likens, G. E.; Bormann, F. H.; Bicknell, S.  H.; Bormann, B. T.; Daube, B. C., Jr.; Eaton,
 25          J. S.; Galloway, J. N.; Keene, W.  C.; Kimball,  K. D.; McDowell, W. H.; Siccama, T. G.; Smiley, D.;
 26          Tarrant, R. A. (1988) Cloudwater chemistry from ten sites in North America. Environ. Sci. Technol.
 27          22: 1018-1026.
 28
 29      Went, F. W. (1960) Blue hazes in the atmosphere. Nature (London) 187: 641-643.
 30
 31      Wesely, M. L.; Hicks, B. B. (1977) Some factors that affect the deposition rates of sulfur dioxide and similar
 32          gases on vegetation. J. Air Pollut. Control Assoc. 27: 1110-1116.
 33
 34      Wexler, A. S.; Lurmann, F.  W.; Seinfeld, J. H. (1994) Modelling urban and regional aerosols—I. model
 35          development. Atmos. Environ. 28:  531-546.
 36
 37      Wexler, A. S.; Seinfeld, J. H. (1990) The distribution of ammonium salts among a  size and composition
 38          dispersed aerosol.  Atmos. Environ. Part A 24: 1231-1246.
 39
 40      Wexler, A. S.; Seinfeld, J. H. (1991) Second-generation inorganic aerosol model. Atmos. Environ. Part A
 41          25: 2731-2748.
 42
 43      Wexler, A. S.; Seinfeld, J. H. (1992) Analysis of aerosol ammonium nitrate: departures from equilibrium
 44          during SCAQS. Atmos. Environ. Part A 26: 579-591.
 45
 46      Whitby, K. T. (1978) The physical characteristics of sulfur aerosols.  Atmos. Environ. 12: 135-159.
 47
 48      Whitby, K. T. (1980) Aerosol formation in urban plumes. Ann. N. Y. Acad. Sci. 338: 258-275.
 49
 50      Whitby, K. T. (1984) Physical and optical behavior of sulfate-nitrate aerosols in equilibrium with atmospheric
 51          water vapor, ammonia and nitric acid. In: Ruhnke, L. H.; Deepak, A., eds. Hygroscopic aerosols:
 52          technical proceedings of the workshop on hygroscopic aerosols in the planetary  boundary layer; April  1982;
-53-   -     Vail,  CO. Hampton, VA: A. Deepak Publishing; pp. 45-63.
 54


         April 1995                                    3-220       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Whitby, K. T.; Sverdrup, G. M. (1980) California aerosols: their physical and chemical characteristics.
  2          In: Hidy, G. M.; Mueller, P. K.; Grosjean, D.; Appel, B. R.;  Wesolowski, J. J., eds. The character and
  3          origins of smog aerosols: a digest of results from the California Aerosol Characterization Experiment
  4          (ACHEX). New York, NY: John Wiley & Sons, Inc.; pp. 477-517. (Advances in environmental science
  5          and technology: v. 9).
  6
  7      Whitby, K. T.; Husar, R. B.; Liu, B. Y. H.  (1972) The aerosol size distribution of Los Angeles smog. J.
  8          Colloid Interface Sci. 39: 177-204.
  9
 10      Whitby, K. T.; Charlson, R. E.; Wilson, W. E.; Stevens, R. K.  (1974) The size of suspended particle matter in
 11          air. Science (Washington, DC) 183: 1098-1099.
 12
 13      Whitby, E. R.; McMurry, P. H.;  Shankar, U.; Binkowski, F. S.  (1991) Modal aerosol dynamics modeling.
 14          Research Triangle Park, NC:  U.S. Environmental Protection Agency, Atmospheric Research and Exposure
 15          Assessment Laboratory; EPA report no. EPA/600/3-91/020. Available from: NTIS, Springfield, VA;
 16          PB91-161729.
 17
 18      White, W. H.; Macias, E. S. (1989)  Carbonaceous particles and regional haze in the western United States.
 19          Aerosol Sci.  Technol. 10: 111-117.
 20
 21      White, W. H.; Anderson, J. A.; Blumenthal, D. L.; Husar, R. B.;  Gillani, N. V.; Husar, J. D.; Wilson,
 22          W. E., Jr. (1976)  Formation and transport of secondary air pollutants: ozone and aerosols in the St.  Louis
 23          urban plume. Science (Washington, DC)  194: 187-189.
 24
 25      White, W. H.; Seigneur, C.; Heinold, D. W.; Eltgroth, M. W.; Richards, L. W.; Roberts,  P. T.;  Bhardwaja,
 26          P. S.; Conner, W. D.; Wilson, W. E., Jr. (1985) Predicting  the visibility of chimney plumes:  an
 27          intercomparison of four models with observations at a well-controlled power plant. Atmos. Environ.
 28          19: 515-528.
 29
 30      Willeke, K.; Whitby, K.  T.  (1975) Atmosperic aerosols: size distribution interpretation. J. Air Pollut. Control
 31          Assoc. 25: 529-534.
 32
 33      Willetts, B. (1992)  [Personal communication]. Aberdeen, Scotland:  University of Aberdeen.
 34
 35      Williams, D. J.;  Carras,  J. N.;  Milne, J. W.; Heggie, A. C. (1981) The oxidation and long-range transport of
 36          sulphur dioxide in  a remote region. Atmos. Environ.  15: 2255-2262.
 37
 38      Williams, A. L.;  Stensland,  G. J.;  Gatz, D. F.; Barnard, W. R. (1988) Development of a PM10 emission factor
 39          from unpaved roads.  In: Gatz,  D. F.; Stensland, G. J.; Miller,  M. V.; Chu, L. C., eds. Alkaline aerosols:
 40          an initial investigation of their  role in determining precipitation  acidity, preprint no. 88-71.4.
 41
 42      Wilson, W. E. (1978) Sulfates in the  atmosphere: a progress report  on project MISTT. Atmos. Environ.
43          12: 537-547.
44
45      Wilson, W. E., Jr.  (1981) Sulfate formation in point source plumes: a review of recent field studies. Atmos.
46          Environ. 15:  2573-2581.
47
48      Wilson, J. C.; McMurry, P. H. (1981) Studies of aerosol formation in power plant plumes—II. secondary
49          aerosol formation in  the Navajo generating station plume. In: White,  W. H.; Moore, D. J.; Lodge, J. P.,
 50          eds. Plumes and visibility: measurements and model components: proceedings of the symposium; November
 51          1980; Grand  Canyon National Park, AZ. Atmos. Environ. 15: 2329-2339.
 52
 53
         April 1995                                    3-221       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Wilson, W. E.;  Spiller, L. L.; Ellestad, T. G.; Lamothe, P. J.; Dzubay, T. G.; Stevens, R. K.; Macias, E. S.;
  2           Fletcher, R. A.; Husar, J. D.; Husar, R. B.; Whitby, K. T.; Kittelson, D. B.; Cantrell, B. K. (1977)
  3           General Motors sulfate dispersion experiment: summary of EPA measurements. J. Air Pollut. Control
  4           Assoc. 27: 46-51.
  5
  6       Wilson, J. C.; Gupta, A.; Whitby, K. T.; Wilson, W. E. (1988) Measured aerosol light scattering coefficients
  7           compared with values calculated from EAA and optical particle counter measurements: improving the utility
  8           of the comparison. Atmos. Environ. 22: 789-793.
  9
10       Winkler, P. (1973) The growth of atmospheric aerosol particles as a function of relative humidity: II.
11           an improved concept of mixed nuclei. J. Aerosol Sci. 4: 373-387.
12
13       Winkler, P.; Junge, C. (1972) The growth of atmospheric aerosol particles as a function of relative humidity:
14           I. methods and measurements at different locations. J. Rech. Atmos. 6: 617-638.
15
16       Winklmayr, W.; Wang, H.-C.; John, W. (1990) Adaptation of the Twomey algorithm to the inversion of
17           cascade impactor data. Aerosol Sci. Technol. 13: 322-331.
18
19       Witten, T. A.; Sander, L. M. (1983) Diffusion limited aggregation. Phys. Rev. B: Condens. Matter
20           27: 5686-5697.
21
22       Wolff, G. T. (1981) Paniculate elemental carbon in the atmosphere. J. Air Pollut. Control Assoc. 31: 935-938.
23
24       Wolff, G. T. (1984) On the nature of nitrate in coarse continental aerosols. Atmos. Environ. 18: 977-981.
25
26       Wolff, G. T.; Korsog,  P. E. (1985)  Estimates of the contributions of sources to inhalable paniculate
27           concentrations in Detroit. Atmos. Environ. 19: 1399-1409.
28
29       Wolff, G. T.; Lioy, P. J.; Wight, G. D.; Meyers, R. E.; Cederwall, R. T.  (1977) An investigation of
30           long-range transport of ozone across the midwestern and eastern United States. Atmos. Environ.
31           11:797-802.
32
33       Wolff, G. T.; Groblicki, P. J.; Cadle, S. H.; Countess, R. J. (1982) Paniculate carbon at various locations in
34           the United States. In: Wolff, G.  T.; Klimsch, R. L., eds.  Paniculate carbon: atmospheric life cycle. New
35           York, NY: Plenum Press; pp. 297-315.
36
37       Wolff, G. T.; Ruthkosky, M. S.; Stroup, D. P.; Korsog,  P. E. (1991) A characterization of the principal
38           PM-10 species in Claremont  (summer) and Long Beach  (fall) during SCAQS. Atmos. Environ. Part A
39           25: 2173-2186.
40
41       Worsnop, D. R.; Zahniser, M. S.; Kolb, C. E.; Gardner, J. A.;  Watson, L. R.; Van Doren, J. M.; Jayne,
42           J. T.; Davidovits, P. (1989)  Temperature dependence of mass accomodation of SC^ and H2O2 on aqueous
43           surfaces. J. Phys. Chem.  93:  1159-1172.
44
45       Wu, P.-M.; Okada, K. (1994) Nature of coarse nitrate particles in the atmosphere—a single particle approach.
46           Atmos. Environ. 28: 2053-2060.
47
48       Xu, G. B.; Yu, C. P. (1985)  Theoretical lung deposition of hygroscopic NaCl aerosols.  Aerosol Sci.  Technol.
49           4: 455-461.
50
51       Yamartino, R. J.; Scire, J. S.; Hanna, S. R.; Carmichael, G. R.; Chang,  Y. S. (1989) CALGRID: a mesoscale
52           photochemical grid model. Sacramento, CA: California Air Resources Board; report no. A049-1.
53
         April 1995                                    3-222       DRAFT-DO NOT QUOTE OR CITE

-------
  1       Yamasaki, H.; Kuwata, K.; Miyamoto, H. (1982) Effects of ambient temperature on aspects of airborne
  2           polycyclic aromatic hydrocarbons. Environ. Sci. Technol.  16: 189-194.
  3
  4       Zafiriou, O. C.; Gagosian, R. B.; Peltzer, E. T.; Alford, J. B.; Loder, T.  (1985) Air to sea fluxes of lipids at
  5           Enewetak Atoll. J. Geophys. Res.: 2409.
  6
  7       Zeldin, M. D.;  Bregman, L. D.; Wylie, B. (1988) Meteorological conditions during the 1985 Southern
  8           California Nitrogen Special Methods Comparison Study. Atmos. Environ. 22: 1541-1543.
  9
10       Zemaitis,  J. F.; Clark, D. M.; Rafal, M.;  Scrivner, N. C. (1986) Handbook of aqueous electrolyte
11           thermodynamics. New York, NY: American Institute of Chemical  Engineers.
12
13       Zeng, Y.; Hopke, P. K. (1989) Three-mode factor analysis: anew multivariate method for analyzing spatial and
14           temporal composition variation. In: Watson, J. G., ed.  Transactions: receptor models in air resources
15           management. Pittsburgh, PA: Air and  Waste Management Association; pp.  173-189.
16
17       Zhang, X.; McMurry, P. H. (1991) Theoretical analysis  of evaporative losses of adsorbed or absorbed species
18           during atmospheric aerosol sampling. Environ. Sci. Technol. 25: 456-469.
19
20       Zhang, H. X.; Sorensen, C. M.; Ramer, E.  R.; Olivier,  B. J.; Merklin, J. F. (1988) in situ optical structure
21           factor measurements of an aggregating soot aerosol. Langmuir 4: 867-871.
22
23       Zhang, S.-H.; Shaw, M.; Seinfeld, J. H.; Flagan, R.  C.  (1992) Photochemical aerosol formation from a-pinene
24           and 0-pinene. J. Geophys. Res. [Atmos.] 97: 20,717-20,729.
25
26       Zhang, X.  Q.; McMurry, P. H.; Hering, S.  V.; Casuccio, G. S. (1993) Mixing characteristics and water
27           content of submicron aerosols measured  in Los Angeles and at the  Grand Canyon.  Atmos. Environ. Part A
28           27: 1593-1607.
29
30       Zimmerman, P. R. (1979) Testing of hydrocarbon emissions from vegetation, leaf litter and aquatic surfaces,
31           and development of a methodology for compiling biogenic emission inventories: final report. Research
32           Triangle Park, NC: U.S. Environmental Protection Agency, Office of Air Quality Planning and Standards;
33           EPA report no.  EPA-450/4-79-004. Available from: NTIS, Springfield, VA; PB-296070.
        April 1995                                    3_223       DRAFT-DO NOT QUOTE OR CITE

-------
 i                     4.   SAMPLING AND ANALYSIS OF
 2         PARTICULATE MATTER AND ACID DEPOSITION
 3
 4
 5     4.1   INTRODUCTION
 6          Assessment of the risks associated with airborne aerosols implies that measurements be
 7     made defining the aerosol characteristics, concentrations and exposures that contribute to, or
 8     simply correlate with, adverse health effects.  The proper selection of an aerosol sampling or
 9     analysis methodology to accomplish these measurements requires that rationales be applied
10     that consider how the resulting data will be applied and interpreted, in addition to the data
11     quality required. As an example, integrated collection of an aerosol sample on a heated
12     substrate may help to stabilize a subsequent measurement technique,  but in the process may
13     dramatically change the character of the aerosol as it existed in the air.  Similarly, integrated
14     collection of acidic fine aerosols, without selectively removing the larger, more basic
15     particles, will cause neutralization (i.e., modification) of the sample on the substrate. The
16     same logic applies to the selective removal of gas phase components during sampling that
17     might react with the deposited aerosol sample, in a manner inconsistent with naturally
18     occurring transformation processes.  The assumption that fixed-location measurements are
19     representative of inhalation exposure implies that the effects of local spatial and temporal
20     gradients are understood and appropriately applied to the  sampler siting criteria.
21     Development of relationships between aerosol characteristics and health or ecological
22     responses requires that the aerosol sampling and analysis processes are truly representative
23     and adequately defined.
24          The application of sampling and analytical systems for aerosols must recognize that
25     particles  exist modally as  size  distributions (see Section 3.3.3), generated by distinctively
26     different  source categories and having distinctly different chemistries. The primary reasons
27     for making size-specific aerosol measurements are (a) to relate the in situ aerosol character to
28     the potential deposition sites, and thus toxicity, of the respiratory system, and (b) separation
29     of the size distribution modes to identify sources, transformation process  or aerosol
30     chemistry.  The interpretation  of particle size  must be made based on the diameter definition
31     inherent in the measurement process.  Since the respiratory system classifies particles of
       April 1995                              4-1       DRAFT-DO NOT QUOTE OR CITE

-------
 1     unknown shapes and densities based on aerodynamic diameter, development of aerosol
 2     relationships with health responses requires that sampling techniques either incorporate
 3     inertial aerodynamic sizers or provide mechanisms to accurately convert the measured
 4     diameters (e.g., optical) to an aerodynamic basis.   All particle diameters described in this
 5     chapter are aerodynamic, unless otherwise specified.
 6           Friedlander (1977a) provided the descriptive matrix shown in Figure 4-1 for placing
 7     measurement techniques that define aerosol characteristics into perspective, in terms  of their
 8     particle sizing capabilities, resolution times and chemical identification attributes.  This
 9     approach defined these characteristics  by resolution (single particle or greater), discretizing
10     ability, and averaging  process.  The author notes that the "perfect"  aerosol sampler
11     characterizes particle size with "perfect" resolution, determines the  chemistry  "perfectly" of
12     each particle, and operates in real-time with no "lumping" of classes.  These characteristics
13     could be amended in "real-world" terms by suggesting that the "perfect" sampler would also
14     have minimal cost and operator intervention. Additionally, if the aerosol measurement
15     design goal is mimicking the respiratory system, physiological averaging characteristics  must
16     be considered.  Size-specific, integrated aerosol measurements have improved significantly
17     and their capabilities are better characterized since the  1987 PMi0 standard, but a "perfect"
18     aerosol sampling system has not been devised.  As discussed subsequently, the methodologies
19     required to adequately define the performance  specifications of aerosol samplers have yet to
20     be devised.
21           Many recent developmental  efforts in aerosol measurement technologies  have addressed
22     the need to perfect the chemical characterization of reactive or volatile species collected on
23     filtration substrates. Some of the most significant recent advances in aerosol measurement
24     technologies have come in the form of analysis system "protocols", rather than individual
25     pieces of hardware. Recognizing that there is no single "perfect" sampler, these protocols
26     attempt to merge several aerosol sampling and analysis technologies into an adaptable and
27     analytically versatile system.  System attributes typically include one or more  size-specific
28     aerosol inlets,  subsequent fractionators to separate the  fine and coarse particle modes, and
29     denuders and/or sequential filter packs to selectively account for reactive gas phase species.
30     Examples include EPA's Versatile Air Pollution Sampler (VAPS), (Conner et al.,  1993), the
31     Southern California Air Quality Study (SCAQS) sampler, (Fitz et al., 1989) and the

       April 1995                                4-2        DRAFT-DO NOT QUOTE  OR CITE

-------

Instrument
Perfect Single
Particle Counter
Analyzer
Optical Single
Particle Counter
Electrical
Mobility
Analyzer
Condensation
Nuclei
Counter
Impactor
Impactor
Chemical
Analyzer
Whole Sample
Chemical Analyzer
1 Resolution
Size Time Chemical
Composition
^\ » ^\ • ^\ •
"" """^J * ^W ' ^J
•
-------
 1     during normal activities.  Miniaturization of aerosol separators stretches the limits of current
 2     technologies to maintain required sampling precisions and accuracies.  One of the most
 3     significant limitations imposed by the low flowrates inherent in personal exposure samplers is
 4     the extremely small sample size available for chemical analysis.
 5           This chapter briefly describes the technical capabilities and limitations of aerosol
 6     sampling and analytical procedures in Sections 4.2 and 4.3, respectively, focusing on those
 7     that (1) were used to collect data supporting other sections in this document, (2) those
 8     supporting the existing PM10, TSP and Pb regulations, (3) those that were used to support
 9     health and welfare  response studies, (4) those having application in development of a possible
10     fine particle standard, and (5) discussing the attributes of several new technologies.  The
11     discussion of aerosol separation technologies is divided between (a) devices used to mimic
12     the larger particle (> 10 ^m) penetration  rationales for the upper airways, and (b) those
13     devices generally used to mimic smaller particle penetration (< 10 /*m) to the sub-thoracic
14     regions.  These device descriptions are followed by sampling considerations for their
15     applications.  The applications of performance specifications to define these measurement
16     systems for regulatory purposes are discussed, along with a number of critical observations
17     suggesting that the current specification process does not always assure the accuracy or
18     representativeness necessary in the field.  The EPA program designating PM10 reference and
19     equivalent sampling systems is then briefly  described, along with a current list of designated
20     devices.  Selected measurement systems used to provide more detailed characterization of
21     aerosol properties for research studies are discussed, with a focus on the determination of
22     particle size distributions.  Aerosol sampling systems for specialty applications, including
23     automated samplers, personal exposure samplers and the sampling systems used in aerosol
24     apportionment studies are briefly described.  This chapter is intended to provide supplemental
25     information to Section 3.3.3 and other discussions of aerosol methodologies in support of the
26     existing standards and possible development of a "fine particle" standard in the 1 to 3 /^m
27     range.  An important contribution of the sampling and analytical sections is the extensive
28     compilation of salient peer-reviewed technical references that can be  obtained by the reader
29     for more detailed information.
        April 1995                                4.4        DRAFT-DO NOT QUOTE OR CITE

-------
  1      4.2    SAMPLING FOR PARTICIPATE MATTER
  2      4.2.1    Background
  3           The development of relationships between airborne paniculate matter and human or
  4      ecological effects requires that the aerosol1 measurement process be accurately,  precisely
  5      and representatively defined.  Improvements in sampling methodologies since the 1982 Air
  6      Quality Criteria Document for Particulate Matter and Sulfur Oxides2 was released,  have
  7      resulted from improved sensor technologies, and more importantly, a better understanding of
  8      the aerosol character in situ3.  Additionally, health studies and  atmospheric chemistry
  9      research in the past decade have focused more closely on smaller, better-defined aerosol size
10      fractions of known integrity, collected specifically for subsequent chemical characterization.
11           The system of aerosols in ambient air is a continuum of particle sizes in a gas phase
12      carrier formed as the summation of all size distributions produced by individual sources and
13      secondary transformations.  Portions of the composite distributions are often found to exist
14      lognormally (Baron and Willeke, 1993;  see also Section 3.3.3). Aerosol  systems also exists
15      as a continuum of particle "ages", resulting from loss and transformation mechanisms such
16      as agglomeration, settling, volatilization, gas-particle  reaction, and rain-out affecting freshly
17      generated particles.  The chemical compositions of the various  portions (modes) of the
18      aerosol size distribution are more discreet, and sampling strategies must consider a specific
19      range of sizes for a given chemical class.  The constantly changing character of the
20      atmosphere (or indoor air) places a premium on a sampling strategy to remove a
21      representative aerosol sample from the air and protect its integrity until analyzed.
22           The 1982 Criteria Document provided basic descriptions of many of the aerosol
23      measurement techniques still used today, and these methods will be briefly mentioned here,
24      but not described in detail.  This section will highlight the more recent peer-reviewed
25      Consistent with recent literature (e.g., see Willeke and Baron, 1993), the term "aerosol" will refer to the continuum
26      of suspended particles and the carrier gas.
27      2U.S. Environmental Protection Agency (1982a), referenced subsequently as an entity as the "1982 PM Criteria
28      Document".
29      3The in situ  characteristics of particles in the ambient air medium can be substantially modified by the sampling
30      and analysis processes. For example, a particle counter which draws particles through a restrictive or heated inlet
31      before reaching  the sensing  volume,  may perceive the particle properties (e.g.  scattering coefficients, size
32      distributions) differently from those that existed in the ambient.
        April 1995                                 4.5        DRAFT-DO NOT QUOTE OR CITE

-------
 1     research on aerosol measurement technologies since 1982 and provide salient points that
 2     should be considered in their application.  The aerosol sampling section is not intended to be
 3     an exhaustive treatise, but is structured to highlight important concepts and technologies
 4     relevant to the development of aerosol measurement/response relationships, or supporting
 5     existing  and potential EPA aerosol  regulations.  Ancillary reference texts,  describing basic
 6     aerosol mechanics (e.g., Hinds, 1982; Reist, 1984) and applied aerosol mechanics and
 7     measurements (e.g., Willeke and Baron, 1993; Hering, 1989; Lundgren et al., 1979;
 8     Friedlander,  1977b, Liu,  1976) should be consulted for more fundamental details.
 9
10     4.2.2   Large Particle Separators
11     4.2.2.1   Cutpoint Considerations
12           The collection of an aerosol sample is defined by the penetration characteristics of the
13     inlet, overlaid on the existing in situ size distribution.  Cooper and Guttrich (1981) describe
14     this  process mathematically, and estimate the influences of non-ideal penetration
15     characteristics.  Miller et al. (1979) described the considerations for the possible selection of
16     15 ptm (designated "inhalable") as a standard for size-selective particle sampling with upper
17     airway respiratory deposition as the primary consideration.  The selection  of the most
18     appropriate aerodynamic  criteria for ambient aerosol sampling was only partially resolved by
19     the  1987 EPA designation (U.S. Environmental Protection Agency, 1987b) of a  10 jum
20     cutpoint, PM10 cutpoint.  The  "ideal" PM10 inlet was referenced to the thoracic penetration
21     model of Lippmann and Chan  (1979).  Ogden (1992) noted that the standardization for
22     aerosol cutpoint sizes and separation sharpness is still under debate across settings (ambient
23     air,  occupational) and across national and international governmental entities. As shown in
24     Figure 4-2 (from Jensen and O'Brien, 1993), the international conventions for cutpoints have
25     been roughly categorized as Respirable, Thoracic and Inhalable (previously, Inspirable).
26     These cutpoints are related to the penetration, respectively, to the gas exchange region of the
27     lung, the larynx, and the nasal/oral plane.  The influences of physiological variables on these
28     cutpoints are described by Soderholm (1989).  The British Standard EN 481:1993 describes
29     size fraction definitions for workplace aerosol sampling, and identifies inhalable
30      "conventions" relative to thoracic,  respirable, extra-thoracic and tracheobronchial penetration
31     (but not necessarily deposition) in the respiratory system.  They define a thoracic cumulative

       April 1995                                4-6       DRAFT-DO NOT QUOTE OR CITE

-------
       100
             Inhalable
                  ACGIH(1994)
                  Proposed ISO (1992)
                  ISO (1983)
             Thoracic
                  ACGIH(1994)
                  Proposed ISO(1992)
                  ISO (1983)
  =5  0/100
             Respirable
                  ACGIH(1994)
                  Proposed ISO(1992)
                  ISO (1983)
                  BMRC(1959)
                  Aerodynamic Diameter
Figure 4-2.     American Conference of Governmental Industrial Hygienists
            (ACGIH), British Medical Research Council (BMRC), and
            International Organization for Standardization (ISO) size-selective
            sampling criteria.
April 1995
4-7
DRAFT-DO NOT QUOTE OR CITE

-------
 1     lognormal distribution with a median of 11.64 /*m and a geometric standard deviation of 1.5,
 2     such that 50% of airborne particles with Da = 10 ^m are in the thoracic region.  The
 3     American Conference of Governmental and Industrial Hygienists (ACGIH, 1994) also
 4     adopted these convention definitions as applied to chemical substance TLV's.  Owen et al.
 5     (1992) provides an extensive  list of the outdoor and indoor particles by type and source
 6     category that are  found in or  overlap these ranges.  Willeke et al. (1992) describe the
 7     sampling efficiencies and test procedures for bioaerosol monitors.
 8          The concept of using an inlet that has the same sampling (penetration) characteristics as
 9     portions of the respiratory system has been discussed by a number of researchers, including
10     Lippmann and Chan (1979), Vincent and Mark  (1981), Soderholm (1989), Liden and Kenny
11     (1991) and John (1993).  They describe sampler design considerations for matching
12     penetration models for respirable, thoracic and inhalable fractions that have been proposed by
13     a number of governing bodies.  Since all models proposed for the same fraction do not
14     necessarily coincide, given the variability and differences in interpretation of respiratory
15     system data, Soderholm (1989) proposed compromise conventions for each fraction. Watson
16     et al. (1983), Wedding and Carney (1983) and van der Meulen (1988) mathematically
17     evaluated the influences of inlet design parameters on collection performance relative to
18     proposed sampling criteria.  These analyses suggested that factors such as extremes in wind
19     speed and Coarse particle concentration could pose significant problems in meeting
20     performance specifications.
21          An analysis of the human head as an aerosol sampler was discussed by Ogden and
22     Birkett (1977), who noted that breathing is an anisokinetic  sampling process.   The concept of
23     a "total inhalable" fraction that passes the  oral and nasal entry planes was refined by Mark
24     and Vincent (1986) with the development of a personal aerosol sampling inlet that mimicked
25     this penetration as a function of aerodynamic size.  The inlet was designated the IOM for the
26     Institute for Occupational Medicine in Edinburgh, Scotland, where it was developed with the
27     cutpoint as a function of wind speed  and aerosol type shown in Figure 4-3. The total
28     inhalable approach has been adopted by the International Standards Organization (ISO, 1983
29     and ISO, 1992)4, European Committee for Standardization (CEN, 1993) and by the
30     American Conference on Governmental and Industrial Hygienists (ACGIH, 1985; ACGIH,
31      4 A proposed ISO convention is described by Soderholm (1989)
        April 1995                                4-8       DRAFT-DO NOT QUOTE OR CITE

-------
»**•
2.
VO
<-"











.£.
t


O
T)
H
b
O
o
H
O
s
w
o
n
120


100



^ 80
&
%^
c
 9m/s •'
	 	 V 	 A 0 \ /T
"'*(., N. •
— , *'*'*•* x / ^fc~
O '•--,fc - 4m/s *
"""'«•. N. •'
n Xx \ m/
J(*°\ •! *'*!• ^^t'"'~'~'"'^^B"" ^^A ^™"
£\J *<^f ^H J^ ^B

. ~]JL —
v n /
— V / ^ A —
<37 / *
n im/s x

O
* o

, 1 , , , , 1 , , ,,,,,,





















1 10 100
Aerodynamic Diameter (|im)
	 Inhalable
Convention

Solid Particles

Aloxite

• 1m/s
A 2m/s
• 4m/s
4 6m/s
T 9m/s
Sodium Fluorescein

O 1m/s
A 2m/s
n 4m/s
O 6m/s
V 9m/s
Droplets

DEHS
!_
-------
 1     1994) for workplace aerosol sampling. The total inhalable fraction using the IOM inlet was
 2     selected for a total human exposure study (Pellizzari et al.,  1994) to provide the total body
 3     burden for metals (lead and arsenic), by the air exposure route.
 4          Similar thoracic penetration conventions have been adopted by ISO, CEN, ACGIH and
 5     EPA, each with D50 values of 10.0 Mm (ISO, 1992, CEN, 1993, ACGIH,  1994, and EPA,
 6     1987).  The EPA definition was based primarily on the data of Chan and Lippmann (1980).
 7     The exact shapes of each efficiency curve were mathematically defined by  Soderholm (1989)
 8     and are slightly different for each convention.
 9          The respirable conventions have had D50 values ranging from 3.5 to 5.0 /mi, but a
10     compromise convention has been accepted internationally be several organizations.  It has a
11     D50 of 4.0 /mi (Soderholm, 1989).  ISO (1992) calls this the "healthy adult respirable
12     convention".  Liden and Kenny (1992) discuss the  performance  of currently available
13     respirable samplers. EPA's emphasis on the 2.5 /xm cutpoint was more  closely associated
14     with separating the Fine and Coarse atmospheric aerosol modes, rather than mimicking a
15     respiratory deposition convention.  The exact location of this minimum in the atmospheric
16     size distribution is currently under debate.  It is noteworthy that ISO (1992) defines  a "high
17     risk" respirable convention which is claimed to relate  to the deposition of particles in the
18     lungs of children and adults with certain lung diseases.  The respirable "high risk"
19     convention has a D50 of 2.4 pm, so it could be identified closely with the EPA samplers
20     having a cutpoint of 2.5 /mi.
21          The PM10 size fraction has become nearly universal for ambient air sampling  in the
22     U.S., with the implementation of the  1987  standard (U.S. Environmental Protection Agency,
23     1987a).  The setting of performance specifications, even with their limitations, has provided
24     a more consistent PM10 data base, with better definition of  the data quality.  As additional
25     information becomes available on the sources of biases in aerosol collection methodologies,
26     further characterizations  of older methods may  be  needed to better  define the quality of
27     collected data. Factors that affect bias, and especially representativeness, should be
28     identified and their influences determined as a function of particle size.  As an example,
29     volatilization losses of nitrates were reported  by Zhang and McMurry (1992), while losses
30     for organics were  reported by Eatough et al.  (1993).  Because of the prevalence of these
31     chemical classes in the Fine fraction,  the effect of  the losses on larger fractions (e.g., PM10,

       April 1995                                4-10       DRAFT-DO NOT QUOTE OR CITE

-------
  1     TSP) would be proportionately smaller and can now be estimated.  The losses of larger
  2     particles through aerosol inlet sampling lines (Anand et al., 1992) has a substantial influence
  3     on Coarse fraction samples.  This was demonstrated for the British smoke shade sampler
  4     inlet line by McFarland et al.  (1982).  Inlet losses would be expected to play only a minor
  5     role  in sampling the Fine particle fraction.  Biases in concentration for samplers with large
  6     particle cutpoints are exacerbated by the large  amount of mass present near the cutpoints and
  7     the steep slope of mass versus aerodynamic size.  Thus, small changes in cutpoint can give
  8     significant and hard-to-predict mass biases.
  9             :._
 10     4.2.24  Total Suspended Particulates (TSP)
 11          The gable roof inlet and sampling system for the TSP high volume sampler have
 12     remained essentially unchanged since the sampler's identification as a reference ambient
 13     sampling device in 1971 (U.S. Environmental  Protection Agency, 1971),  The sampling
 14     performance (e.g., wind speed and direction sensitivity) was described in detail in the 1982
 15     Criteria Document, and shown by McFarland and Ortiz  (1979) to collect particles with
 16     aerodynamic diameters exceeding 40 /mi.  More importantly,  its particle collection
 17     characteristics were shown  to be significantly wind speed (2 to 24 km/hr) and wind direction
 18     sensitive.  Only minor technical updates have been incorporated in commercially available
 19     units, such as  in the types of available sequence and elapsed timers (mechanical, electronic)
 20     and in the types of flow controllers (mass flow, volumetric).  Cassettes are now available
 21      that protect the fragile glass or quartz fiber filters during handling and transport.  Size
 22     fractionating inlets for smaller size cutpoints (e.g., 2.5, 6.0 and 10.0 ^m) and cascade
 23      impactors have been developed that can be  retro-fitted in place of the gable roof.  Similar to
 24      the Pb strategy of using the TSP high volume sampler to collect a "total" sample, asbestos
 25      sampling utilizes an aerosol inlet that attempts  to collect a "total" sample, by using an open-
 26      faced filter holder with a conductive inlet cowling.  Baron (1993) discusses the potential
 27      anisokinetic problems that can occur with such a simple  inlet, but notes that the small Stokes
28      number for typical asbestos fibers provides  efficiencies close to 100%.
29
30
        April 1995                                4_H       DRAFT-DO NOT QUOTE OR CITE

-------
 1      4.2.2.2   Total Inhalable
 2           The toxicity of contaminants such as lead pose health concerns as total body burdens,
 3      suggesting that penetration of all aerosols inhaled into the nose and mouth must be
 4      considered, rather than just thoracic penetration.  The TSP sampler for atmospheric lead is
 5      considered (U.S. Environmental Protection Agency, 1978) to more closely capture this larger
 6      size fraction than would a PM10 counterpart, but was not specifically designed  to mimic
 7      inhalability.  The ISO "inhalable" draft sampling convention (ISO, 1993) is intended to apply
 8      to such situations, defining collection of all particles passing the oral/nasal  entry planes.  The
 9      total  inhalable cutpoint is currently available only in a personal sampler version.  Mark and
10      Vincent (1986) described the development of an inhalable inlet (designated as the IOM)
11      meeting the ISO (1992), CEN (1993) and ACGIH (1994) conventions for inpirable dust.
12      This  inlet was improved by Upton et al. (1992) and tested by Mark et al. (1992)  and shown
13      to satisfy the ACGIH criteria for wind speeds of 0.5 and  1.0 m/s.
14
15      4.2.2.3   PM10
16           The penetration of ambient aerosols through a  size-fractionating inlet  to the  collection
17      substrate must be characterized over the ranges of operating conditions (typically,
18      meteorology and aerosol types) that may be encountered.  The  range of conditions currently
19      required by EPA PM10 performance specifications were given in U.S. Environmental
20      Protection Agency (1987b). Ranade et al. (1990) and John (1993) described the required
21      testing, which specifies a controlled flow wind tunnel, mono-dispersed, fluorescently-tagged
22      wet and dry aerosols, and an iso-kinetic nozzle aerosol sampling reference  to determine
23      aerodynamic penetration through candidate PM10 inlets.
24           Marple and Rubow (1976) suggested an  alternate approach to fluorescent  tracer
25      chemistry, using a representative poly-dispersed aerosol and monitoring the size distributions
26      of the challenge aerosol  entering and exiting the inlet in a static chamber with an optical
27      particle counter (OPC).  Buettner (1990) showed that this technique is only accurate if the
28      OPC particle responses are  aerodynamically calibrated to account for factors affecting the
29      optical response, including particle shape and  refractive index.  Maynard (1993) used this
30      approach to determine the penetration of a respirable cyclone to poly-disperse glass micro-
31      spheres, using the TSI, Inc. Aerodynamic Particle Sizer.  John and Wall (1983) noted that

        April 1995                               4-12       DRAFT-DO NOT QUOTE OR CITE

-------
  1      inaccurate inlet sizing results may be obtained using poly-disperse AC test dust, as the result
  2      of agglomeration.  Kenny and Liden (1991) used the APS to characterize personal sampler
  3      inlets, and observed that on theoretical grounds, calm air sampling would be expected to
  4      provide unity aspiration efficiencies for particles below about 8 jum.   Tufto and Willeke
  5      (1982) used an OPC to monitor monodispersed aerosols in a wind tunnel setting to determine
  6      the performance of aerosol sampling inlets relative to an iso-kinetic nozzle. Yamada (1983)
  7      proposed using electron microscopy to determine the size distributions of poly dispersed
  8      particles using manual counting techniques before and  after a candidate aerosol separator.
  9      Penetration data  from this technique were found to be  significantly less precise and difficult
10      to interpret compared with data for the same separators using fluorometric methods.
11           The aerosol cutpoint performance of two PM10 samplers that have met the EPA
12      performance specifications are  illustrated (see Figure 4-4) by the data for the Andersen 321A
13      and Wedding IP10 high volume sampler  inlets at 8 km/hr from Ranade et al.  (1990).  The
14      data show that the cutpoint requirements, defined as a  D50  of 10.0 /zm + 0.5 /xm and
15      mimicking a modeled cutpoint  sharpness (ae), were met for each of the tested wind speeds.
                                                &
16      These performance results were verified by repeating the tests in wind tunnels located at two
17      other research facilities. A diagram (U.S. Environmental Protection Agency, 1990)  of the
18      two-stage Sierra-Andersen PM10 high  volume sampler  inlet with a design flowrate of 1.13
19      m3/min is shown in Figure 4-5.  The buffer chamber of this inlet serves to dampen the
20      influences of variable  wind speeds  and directions.  Aerodynamic separation occurs as the
21      particle-laden air stream passes  through two sets of acceleration nozzles, which deposit the
22      particles larger than PM10 on internal  collection surfaces.  The PM10 fraction is typically
23      collected by  a  glass fiber filter.   An oiled impaction shim was incorporated into the first
24      stage fractionator of the 321A to minimize reentrainment of deposited particles during field
25      sampling.  This modified version (Sierra-Andersen 32IB) was designated as an EPA
26      reference method for PM10 in 1987.  A  subsequent single-stage fractionator (Sierra-Andersen
27      1200) was developed5 and designated as  an EPA  reference  method, with a D50 of 9.5
28      and a hinged design to facilitate cleaning and oiling of the oiled impaction shim.
29      5Graseby-Andersen, Inc., Atlanta, GA.
        April 1995                                4_13       DRAFT-DO NOT QUOTE OR CITE

-------
          100
        S- 80
        ^
        ¥
        0)
        I 60
        1
        LLJ
           40
           20
                                                                  • Wedding IP10
                                                                  0 Model 321A
                                              J_L
                                     4   5   6  7  8  910     15    20
                                          Aerodynamic Diameter (jj.m)
      Figure 4-4. Liquid particle sampling effectiveness curves with solid particle points
                  superimposed for the Wedding IP10 (•) and the Andersen Samplers Model
                  321A inlets at 8 km/h.
1
2
3
4
5
6
1
      A diagram of the cyclone-based Wedding6 PM10 high volume sampler inlet (U.S.
Environmental Protection Agency, 1990) with a design flowrate of 1.13 m3/min is shown in
Figure 4-6.  This inlet uses an omni-directional cyclone to accelerate the particle-laden air
stream to deposit particles larger than PM10 on an oiled collection surface.   Two additional
turns are made to alter the flow  into a downward trajectory toward the collection filter.  A
brush is used to clean the deposited aerosol from the absorber surface through an access port.
This inlet was designated as an EPA reference method for PM10 in 1987.
      6Wedding and Associates, Fort Collins, CO.
      April 1995
                                         4-14
DRAFT-DO NOT QUOTE OR CITE

-------
VO
0
o

z
o
H

O
c;
o
H
w

o
5«
O
HH
H
W
Figure 4-5. Andersen sampler.
                                                              Buffer Chamber



                                                          V^ Air Flow

                                                              Acceleration Nozzle



                                                              Impaction Chamber


                                                              Acceleration Nozzle


                                                              Impaction Chamber

                                                              Vent Tubes



                                                              Filter Cassette

                                                              Filter
                                                              Filter Support
                                                              Screen
                                                              Motor Inlet

-------
 1           The aerosol collection performances for 16.67 1pm PM10 i^618 for tne dichotomous
 2      sampler are described by Wedding et al. (1982) and McFarland and Ortiz (1984), and
 3      illustrated by the penetration data in Figure 4-7.  The variability of the performance as a
 4      function of wind speed for the Andersen 321A PM10 inlet is shown in Figure 4-8 from data
 5      by McFarland et al. (1984).  This is a dramatic improvement over the variability shown by
 6      the TSP high volume sampler (McFarland and Ortiz,  1979) for the same  speed range.  An
 7      attempt to simplify the complexity and improve the availability of wind tunnels to test PM10
 8      inlets was addressed by Teague et al. (1992), who describe a compact tunnel 6 m long by
 9      1.2 m high that  is  capable of testing inlets against the EPA PM10 specifications.
10           Watson and Chow (1993) noted that the EPA PM10 performance specifications allowed
11      a tolerance range around the D50 that permitted inlets to be undesirably "fine tuned" to
12      provide a cutpoint on the lower or upper end of the range.  Since a significant amount of
13      mass in the atmospheric aerosol may be associated with particles in the allowable tolerance
14      range, a "reduction" in reported concentrations could be achieved by simply using a lower
15      (e.g., 9.6 pim) cutpoint inlet that is still within the acceptable D50  range.  The biases
16      between acceptable samplers have been apparent in the data from field aerosol comparison
17      studies (e.g., Rodes et al., 1985; Purdue et al., 1986; Thanukos et al.,  1992).  Most of the
18      reported biases between samplers were less than 10%, although differences of up to 30%
19      were reported.   The data suggested that the high volume sampler PM10 inlets based on
20      cyclonic separation (Wedding, 1985) were consistently lower, while those based on low
21      velocity impaction (McFarland et al., 1984) were consistently higher.  Sweitzer (1985)
22      reported results  of a field comparison of these two high volume sampler types at an industrial
23      location, and reported average biases of 15%.  It was noted that this amount of bias was
24      unacceptable for compliance monitoring and more stringent performance requirements should
25      be used.  Rodes  et al. (1985) observed that the PM10 concentration data from the
26      dichotomous sampler (regardless of the inlet design) gave the most predictable results.
27           Wang and  John (1988) were critical of the EPA PM10 performance  specification on
28      allowable particle bounce (U.S. Environmental Protection Agency, 1987b), stating that the
29      criteria can lead to a 30% overestimation of mass under worst-case conditions.  In a related
30      paper, John et al. (1991) reported that although reentrainment by air flow alone of particles
31      deposited in an aerosol inlet is typically negligible, reentrainment caused  from subsequent

        April 1995                               4-16       DRAFT-DO NOT QUOTE OR CITE

-------
                                    Maintenance Access Port
 Housing
Deflector
Spacing
                              Perfect
                              Absorber
                              No-Bounce
                              Surface

                              Middle
                              Tube
                                 Plug
                              Flow
                              Y
                                                    Vanes

                                                    Vane
                                                    Assembly
                                                    Base
                                                     Insect
                                                    Screen
                     Protective
                     Housing

                    Aerodynamic
                    Inlet
                    Pathway
                Aerodynamic Flow
                   Deflector

                Outer Tube
Figure 4-6.  Sampling characteristics of two-stage size-selective inlet for liquid aerosols.
April 1995
4-17     DRAFT-DO NOT QUOTE OR CITE

-------
           100
            80
         0)
         0
         I
         *=  40
         HI  HU
            20
              0
                              4       6      8  10           20
                         Aerodynamic Particle Diameter (jim)
      Figure 4-7.  Penetration of particles for 16.67 1 pm PM10 inlets.
                       40
1     particle "bombardment" can be substantial. John and Wang (1991) suggested that particle
2     loading on oiled deposition surfaces can bias the collection 2.2%/gram deposited, and
3     strongly suggested that a periodic cleaning schedule should be required for PM10 inlets.
      April 1995
4-18
DRAFT-DO NOT QUOTE OR CITE

-------
           100 -
            80
            60
          o
         I 40
            20
                        o    2 km/h
                        A    8 km/h
                        n  24 km/h
                                              I	I
              U2             4       6     8   10           20
                         Aerodynamic Particle Diameter (urn)
      Figure 4-8.  Collection performance variability as a function of wind speed.
1         The EPA PM10 performance specification program should be considered successful
2     (John, 1993) in providing consistent aerosol collection results during field sampling.  As
3     noted by Thanukos et al. (1992), the cases of greatest concern were those where the
4     measured concentrations were near an exceedance level.  A review of the current PM10
      April 1995
4-19
DRAFT-DO NOT QUOTE OR CITE

-------
 1     performance requirements and possible amendments of the existing specifications may be
 2     appropriate, given the information base now available.
 3
 4     4.2.3   Fine Particle Separators
 5     4.2.3.1   Cutpoint Considerations
 6           Although a particle separation at 2.5 /im has been utilized by the dichotomous sampler
 7     for a number of years, the 1987 standard reassessment (U.S. Environmental Protection
 8     Agency, 1987a) did not specifically require routine monitoring for Fine particles.  It has
 9     become apparent (see Chapter 14) that certain health and ecological responses are most
10     strongly correlated with fine particles, significantly smaller than 10 /*m, and their related
11     chemistry.  Since the mass of a particle is proportional to the cube of its diameter, larger
12     particles (especially above  10 /*m) can totally dominate the mass of PM10 and TSP samples.
13     The 2.5 /im cutpoint generally occurs near a minimum in the mass distribution, minimizing
14     mass concentration differences between samplers with cutpoint biases. The development of
15     control strategies based on mass concentrations from a smaller cutpoint standard must be
16     carefully constructed, especially if large particle interference problems (e.g., particle bounce)
17     cannot be appropriately minimized.  This issue was highlighted by an EPA workshop in
18     May, 1994 that focused on the  implications of introducing a "fine" particle standard and
19     possibly changing the current 2.5 /zm cutpoint for fine  fraction sampling to 1.0 /un. A
20     background paper by Lundgren and Burton (1994) notes that the size fraction less than 1.0
21     /xm typically contains only 0.1% of the total aerosols by mass for particles less than 100 /mi,
22     while the less than 2.5 /mi fraction contains ~ 1%.  By comparison, it was estimated that  at
23     least 50% of the aerosol particles by number are less than 1.0 jum, while at least 80% are
24     less than 2.5 pm.
25           Practical considerations would be the time and expense required to develop inlets with
26     1.0 nm cutpoints that meet required  specifications and  retrofit existing samplers.  Given the
27     body of data available at 2.5 /xm, a focused effort may prove practical that defines the
28     characteristics of the particle mass and chemistry between 1.0 and 2.5 /mi. This would add
29     to the technical knowledge base, allow interpretive corrections between cutpoints to be made,
30     and permit continued sampling  at 2.5 /mi with a minimum of additional resources.
31     Compositional analysis of the PM2 5 to PM10 coarse fraction at eastern U. S. sites show that

       April 1995                                4-20      DRAFT-DO NOT QUOTE OR CITE

-------
 1     more than one half of anthropogenically -produced sulfates would be lost by adoption of a
 2     1.0 /xm cutpoint during the important summer haze period (Eldred et al., 1995).  Similar
 3     losses would also be suggested at some California sites.
 4
 5     4.2.3.2  Virtual Impactors
 6           The dichotomous sampler utilizes virtual impaction to separate the Fine and Coarse
 7     fractions into two separate flowstreams (see Novick and Alvarez, 1987). The calibration of a
 8     nominal 2.5  /mi impactor, including wall loss data,  is shown in Figure 4-9 (from Loo and
 9     Cork, 1988). A virtual impactor has been designed with a 1.0 jum cutpoint (Marple et al.,
10     1989), and for cutpoints  as small as 0.12 /-tm (Sioutas et al.,  1994).  After applying a cross-
11     channel correction factor for the Coarse mode, the mass concentrations of each fraction and
12     total  (using a PM10 inlet) can be determined gravimetrically.  An inherent consideration with
13     virtual separation  is contamination of the Coarse fraction by a portion of the Fine fraction,
14     equivalent  to the ratio of the Coarse channel flow to the total flow (typically 10%). This can
15     influence subsequent  chemical and physical characterizations, if significant differences exist
16     between the  chemistry of each fraction (e.g., acidic Fine fraction and basic Coarse fraction).
17     Stevens et  al. (1993)  utilized this limited addition of Fine  particles to the Coarse fraction to
18     advantage in the SEM analysis of samples  collected on Nuclepore filters.  The current
19     separator design provides a relatively sharp cutpoint with  minimal internal losses.  Keeler
20     et al. (1988) showed  that the growth of Fine aerosols at elevated relative humidities can
21     significantly  alter the ratio of Fine to Coarse collection for the dichotomous sampler.  During
22     early morning periods when the humidity approached 100%, an apparent loss of up to  50%
23     of the Fine mass (to the  Coarse channel) was observed. They commented that analyzing
24     only  the fine fraction of the measured aerosol may not be appropriate,  especially for short
25     integration intervals.
26           A high volume  (1.13 mVmin) virtual impactor assembly was developed by Marple, Liu
27     and Burton (1990) that can be placed on an existing  high  volume sampler to permit larger
28     total  collections than  the dichotomous sampler for chemical speciation by size fraction.  By
29     placing a number of virtual impactors in parallel, a separation can be achieved at higher
30     flows, while reducing the total pressure drop.  Marple et  al.  (1993) provide a list of
        April 1995                                4-21       DRAFT-DO NOT QUOTE OR CITE

-------
                                                      I

                                                       I
                                                       IT)
                                                       I


                                                       I
                                                       U


                                                        5
                                                        s
April 1995
                                4-22
                                        DRAFT-DO NOT QUOTE OR CITE

-------
 1      commercially available virtual impactors by flowrate and available cutpoints.  They also note
 2      that virtual separators  inherently concentrate the particles in the Coarse fraction (typically by
 3      a factor of 10), making them useful as pre-concentrators for sensors with marginal
 4      sensitivities.   John et al. (1983) found that an oiled Nuclepore filter with a nominal 8 /^m
 5      porosity could provide a D50 cutpoint of 2.5 /xm, similar to that of a virtual impactor, if
 6      operated at the appropriate face velocity.
 7
 8      4.2.3.3  Cyclones
 9           Cyclones have been used as aerosol separators in personal exposure sampling in
10      occupational settings for many years. Lippmann and  Chan (1979) summarized the cyclones
11      for sampling aerosol sizes below 10 /xm, and note that the aerosol penetration through a
12      cyclone can be designed to closely  mimic respiratory  deposition.  An intercomparison of
13      three cyclone-based  personal exposure samplers under occupational conditions (concentrations
14      typically  > 1 mg/m3) was described by Groves et al. (1994).  They reported that even
15      though the cyclones  were reportedly design to mimic  similar respirable conventions,  biases as
16      large as a factor of two were noted, possibly attributable to over-loading problems.   Marple
17      et al. (1993) provided a list of commercially available air sampling cyclones, by sampling
18      flowrate and D50 range.  Cyclones  can be used individually or in a cascade arrangement to
19      provide a size distribution.  Hartley and Breuer (1982) describe methods to reduce biases
20      when using a 10 mm (diameter) personal air sampling cyclone, especially as related to
21      cutpoint shifts caused by flowrate changes.  Saltzman (1984) provided a similar analysis for
22      atmospheric sampling  cyclones. Sass-Kortsak et al. (1993) observed that substantial
23      uniformity-of-deposition problems can occur on the filters downstream of personal sampling
24      cyclones.  Wedding  (1983) used a cyclone within a high volume aerosol inlet to provide a
25      PM6 cutpoint for ambient sampling that did not allow penetration of particles greater than
26      10.0 nm.
27           The simplicity  of cyclones has prompted their use as inlets and subsequent separators  in
28      samplers designed to fractionate the aerosol sample for chemical analysis.  The "Enhanced
29      Method"  employed by EPA for sampling acidic aerosols, uses a glass cyclone with a 2.5 /im
30      cutpoint as the sampler inlet (EPA, 1992).  The percent collection as a function of
31      aerodynamic diameter is shown in Figure 4-10 (Winberry et al., 1993).  Bering et al. (1990)

        April 1995                               4-23       DRAFT-DO NOT  QUOTE OR CITE

-------
to
o
o

2
o

o
C
O

s
o
»
o
HH

a
IUU
SO
Cp*
- 60
C
o
ts
0)
o 40
O
20
n







C
/
X
0^^^


: (
/
/
y4
,/ !
k !
) ;



;/(.
s
s






^
X"
)








r— ' '









— i •-







































          1
2      2.5              4


    Aerodynamic Diameter (jam)
S
10
     Figure 4-10.  Percent collection as a function of aerodynamic diameter.

-------
  1     describe several validated aerosol systems for sampling carbonaceous particles that utilize
  2     cyclones with 2.5 jum cutpoints to sample the Fine fraction on either Teflon or quartz
  3     substrates.  Spagnolo and Paoletti (1994) describe a dual cyclone ambient aerosol sampler
  4     with a 15 /mi inlet (described by Liu and Piu,  1981).  This sampler was designed to collect
  5     an "inhalable" 0 to 15 /im fraction,  an "extra-thoracic"7 0 to 4.0 /mi fraction, and a
  6     "respirable" 0 to 2.5 /im fraction.  Malm et al. (1994) describe a sampling system with a
  7     PMIO inlet and three parallel channels following a 2.5 /mi cutpoint cyclone that  was used for
  8     the 40 site IMPROVE network.  Over 120,000 fine particle filter substrates of Teflon6,
  9     nylon and quartz were collected for  chemical analysis  over a 6 year period.
 10
 11     4.2.3.4  Impactors
 12          Impactors have been developed for a wide range of cutpoints and flowrates. In cascade
 13     arrangements (see Section 4.2.7.1.1) with a characterized inlet, impactors provide
 14     distributional information over a range of aerodynamic sizes.  Impactors used as  components
 15     of inlets or as in-line fractionators, stop and  retain the aerosol on a surface  (e.g., oil-soaked,
 16     sintered metal or glass) that hopefully provides consistent performance (primarily minimal
 17     bounce) over the entire sampling interval. Recovery and analysis of the deposited particles
 18     in these situations are usually not considerations.  Koutrakis et al (1990) described the design
 19     of an 2.1 /mi cutpoint impactor for a single stage annular denuder system that exhibited
 20     internal losses of less than 3 %.
 21          Marple et al. (1993) noted that the three primary limitations of impactors are particle
 22     bounce, overloading of collection stages and interstage losses.  Particles can bounce from a
 23     stage after impaction if the surface forces are not adequate for their retention.  Wang and
 24     John (1988) described the effects of  surface loading and relative humidity on particle bounce
 25     and growth, and noted that if less than 6% of the impact area was covered by deposited
 26     particles, particle-to-particle collisions (and bounce) could be neglected.  They also showed
 27     that ammonium sulfate aerosol growth with increasing  humidity resulted in a 25% shift in
28     cutpoint as the relative humidity increased to 64%.  Biswas et al. (1987) showed that,
29     especially in  low pressure zones, the relative humidity and temperature can change rapidly
30     within a cascade impactor, significantly altering cutpoints and losses.  Turner and Hering
31      7The term "extra-thoracic" is most often used to refer to those panicles > 10 ^m.
        April 1995                                4_25       DRAFT-DO NOT QUOTE OR CITE

-------
 1      (1987) noted that the stage substrate materials (Mylar*, stainless steel and glass) with the
 2      same grease (Vaseline*) could produce substantially different particle adhesion
 3      characteristics.  Vanderpool et al. (1987) showed (see Figure 4-11) that using glass fiber
 4      filters as impactor surfaces can produce drastically  reduced performance as compared to a
 5      greased substrate.  Markowski (1987) suggested that adding a duplicate (same cutpoint),
 6      serial impactor stage can permit reasonable bounce and re-entrainment corrections to be
 7      made.
 8
 9      4.2.4   Sampling Considerations
10      4.2.4.1   Siting Criteria
11           The selection of an aerosol sampling location is partially guided by siting criteria
12      provided as part of the 1987 PM10 regulation (U.S. Environmental Protection Agency,
13      1987c), which provided limited  guidance for Pb and PM10 samplers. The details behind
14     these guidelines for PM10 are provided by EPA in  a guidance document (U.S.  Environmental
15      Protection Agency, 1987d), which relates the physical and chemical characteristics of
16     aerosols to the spatial scales (regional, urban, neighborhood, middle and micro) required  to
17     define the influences of sources on various populations. Guidance was also provided on the
18     influences of nearby point, line and area sources on sampling location as a general function
19     of particle size.  Only limited information was noted to be available on specific influences of
20     local obstructions and topography (e.g., trees, buildings) on the measured aerosol
21     concentrations.  The primary focus was establishment of the degree that a sampling location
22     was representative of a specific scale.
23           The high purchase cost,  and occasionally physical size, of aerosol samplers have  tended
24     to restrict the number of sampling sites used in air monitoring studies.  In an attempt to
25     address the biases resulting from too few aerosol samplers in a field study, a "saturation"
26     sampler approach has been used, utilizing an inexpensive, miniature and battery-powered
27     PM10 sampler  that can be deployed at a large number of sites.  Phillips et al. (1994) reported
28     the  application of this approach using 15 PM10 saturation  samplers in conjunction with one
29     dichotomous sampler to study the contribution of diesel emissions to the total paniculate
30     levels in Philadelphia.  Although the mean PM10 concentrations of the saturation samplers
        April 1995                               4-26      DRAFT-DO NOT QUOTE OR CITE

-------
             100
                    O Greased Substrate
                    D Glass-Fiber Filter
                                    5            10            20
                                      Aerodynamic Particle Diameter
       Figure 4-11. Performance of glass fiber filters compared to greased substrate.
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
was essentially identical to that of the dichotomous sampler, the saturation data showed site-
to-site mean differences of as much as 30
4.2.4.2  Averaging Time/Sampling Frequency
     The collection frequency for samples to support the EPA PM10 standard has typically
been on an every-6th-day schedule.  A statistically-based concern (Shaw et al., 1984) was
raised that infrequent collection increases the coefficient of variation about the overall mean
concentration value.  They observed that the variability of computed Fine mass concentration
means  increased as the square root of the number of intervals between individual
measurements.  Symanski and Rappaport (1994), using time series analyses, described the
influences of autocorrelation and non-stationary behavior in occupational  settings on the
concentration distributions constructed from infrequent sampling.  They recommended a
random sampling design where a  sufficient number of locations are sampled repeatedly over
April 1995                               4-27      DRAFT-DO NOT QUOTE OR CITE

-------
 1     an adequate period of time to account for the full range of exposure possibilities.  Hornung
 2     and  Reed (1990) described a method of estimating non-detectable (or missing) values to
 3     improve the variance about the estimate of the geometric mean, by assuming the
 4     concentration distribution is log-normal.
 5           Insufficient sample collections can be remedied by more frequent operation of manual
 6     samplers.  The recent PM10 equivalency designations (see section 4.2.5) of two beta gauge
 7     samplers and  the TEOM sampler can provide the necessary information, with hourly, rather
 8     than daily, resolution.  The initial cost of an automated sampler is typically 2-3  times that  of
 9     a manual, single channel PM10 sampler, but can be offset by the savings in operator labor
10     costs. If the  inherent biases described  in section 4.2.3.4 for the beta and TEOM samplers
11     can  be accommodated (and they  are field reliable), these approaches should prove very useful
12     in routine regulatory and research monitoring studies.  The potential also exists  that the
13     integrating nephelometer may be an acceptable exceedance monitor8, using site specific
14     calibrations relating the measured scattering coefficient, bsp, to Fine aerosol mass
15     concentrations (e.g., Larson et al., 1992).
16           Another consideration for setting  the sampling interval concerns  the setting of start and
17     stop clock times.  Daily 24-h sampling is most often accomplished from midnight-to-
18     midnight, but occasionally from  noon-to-noon to either reduce the number  of samplers
19     required or reduce operator burden.  Sampling locations with highly variable diurnal aerosol
20     concentration patterns (e.g., from night time wood smoke influence or day time traffic dust),
21     or substantial differences between week days and weekend days may require special
22     consideration. These influences  can be especially significant for sampling periods less than
23     24 h.
24
25     4.2.4.3  Collection Substrates
26           The selection of a filtration substrate for integrated collection of particles must be made
27     with some knowledge of the expected particle characteristics  and  a pre-determined analytical
28     protocol.  The expected sampled size distribution places a requirement on the porosity of the
29     filter media to effectively trap a  reasonably high percentage of the particles with a minimum
30      8A Pollutant Standard Index (PSI) monitor used to estimate when a pre-determined exceedance level has been
31      reached or exceeded, to potentially trigger the operation of an equivalent PM10 gravimetrically-based sampler.
        April 1995                                4-28       DRAFT-DO NOT QUOTE OR CITE

-------
 1      of pressure drop.  The most common filter types used in air sampling are fiber and
 2      membrane.  Fiber filters tend to be less expensive than membrane filters, have low pressure
 3      drops, and have high efficiencies for all particle sizes.  They are most commonly available  in
 4      glass fiber, Teflon coated glass fiber and quartz materials.  Membrane filters retain the
 5      particles on the surface for non-depth analyses (e.g., X-Ray Fluorescence), can have specific
 6      porosity's, and are available in a wide  variety of materials.  Teflon is a popular membrane
 7      material because of its inertness, but is 2 to 4 times as expensive as more common materials.
 8      Liu et al.  (1978) summarize the effective penetration characteristics as a function of particle
 9      size and the pressure drops for a wide  variety of fiber and membrane filters.  Polycarbonate
10      filters with well defined porosity's (e.g., Nuclepore0) have been used in "stacked"
11      arrangements as fine particle separators.  John et al.  (1983) describe using an 8 ^m porosity
12      filter in series with a back-up filter to effectively provide a 3.5 /xm separation of fine and
13      coarse particles in a small,  inexpensive package.  Samplers based  on this principle were
14      widely used in the early 1980's (Cahill et al., 1990) and their performance under field
15      conditions was shown to be equivalent  to later cyclone base PM2 5 samplers  in the
16      IMPROVE network.
17           The  reactivities of filter substrates with  the aerosol have been reported extensively.  A
18      common problem with glass fiber filters used on high volume samplers is the basic pH of the
19      glass material and its effective conversion of acid gases to particulate sulfates (e.g., Pierson
20      et al., 1976).  Appel et al.  (1984) also  reported similar conversions of nitrogen oxides to
21      particulate nitrates on glass fiber filters.  Witz et al. (1990) reported losses of particulate
22      nitrates, chlorides and ammonium (19,  51 and 65 %, respectively) from quartz fiber filters
23      during storage. No significant losses of sulfates were reported from quartz filters.
24      Similarly,  Zhang and McMurry (1992) reported the anomalous loss of fine particle nitrates
25      from Teflon filters and noted that predictive loss theories were insufficiently accurate to
26      permit corrections.  Lipfert (1994) also observed that nitrate artifacts on glass fiber filters
27      were difficult to quantify on a  routine basis.  Measurements of particulate nitrate using nylon
28      filters by the IMPROVE protocols show, however, that such effects are minor except in
29      California  (Malm et al., 1994).  Eatough et al. (1993) found significant losses of particulate
30      organic compounds on quartz filters due to  volatilization, such that the ambient
31      concentrations of particulate carbon may be underestimated substantially.  Lipfert (1994)

        April 1995                               4_29       DRAFT-DO NOT QUOTE OR CITE

-------
 1      investigated filter artifacts in a field study in New York, and concluded that positive sulfate
 2      artifacts inflated PM10 values from glass fiber filters by 6 jwg/m3.  It was noted that the
 3      combination of sulfate and nitrate artifacts on glass fiber filters may inflate TSP
 4      concentrations by as much as 10 to 20 jug/m3.
 5
 6      4.2.4.4  Chemical Speciation Sampling
 7           The collection of aerosol samples for chemical speciation analysis adds another
 8      dimension to the complexity of the sampling protocol (also see Section 4.3 for additional
 9      discussion).  The simplest approach utilizes a characterized inlet or separator to define a size
10      fraction, provides an aerosol collection substrate compatible with the analytical technique,
11      and collects an adequate quantity of sample for analysis.  This approach is applicable for
12      relatively nonreactive and stable components such as heavy metals.  An important
13      consideration  is the potential reactivity of the sampling substrate with either the  collected
14      aerosols or the gas phase. Appel et al. (1984) predicted the effect of filter alkalinity on the
15      conversion of acid  gases to sulfates and nitrates.  They provided an upper limit estimate on
16      artifact sulfate formation  (added mass) for TSP high volume sampling  of 8-15 /xg/m3 for a
17      24-h sample.
18           Analyses for  semi-volatile organics which are found in both  the particle and vapor
19      phases must be collected  by adding a vapor trap (e.g., polyurethane foam plug)  downstream
20      of the sampling filter.  Arey et al. (1987) noted that this arrangement of sequential sampling
21      reservoirs may account for the total mass of organics, but not accurately describe their phase
22      distribution in situ, due to "blow-off" from the filter during sampling.  Van Vaeck et al.
23      (1984) measured the volatilization "blow-off" losses of organic species from cascade
24      impactor sampling  to be up to 30%, while the loss of total mass was only  10%.  McDow and
25      Huntzicker (1990)  characterized the face velocity dependence  for organic carbon sampling
26      and provided  correction models, based on adsorption losses to a backup filter.  Turpin et al.
27      (1994) examined organic  aerosol sampling artifacts and highlighted the distinction between
28      "organic carbon" and individual organic species.  They observed that organic carbon
29      sampled from the atmosphere is unlikely to attain equilibrium between that in the gas phase
30      and that adsorbed on a quartz fiber back-up filter.  They also  noted that under typical
31      sampling conditions, adsorption is the dominant artifact in the sampling of particulate organic

        April 1995                                4-30      DRAFT-DO NOT QUOTE OR CITE

-------
  1      carbon, and longer sampling periods reduce the percentage of collected material that is
  2      adsorbed vapor.  It was recommended that collection of aerosols for carbon analyses by
  3      made on  a pre-fired quartz filter, with estimates of the adsorption artifact made from a quartz
  4      filter placed behind a Teflon filter in a parallel sampler.
  5           For more highly reactive and unstable species, the recognition of the in situ character
  6      of the aerosol  in the air must be identified and preserved during all facets of the sampling
  7      process to provide a representative and accurate sample.  Durham et al.  (1978) described a
  8      denuder to  remove sulfur dioxide while sampling for sub-micron aerosols.  Spicer and
  9      Schumacher (1979) observed  that many artifact reactions may occur if stripping of nitric
10      acid, sulfuric acid and ammonia is not performed during speciated aerosol sampling.  Appel
11      et al. (1988) described  the various  loss mechanisms that apply to the aerosol and vapor
12      phases while sampling for nitric acid. They noted that residence time, surface material
13      compositions,  and conditioning prior to sampling were the predominant variables  affecting
14      transmission efficiency.
15           The determination of strong acidity  for atmospheric aerosols (U.S. Environmental
16      Protection Agency, 1992)  describes an "Enhanced" method that recognizes the inter-
17      relationships between the vapor and aerosol phases for each constituent and the potential
18      interferences.  An inlet cyclone or impactor is used to provide a 2.5 /*m cutpoint  to exclude
19      the higher pH  aerosols  found in the Coarse fraction. As shown in Figure 4-12, denuders are
20      used in the flowstream  which selectively remove gas phase components with minimal,
21      characterized losses of  aerosol.  Ye et al. (1991) determined the aerosol losses through an 10
22      1pm annular denuder system as a function of particle size. They noted that total particle
23      losses were  less than a  few percent whether the denuders were coated or uncoated.  Forrest
24      et al. (1982) using parallel annular denuders,  found aerosol losses of only 0.2-2.2% for 0.3-
25      0.6 jum particles, and 4-5% for 1-2 /mi particles.
26           Filter packs have  been developed, consisting of a sandwich of filters and collection
27      media of various types  in series, to collect  aerosols and selectively trap gases and aerosol
28      volatilization products.   Benner et al. (1991) described an annular denuder sampling system
29      using Teflon and nylon filter packs and annular denuders to quantitatively collect  the
30      distributed ammonium nitrate, nitric acid  and ammonia in the vapor and aerosol species.
31      They observed that volatile nitrates were 71 % ±27% of the total nitrates during the day and

        April 1995                                4.31        DRAFT-DO NOT QUOTE OR CITE

-------
                          Filter Pack
                                 Nylon
                                 Teflon
                H
                 +3
d4
1
d3
i
d2
Coupler (Typical) — — [
d1
£
'tk
ff'>r
OBI
nu
v4
;!,(!
8
i
Hj||
yarn
mm
m
4
'ff,
n^**
X5
m
MS
"V?*
*
-tlli
•O
3
o
*k_
+-I
b
f
NH3
V

m
O

-------
 1           Vossler et al. (1988) reported the results of improvements in an annular denuder
 2      system, including Teflon coating of the internal glass surfaces.  They found an apparent
 3      particle bounce problem with the cyclone inlets (with or without Teflon coating), and
 4      proposed adding an additional in-line,  greased impactor.  John et al. (1988) found that
 5      anodized  aluminum surfaces  absorb nitric acid efficiently and irreversibly.  A number of
 6      method comparison studies have been reported for systems utilizing annular denuder/filter
 7      pack technologies, including  Harrison  and Kitto (1990), Sickles et al. (1990), and Benner
 8      et al. (1991).
 9
10      4.2.4.5  Data Corrections/Analyses
11           Aerosol concentration data are reported in units of mass per volume (e.g., /zg/m3).  The
12      current EPA regulations for  sampling TSP, PM10 and Pb require that sampler  flowrates be
13      controlled and the sampled volumes be standardized  to 760 mm Hg and 25 °C. These
14      requirements may pose problems in the interpretation of concentrations from aerosol
15      samplers.  The flowrate through inertial  impactors should be maintained at "local"
16      temperatures and pressures to retain the separator's aerodynamic calibration.  Mass  flow
17      controllers may significantly  affect the separator flow velocity during large diurnal
18      temperature changes, excessively biasing the resulting cutpoint diameter.
19           Subsequent correction of the sampled aerosol volume to "standard"  conditions by
20      mathematically compensating for average meteorological  conditions may improperly report
21      the aerosol concentration measurement.  If the rationale for aerosol sampling was to mimic
22      respiratory penetration (which occurred at local conditions), a correction after-the-fact may
23      not be appropriate.  These corrections  are typically small (less than a few percent) except in
24      locations  at higher altitudes and those with large diurnal or seasonal temperature changes.
25      The basis for mandating flowrate controller performance  for aerosol samplers is sound, but
26      and  the subsequent requirements for concentration corrections for temperature  and pressure
27      are complex.  Although the issue of sampled volume correction for local  temperature and
28      pressure is beyond the scope of this document, the scientific bases should  be reassessed  for
29      aerosol sampling to determine if this requirement is consistent with EPA goals.
30           The matching of aerosol measurement capabilities with data quality requirements is
31      discussed by Baron and Heitbrink (1993).  They note that although aerosol sampler precision

        April 1995                                4.33      DRAFT-DO NOT QUOTE  OR CITE

-------
 1     can be determined from collocated measurements, field sampling accuracy is more difficult to
 2     define.  Generation of mono- or polydisperse calibration aerosols are rarely  done in field
 3     settings because of the complexity of the calibration process. Typically, only the aerosol
 4     sampler flowrate accuracy is determined in the field.  Biases between the means from
 5     collocated aerosol samplers using different separation techniques, may result from sampler
 6     operational errors, or from inadequacies in determining the performance specifications during
 7     laboratory testing.
 8
 9     4.2.5  Performance Specifications
10     4.2.5.1 Approaches
11           A significant step in the standardization process for aerosol sampling was the EPA
12     definition (U.S. Environmental Protection Agency, 1987a) of the PM10  size fraction, based
13     on the aerodynamic diameter of particles capable of penetrating  to the thoracic  region of the
14     respiratory system.  This definition was followed by implementation of  the PM10 provisions
15     of EPA's Ambient Air Monitoring Reference and Equivalent Methods regulation (U.S.
16     Environmental Protection Agency, 1987b).  The format of the latter regulation (see section
17     4.2.5 for  specifics) was the adoption of performance specifications for aerosol samplers,
18     based on controlled wind  tunnel testing with mono-dispersed aerosols.   Controlled laboratory
19     testing is  followed by limited field testing, including tests of candidate equivalent methods to
20     demonstrate comparability to designated reference methods.   This approach  was chosen,
21     rather than the design  specification approach taken in 1971 (U.S. Environmental Protection
22     Agency, 1971), which identified the high volume sampler and associated operational
23     procedures as the reference method for Total Suspended Particulates (TSP)9. The 1971
24     regulation had no provisions for the  use of alternative or equivalent methods.  Subsequent to
25     this design designation, significant problems of the TSP high volume sampler, such as wind
26     speed and direction dependency (McFarland and Rodes,  1979) and off-mode collection (Sides
27     and Saiger, 1976), were reported. These inherent biases complicated the interpretation of
28     TSP concentration data (U.S. Environmental Protection Agency, 1982c) and weakened
29      Subsequent identifications in this section:  "TSP" for Total Suspended Particulates by high volume sampler,
30      "PM10" for the fraction less than 10 /xm, "Fine" (capitalized) for the fraction less than 2.5 /mi, and "Coarse" for
31      the fraction between 2.5 and 10 /j.m.
        April  1995                               4-34       DRAFT-DO NOT QUOTE OR CITE

-------
  1     correlations with other measures. The problems were estimated to have induced biases of
  2     less than 10% for most situations, but occasionally as high as 30%. The subsequent
  3     development of aerosol testing programs for size selective aerosol samplers (e.g., McFarland
  4     and Ortiz,  1979; Wedding, 1980; John and Wall, 1983; Ranade et al., 1990; Hall et al.,
  5     1992) more rapidly identified weaknesses in existing technologies and facilitated the
  6     development of better methods.  No reference standard exists for aerosol concentration
  7     measurements in air.  The calibration of aerosol samplers relies primarily on
  8     characterizations under controlled conditions of the sampler sub-systems, including the size
  9     selective inlet, sample conditioning and transmission system, the flow control system, and, if
 10     used,  subsequent size separators, sample collection and storage elements, and sensors and
 11     associated electronics.  Although the precision of an aerosol  sampler is readily  obtained by
 12     using  replicate, collocated samplers, the accuracy can only be estimated  by comparison with
 13     either designated "reference" samplers  or with computations of expected aerosol mass
 14     collections.  Performance specification  limits are used to control the overall aerosol sampling
 15     accuracy.  As noted by John (1993) the selection of a comprehensive list of sampling
 16     elements requiring inclusion and the setting of the performance limits for each element is a
 17     difficult task, especially when the range of "real-world" sampling situations is considered.
 18          Performance specifications were utilized for the PM10 standard to allow the broadest
 19     spectrum of measurement technologies, hopefully encouraging the development of new and
 20     better methods.  A research program was implemented by  EPA in parallel with the 1982
 21     Criteria Document to identify the critical specifications and understand the inter-relationships
 22     among the parameters influencing the aerosol sampling process. Studies of the influences of
 23     factors such as wind velocity, particle character, flow rate  stability, particle bounce and wall
 24     losses on precision and accuracy substantially advanced the science of large particle
 25     sampling.  The performance specification approach was a significant improvement over the
 26     design specification approach used for the TSP high volume sampler, in  that it fostered the
27     development of new information and technologies, and provided for the use of alternative
28     methods. In retrospect, the primary weakness  of the design specification approach for the
29     TSP reference method was not the process per se, but the technical inadequacy of the
30     development and testing program that produced the high volume sampler design.
       April 1995                                4.35       DRAFT-DO NOT QUOTE OR CITE

-------
 1           The utilization of a performance specification approach requires that a minimum level
 2     of knowledge be available about the measurement process and the associated test procedures.
 3     Some significant drawbacks subsequently observed in the performance specification approach
 4     for PM10 included the complexity, expense and scarcity of aerosol wind tunnel test facilities,
 5     and the difficulty in defining comprehensive specifications that considered all of the nuances
 6     of aerosol sampling.  Wind tunnel evaluation and limited field tests do not always identify
 7     sampler related problems encountered during extended  periods of ambient sampling (e.g.,
 8     John  and  Wang, 1991).
 9
10     4.2.5.2  Critiques
11           Aerosol sampling research studies since the 1982  Criteria Document have identified a
12     number of factors that influence the precision and accuracy of both wind tunnel sampler
13     performance testing and  individual aerosol samplers, demonstrated to meet performance tests.
14     Rodes et al.  (1985) and Purdue et al.  (1986) showed in field evaluations that PM10 samplers
15     meeting the EPA performance specifications did provide consistent aerosol concentration
16     measurements within 10%, under a variety of sampling situations,  but reported that
17     significant biases were evident.  The biases were based on expected collections computed
18     from a knowledge of aerosol inlet penetration by particle size and the size distributions of
19     ambient aerosol mass by particle size.  They also noted that sampler precisions (coefficients
20     of variation) were better than +10%, with several samplers better than ±5%.
21           Mark et al. (1992)  reviewed the attributes of wind tunnel testing,  and noted that tests
22     using controlled conditions are a necessity to determine whether  an aerosol sampler meets a
23     basic cadre  of established performance specifications.  Hollander (1990) suggested  that
24     sampler performance criteria should be evaluated in controlled outdoor tests, given the
25     inability of wind tunnels to  accurately mimic the influences of outdoor meteorological
26     conditions on sampling.  The current EPA PM10 performance  testing requires field tests to
27     demonstrate  sampler precision and flow rate stability, and the comparability  of equivalent
28     methods to designated reference methods. The stringency of such tests are highly dependent
29     on the sampling location chosen, local aerosol sources,  the existing meteorology and the
30     season.
        April 1995                                4-36       DRAFT-DO NOT QUOTE OR CITE

-------
  1           Kenny and Liden (1989) noted that the EPA PM10 sampler performance specifications
  2      (U.S. Environmental Protection Agency, 19875) provided inadequate consideration for
  3      defining the uncertainty in each parameter, and  suggested that bias mapping approaches be
  4      considered.  This approach relates the allowable precision of a parameter to the critical
  5      values of expected bias that just meet the specifications.  Botham et al. (1991) recommended
  6      that the wind tunnel test system duplicate the expected field sampling scenarios as  closely as
  7      possible,  including characteristic flow obstructions.  They described the wind tunnel testing
  8      of personal aerosol samplers mounted on an  anthropogenically consistent (e.g., breathing,
  9      heated) mannequin. Hoffman et al.  (1988) and  John et al. (1991) described the adverse
10      influence of internal surface soiling on aerosol collection performance during extended field
11      operation, and noted that the existing EPA PM10 performance specifications only considered
12      clean samplers.
13           Significant new innovations in  aerosol sensing technologies, that met the PM10
14      performance specification and earned designations as equivalent methods (see section 4.2.6)
15      have occurred since the 1982 Criteria Document.  These indirect10 methods include
16      automated beta attenuation monitors  (e.g., Merrifield,  1989; Wedding and Weigand, 1993),
17      and the automated Tapered Element  Oscillating  Microbalance (TEOM) technology
18      (Patashnick and Rupprecht, 1991).  These designations added automated  sampling  capabilities
19      to the previously all-manual list of sampling  methods.  Recent field tests of both the beta and
20      TEOM methodologies suggest that biases compared to gravimetrically-based samplers may
21      exist that were not identified by the EPA performance test requirements. Arnold et al.
22      (1992) provide data suggesting that the mass concentration data from  a Wedding beta gauge
23      averaged  19% lower than a collocated Wedding  PM10  gravimetric sampler.  Several
24      researchers,  including Hering (1994) and Meyer (1992), have suggested that the TEOM
25      method can exhibit biases (not identified by performance testing) caused by excessive heating
26      (and desiccation) of the sampled aerosol  due to operation at an elevated reference
27      temperature (30 or 50 °C) during the measurement process.  Devising comprehensive
28      performance specifications and test procedures for aerosol samplers, given the complexities
29      of aerosol mechanics,  is a demanding task.
30      10An alternate technology used instead of direct gravimetric analysis to infer mass concentrations from developed
31      relationships.
        April 1995                               4.37       DRAFT-DO NOT QUOTE OR CITE

-------
 1          The size-selective, gravimetrically-based, 24-h manual aerosol concentration
 2     measurement has been the mainstay of compliance sampling for at least two decades.
 3     Although several new sensor technologies have been designated as Equivalent methods for
 4     PM10 by EPA, no superior technology has been developed that is a better reference method
 5     than that based on collection of a discreet aerosol sample followed by gravimetric analysis.
 6     Improvements have been made since 1982 in the accuracy and precision of integrated,
 7     manual aerosol sampling. Some of the most significant advances have occurred in aerosol
 8     size separation technologies, improved performance characterization test methods, and
 9     speciation sampling techniques.
10          As discussed by Lippmann (1993), there  may be no threshold for health responses
11     down to the lowest aerosol concentrations.  This implies that the precision and lower
12     detection limit requirements will continue to be important for aerosol measurements  across
13     the concentration spectrum.   These factors become even more critical as the size fraction of
14     interest becomes smaller and fewer total particles are collected.   At low concentrations
15     (especially with small size fractions), normally insignificant factors can become important
16     contributors to biases.  Witz et al. (1990) reported rapid and substantial  losses of nitrates,
17     chlorides and ammonium ion (19, 65 and 51%, respectively) from quartz high volume
18     sampler filters during storage periods of one week prior to analyses.  Transformations can
19     also occur on glass fiber substrates during sampling, as reported by Sickles and Hodson
20     (1989) for the rapid conversion of collected  nitrites to nitrates in the  presence of ozone.
21     Zhang and McMurry (1992) showed that nearly complete evaporative losses of Fine particle
22     nitrate can occur during sampling on Teflon filters. Lioy et al. (1988),  in a study using
23     PM10 samplers, reported 25-34% lower concentration values resulting from losses of glass
24     fibers from the filter to the  filter holder gasket during sampling. Feeney et al. (1984)
25     reported weight gains in Teflon filters  used in  contaminated ring cassettes, that posed
26     significant problems for light aerosol loadings.  Grinshpun et al. (1993)  suggest that if
27     unavoidable changes in the  aerosol occur during sampling, development of a model that
28     permits back-calculation of  the in situ characteristics is required.
29
30
        April 1995                               4-38       DRAFT-DO NOT QUOTE OR CITE

-------
 1     4.2.6  Reference and Equivalent Method Program
 2          Ambient air PM10 measurements are used (among other purposes) to determine whether
 3     defined geographical areas are in attainment or non-attainment with the national ambient air
 4     quality standards (NAAQS) for PM10. These measurements are obtained by the States in
 5     their state and local air monitoring station (SLAMS) networks as required under 40 CFR Part
 6     58.  Further, Appendix C of Part 58 requires that the ambient air monitoring methods used
 7     in these EPA-required SLAMS networks must be methods that have been designated  by the
 8     EPA as either reference or equivalent methods.
 9          Monitoring methods for paniculate matter (i.e., PM10) are designated by the EPA as
10     reference or equivalent methods under the provisions of 40 CFR Part 53, which was
11     amended in 1987 to add specific  requirements for PM10 methods. Part 53 sets forth
12     functional specifications and other requirements that reference and equivalent methods for
13     each criteria pollutant must meet, along with explicit test  procedures by which candidate
14     methods or samplers are to be tested against those specifications. General requirements  and
15     provisions for reference and equivalent methods  are also given in Part 53, as are the
16     requirements for submitting an application to the EPA for a reference or equivalent method
17     determination.  The distinction between reference and equivalent methods  is a technical one.
18     On one hand,  it provides for detailed, explicit specification of a selected measurement
19     technology for reference methods.  On the other hand, it  allows alternative (including
20     innovative and potentially improved) methodologies for equivalent methods,  based only on
21     meeting specified requirements for functional performance and for comparability to the
22     reference method. For purposes  of determining  attainment or non-attainment with  the
23     NAAQS, however, the distinction between reference  and  equivalent methods is largely, if not
24     entirely, immaterial.
25          Under the Part 53 requirements,  reference  methods for PM10 must be shown to use the
26     measurement principle and meet the other specifications set forth in 40 CFR 50, Appendix J.
27     They must also include a PM10 sampler that meets the requirements specified in Subpart D of
28     40 CFR 53.  Appendix J specifies a measurement principle based on extracting an  air sample
29     from the atmosphere with a powered  sampler that incorporates inertial separation of the PM10
30     size range particles followed by collection of the PM10 particles on a filter over a 24-h
31     period.  The average PM10 concentration for the sample period is determined by dividing the

       April 1995                               4.39      DRAFT-DO NOT QUOTE OR CITE

-------
 1     net weight gain of the filter over the sample period by the total volume of air sampled.
 2     Other specifications are prescribed in Appendix J for flow rate control and measurement,
 3     flow rate measurement device calibration, filter media characteristics and performance, filter
 4     conditioning before and after sampling, filter weighing, sampler operation, and correction of
 5     sample volume to EPA reference temperature and pressure.  In addition,  sampler
 6     performance requirements in Subpart D of Part 53 include wind tunnel tests for "sampling
 7     effectiveness"  (the efficacy of the PM10 particle size separation capability) at each of three
 8     wind speeds and "50 percent cutpoint" (the accuracy of the primary 10-micron particle size
 9     separation).  Field tests for sampling precision and  flow rate stability are also specified.  In
10     spite of the instrumental nature of the sampler, this method is basically a manual procedure,
11     and all designated reference methods for PM10 are therefore defined as manual methods.
12          Equivalent methods for PM10, alternatively, need not be based on the measurement
13     principle specified in Appendix J nor meet the other Appendix J requirements.  Instead,
14     equivalent methods must  meet the "sampler" performance specifications set forth in Subpart
15     D of Part 53 and demonstrate comparability to a  reference method as  required by Subpart C
16     of Part 53.  The provisions of Subpart C specify  that a candidate equivalent method must
17     produce PM10 measurements that agree with measurements produced by collocated reference
18     method samplers at each  of two field test sites.  For this purpose, agreement means a
19     regression slope of 1 +0.1, a regression intercept of 0  +5 /ig/m3, and a correlation >0.97.
20     These requirements allow virtually any type of PM10 measurement technique, and therefore
21     an  equivalent method for PM10 may be either a manual method or a fully automated
22     instrumental method (i. e., analyzer).
23          As of this writing, the EPA has designated  seven reference methods and  three
24     equivalent methods for PM10, as listed in Table 4-1.  The reference methods include four
25     methods featuring high-volume samplers from two manufacturers, with one using a cyclone-
26     type size separator and the others using an impaction-type separator.  The other reference
27     methods include a low-volume sampler (from a third manufacturer), a low-volume sampler
28     featuring a secondary size separation at 2.5 microns (dichotomous sampler), and a medium-
29     volume, non-commercial sampler.  The three designated equivalent methods are all
30     automated PM10 analyzers and include two operating on the beta-attenuation principle and
31     one based on a tapered element oscillating microbalance (TEOM™).  It should be noted that

       April 1995                                4-40      DRAFT-DO NOT QUOTE OR CITE

-------
                      TABLE 4-1. EPA-DESIGNATED REFERENCE AND EQUIVALENT METHODS FOR PM10
           Method No.
                                Identification
                                                 Description
                                                    Type
                        Date
      RFPS-1087-062
                      Wedding & Associates3 PM
                      Critical Flow High-Volume
                      Sampler.
                        10       High-volume (1.13 m3/min) sampler with cyclone-
                                type PM10 inlet; 203 x 254 cm (8 x 10 in) filter.
                                             Manual reference
                                             method
                   10/06/87
      RFPS-1287-063
      RFPS-1287-064
                      Sierra-Andersen5 or General Metal
                      Works0 Model 1200 PM10 High-
                      Volume Air Sampler System

                      Sierra-Andersenb or General Metal
                      Worksc Model 321-B PM,0 High-
                      Volume Air Sampler System
                                High-volume (1.13 m3/min) sampler with
                                impaction-type PM10 inlet; 203 x 254 cm (8 x 10
                                in) filter.

                                High- volume (1.13 nrVmin) sampler with
                                impaction-type PM10 inlet; 203 x 254 cm (8 x 10
                                in) filter.  (No longer available.)
                                             Manual reference
                                             method
                                             Manual reference
                                             method
                    12/01/87
                    12/01/87
      RFPS-1287-065
                      Sierra-Andersenb or General Metal
                      Works0 Model 321-C PM10 High-
                      Volume Air Sampler System
                                High-volume (1.13 m3/min) sampler with
                                impaction-type PM10 inlet; 203 x 254 cm (8 x 10
                                in) filter.  (No longer available.)
                                             Manual reference
                                             method
                    12/01/87
      RFPS-0389-071
                      Oregon DEQ Medium Volume
                      PM10 Sampler
                                Non-commercial medium-volume (110 L/min)
                                sampler with impaction-type inlet and automatic
                                filter change; two 47-mm diameter filters.
                                             Manual reference
                                             method
                   3/24/89
      RFPS-0789-073
T)
H
6
O
z
s
o
c
s
w
o
h-H
H
W
EQPM-0990-076
EQPM-1090-079
                      Sierra-Andersenb Models SA241 or  Low-volume (16.7 L/min) sampler with impaction-  Manual reference     7/27/89
                      SA241M or General Metal Works0   type PM10 inlet; additional particle size separation   method
Models G241 and G241M PM10
Dichotomous Samplers

Andersen Instruments'5 Model
FH62I-N PM10 Beta Attenuation
Monitor

Rupprecht & Patashnickd TEOM
Series 1400 and Series 1400a PM-
10 Monitors
at 2.5 micron, collected on two 37-mm diameter
filters.
                     \
Low-volume (16.7 L/min) PM10 analyzers using
impaction-type PM10 inlet,  40 mm filter tape, and
beta attenuation analysis.

Low-volume (16.7 L/min) PM10 analyzers using
impaction-type PM10 inlet,  12.7 mm diameter
filter, and tapered element oscillating microbalance
analysis.
Automated
equivalent method
Automated
equivalent method
9/18/90
10/29/90

-------
>   	TABLE 4-1 (cont'd). EPA-DESIGNATED REFERENCE AND EQUIVALENT METHODS FOR PM10
_,   ====1^=:^===:=:^=^============:^=^=^====^=^=:===========:^=^=========^===:^=:^===^====^=^=====^==:^===:^=^=========:^=^=^=:^==^=^=^=^=^=^=
~        Method No.                Identification                           Description                         Type            Date

Jg   EQPM-0391-081       Wedding & Associates'1 PM10 Beta   Low-volume (16.7 L/min) PM10 analyzer using     Automated         3/5/91

^>                        Gauge Automated Particle          cyclone-type PM10 inlet, 32 mm filter tape, and     equivalent method

                          Sampler                        beta attenuation analysis.



     RFPS-0694-098        Rupprecht & Patashnickd Partisol    Low-volume (16.7 L/min) PM10 samplerwith      Manual reference    7/11/94

                          Model 2000 Air Sampler           impaction-type inlet and 47 mm diameter filter.     method
-f.

N)
H
6
o
Z!
S
O
d
s
m
n
HH
H
W

-------
  1      although these latter three automated PM10 analyzers may be capable of providing continuous
  2      or semi-continuous PM10 concentration measurements, only 24-h average PM10
  3      measurements are recognized as approved under their equivalent method designations.
  4
  5      4.2.7  Determination of Size Distribution
  6           The determination of aerosol size distributions can be a powerful research tool when
  7      studying source contributions and transformation processes.  A number of techniques are
  8      available as described by texts such as Willeke and Baron (1993) to make near real-time,
  9      single particle aerosol measurement in addition to cascade impactors.
 10
 11      4.2.7.1  Cascade Impactors
 12           In cascade applications, the aerosol is impacted and trapped onto a series of removable,
 13      coated substrates (e.g., greased foils), including a final total stage collection on a filter for
 14      gravimetric analysis.  Marple et al. (1993) list over 30 single stage and cascade impactors
 15      that are either commercially available or still commonly  used.  The design and calibration of
 16      a miniature eight-stage cascade impactor for personal air sampling in occupational settings is
 17      described by Rubow et al.  (1987), operating at 2.0 1pm.   Evaluations of the most commonly
 18      used cascade impactor systems have been reported by Vaughan (1989) for the Andersen MKl
 19      and MK2 7-stage cascade impactors, Marple et al. (1991) for the 10-stage Micro-Orifice
 20      Uniform Deposit Impactor  (MOUDI), and Wang and John (1988) and Hillamo and
 21      Kauppinen (1991) for the 6-stage Berner, low pressure cascade impactor.  The smallest
 22      particle stages of these impactors can have very small diameter jets and/or very low total
 23      pressures to achieve the sub-micron separations.  The MOUDI  impactor has 2000 holes on
 24      the lowest cutpoint stage. Raabe et al. (1988) describe an 8 stage cascade slit impactor with
 25      slowly rotating impactor  drums instead of flat plates.  This arrangement, in combination with
 26      a PIXIE analyzer, permitted aerodynamic sizing of elemental components, with temporal
27      resolution.   The skill and care required in the operation of cascade impactors suggests that
28      they are research rather than routine samplers.
29           The importance of the aerosol calibration of a cascade impactor is illustrated by
30      Vaughan (1989) in Figure 4-13,  which compares the experimental data with the
31      manufacturer's calibrations  and indicates biases as large as 1.0 /*m.  Marple et al. (1991)
       April 1995                               4.43       DRAFT-DO NOT QUOTE OR CITE

-------
2J

8

O
d


tr)

O


n
t—i
H
W
         100
          80
          60
       W
       0

       '€ 40
       co
          20
             n Liquid Particles


             • Solid Particles
                            10   9
            Stage —=

          Cut-Points
                                  8
    StageNumber


7654
1   Inlet

      0.01                  0.1                   1


                              Aerodynamic Particle Diameter (|im)


Figure 4-13.  Aerosol calibration of a cascade impactor.
                                                                           10
                                                    100

-------
  1      provided a similar type of stage calibration for the MOUDI impactor, and included data on
  2      the internal particle losses (see Figure 4-9).  These loss data showed that an improperly
  3      designed inlet to the impactor, combined with the inertial and interception losses of the larger
  4      particle sizes, can substantially bias the first stage collections.  This was also demonstrated
  5      for the inlet to the Andersen impactor by McFarland et al. (1977).
  6           Cascade impactors can be used to construct distributions of mass and speciated
  7      constituents  as a function of aerodynamic diameter. These distributions can be constructed
  8      graphically or using matrix inversion techniques (e.g., Crump and Seinfeld, 1982,
  9      Wolfenbarger and Seinfeld, 1990).  Marple et al.  (1993) notes that impactor stage
10      calibrations which do not demonstrate sharp cutoffs can cause significant between-stage
11      sizing errors if not accommodated.  John et al. (1990) measured distributions over the 0.08
12      to 16 fj,m range for mass and inorganic ions for several sites in Southern California.  They
13      identified the standard Coarse mode, and two separate and previously unreported modes in
14      the 0.1 to 1.0 ftm range. This range was referred to by Whitby (1978) as a single
15      "accumulation" mode.  John et al. (1990) described a "condensation" mode at 0.2  ±0.1 fj.m
16      containing gas phase reaction products, and a "droplet" mode at 0.7 ±0.2 /zm which grows
17      from the "condensation" mode by the addition of water and sulfates.  Fang et al. (1991)
18      described the effects of flow-inducted relative humidity changes on the sizing of acid aerosols
19      in the MOUDI impactor.  They noted that it may not be possible to measure size
20      distributions of small (less than about 0.2 to 0.5 |um) particles with impactors at relative
21      humidities exceeding 80%.
22
23      4.2.7.2 Single Particle Samplers
24           Aerosol size distribution data are useful in studies of particle transport and
25      transformation processes, source characterization, and particle sizing and collection device
26      performance. In  addition to cascade impactors, a number of real time or near real  time
27      sizing instruments are available and described in texts such as Willeke and Baron(1993).
28      While cascade impactors provide distributions in  terms of aerodynamically sized mass, single
29      particle sampling  devices can produce optically sized distributions as a function of particle
30      number (count), with surface area and volume distributions computed during the data
31      reduction, assuming spherical particles.  Particle density and shape information as a function

        April 1995                               4.45       DRAFT-DO NOT QUOTE OR CITE

-------
 1      of size are required to convert from volume distributions to an estimated mass basis.
 2      Individual particle sizing and counting instruments are generally limited to a particle
 3      detection range of a decade or so, but several devices can overlap to cover the range of
 4      approximately 0.001 to  10 pm. The principle of detection of an instrument restricts the
 5      particle sizes  which can be detected.  For example, instruments using electrical mobility
 6      analysis are limited to particle sizes less than about 1 jum.  Optical methods are typically
 7      used to measure particles larger than about 0.1 to 0.3 /mi.  Inlet and transport system losses
 8      of coarse particle above about 2 /mi, prior to the sensing volume, must be factored into
 9      reported size  distributions.
10           The three most commonly used single particle sampler types are aerodynamic particle
11      sizers,  electrical mobility analyzers and optical particle counters (OPC's). Aerodynamic
12      particle sizers use laser doppler anemometry to measure the velocity of particles in a jet.
13      The acceleration of the particle is related to the aerodynamic particle diameter.  This
14      technique is typically applied to particles  larger than about 0.5 /mi.  In electrical mobility
15      analysis, aerosol with a known charge distribution flows through an electric field.   The
16      particles migrate according to their mobility which can be related to size.  The original TSI
17      electrical aerosol analyzer (EAA) performed this separation in an integrated manner over the
18      total size distribution and detected the particles by unipolar diffusion charging.  A more
19      versatile approach is the differential mobility analyzer, or DMA (Knutson and Whitby,  1975;
20      Liu et al., 1978) is able examine a narrow slice of the size distribution in an equilibrium
21      charge state,  detected by a condensation nucleus counter (CNC).  Differential mobility
22      analyzers have been employed in pairs (Tandem Differential Mobility Analyzer, or TDMA)
23      to examine particle characteristics such as NH3 and H2SO4 reaction rates (McMurry et al.,
24      1983), and the sensitivity of the size distributions of Los Angeles aerosol to relative humidity
25      (McMurry and Stolzenburg,  1989).  The  latter research used the first DMA to select particles
26      of known mobility from the  input aerosol, a humidification system to condition the selected
27      particles, and the second DMA to determine mobility changes.  Optical particle counters pass
28      a jet of aerosol through an optical system.  Light scattered from individual particles is
29      detected and  the signal in processed in a  multi-channel analyzer.  Discreet signals  are
30      counted and sorted by intensity by optical size.  An example forward-scattering counter with
31      an open sensing volume (for use on aircraft), is the Particle Measuring  Systems, Inc.,  FSSP-

        April 1995                                4-46       DRAFT-DO NOT QUOTE OR CITE

-------
 1      300, which can provide high resolution (31 channel) count distributions over the size range
 2      of 0.3 to 20 urn (Rader and O'Hern, 1993).  Gebhart (1993) described currently available
 3      OPC's and their counting efficiencies over a range of diameters.
 4           Single particle samplers have common considerations.
 5      Calibration: They are calibrated with reference aerosol either by the manufacture or by the
 6      user.  If the properties of the aerosol measured are quite different than the calibration, the
 7      indicated size distribution may be quite different than actual distribution.  Brockman et al.
 8      (1988) demonstrated that the APS calibration can vary significantly with the type of test
 9      aerosol, and showed substantial response biases between oleic acid and polystyrene latex
10      spheres above 10 /mi.  Wang and John (1989) described a procedure to correct the APS
11      response for aerosol particle density.  Particle shape can also provide serious sizing errors,
12      and specific calibrations are needed for particles with shape factors significantly different
13      from unity (spherical).  Yeh (1993) commented that the calculated geometric standard
14      deviations (#„) determined by the EAA and DMA are generally larger than 1.3, even if the
                    &
15      correct value is significantly closer to unity.  Woskie et al. (1993) observed, as did Willeke
16      and Degarmo (1988), that optical particle counting devices must be appropriately calibrated
17      using  realistic aerosols, especially for low concentration applications. Harrison and Harrison
18      (1982) suggested that the ratio of fine particle mass concentration to optical scattering
19      extinction will be more variable when a significant contribution is made by irregular (shaped)
20      particles - an event likely to occur when the mean mass diameter exceeds 1 pirn.
21
22      Particle Concentration Effects:  Gebhart  (1993) noted that the response of single particle
23      counters may be influenced by extremely high particle concentrations. Wake (1989) and
24      Heitbrink et al. (1991) described the coincidence problems of the  APS when sampling high
25      total particle concentrations, especially for sizes greater than 1 /an.  Baron et al.  (1993)
26      reported that the concentration levels giving 1 % coincidence in an aerodynamic particle sizer
27      for 0.8, 3 and 10 jum particles, are the relatively low values of 558, 387 and 234
28      particles/cm3,  resectively.   Optical particle counters  experience coincidence errors (two
29      particles are detected as a single particle) and counter saturation at high particle
30      concentrations.  Hinds and Kraske (1986) described  the performance of the PMS, Inc. LAS-
31      X and noted a sizing accuracy of ±2 channel widths, and coincidence errors of less than

        April  1995                                4.47       DRAFT-DO NOT QUOTE OR CITE

-------
 1      10% for concentrations below 10,000 particles/cm3.  Clearly, typical particle concentrations
 2      found in the atmosphere may produce significant errors if sample dilution is not utilized.
 3
 4      4.2.8  Automated Sampling
 5          Automated methods to provide measures of aerosol  concentrations in the air have
 6      existed  for decades in an attempt to provide temporal definition of suspended particles with a
 7      minimum labor expense.  Some  of the automated samplers described in the 1982 Criteria
 8      Document (e.g., British Smoke Shade and AISI tape samplers), were indicator measures of
 9      aerosol  concentration, using calibrations relating aerosol concentrations to reflected or
10      absorbed light. Tape samplers were used in the U. S. primarily as exceedance (index)
11      monitors.
12          The beta attenuation and integrating nephelometer techniques were described in 1982
13      primarily as research methods.  Refinements to the beta gauge sampling approach and the
14      addition of the Tapered Element Oscillating Microbalance (TEOM) principle have resulted in
15      their designation as equivalent methods for PM10.
16          Presently there are no commercially available, automated high volume (>  1 m3/min
17      flowrate) aerosol  samplers, excluding the possibility of the timed operation of an array of
18      manual  samplers.  The physical  size of such a sampling system using 8 x 10 inch filters is
19      impractical.  The dichotomous sampler is currently the only low volume,  gravimetrically-
20      based sampler, commercially available in an automated version.
21
22      4.2.8.1  TEOM
23          The Tapered Element Oscillating Microbalance (R & P, Inc.) sensor was described by
24      Patashnick and Rupprecht (1990), and consists of an oscillating tapered tube with a filter on
25      its free  end (see the diagram in Figure 4-14).  The change in mass  of the filter and collected
26      aerosol  produces  a shift in  the oscillation frequency of the tapered tube that can be directly
27      related  to mass.   Rupprecht et al. (1992) suggested that the filter can be archived after
28      sampling for subsequent analysis.  The sampler inlet has  a PM10 cutpoint and operates at
29      16.67 1pm. A flow splitter samples a 3  1pm portion of this flow to be filtered.  Since the
30      fraction of volatile species  (e.g., water,  nitrates, organics) in the aerosol is a function of
31      ambient temperature, the TEOM heats the inlet air stream to  a constant 50 °C to keep

        April 1995                               4-48      DRAFT-DO NOT QUOTE OR CITE

-------
                       Flow
Flow
                                                              Sampling Head
                                                       Heated Air Inlet
          Filter Cartridge
           Tapered Element
                                                                   Electronic
                                                                Feedback System
                                                                 Microprocessor
                                                      to Flow Controller
      Figure 4-14. TEOM
1     moisture in the vapor phase.  The mass transducer is also heated to 50 °C to stabilize the
2     measurement process.  Operation with the flow stream heated to a lower temperature (e.g.,
3     30 °C) is possible, but care must be taken to avoid moisture condensation that will confound
      April  1995                             4.49      DRAFT-DO NOT QUOTE OR CITE

-------
 1     the measurement.  The transducer is also heated to 50 °C to stabilize the mass measurement.
 2     A  factory calibration regression is used to electronically correct the computed mass from the
 3     TEOM to that measured by a reference PM10 sampler.
 4          As  previously described, several researchers, including Cahill et al. (1994), Hering
 5     (1994) and Meyer et al. (1992) have reported that the modification of the aerosol by the
 6     elevated  operating temperature appears to have a significant effect (loss) on mass
 7     concentration.  Meyer et al.  (1992) collocated a TEOM sampler with an PM10 SA1200
 8     gravimetric sampler in Mammoth Lakes, CA during a winter heating season (heavy wood
 9     stove usage).  The regressions between the TEOM and PM10 sampler gave strong
10     correlations (r2 > 0.98), with slopes of 0.55 for operation at 50 °C,  and 0.66 for operation
11     at 30 °C. The negative bias of the TEOM was attributed primarily to losses of semi-volatile
12     organics  from the filter.  Cahill et al.  (1994) reported that the TEOM showed biases on the
13     order of  30%  low and poor correlations with PM10  samplers in dry, dusty conditions.  The
14     reasons for this discrepancy were unknown.  The field comparison data of Patashnick and
15     Rupprecht (1990) showed near unity (1±0.06)  regression slopes for the TEOM with the
16     Wedding IP10 and Sierra-Andersen dichotomous samplers in El  Paso, TX and Birmingham,
17     AL. Since aerosol composition is highly dependent on local sources and meteorology,
18     volatilization losses could be expected to be site- and season-dependent.  More data are
19     needed to determine the implications of these problems on the ability of the TEOM to  be
20     used in a regulatory setting.
21
22     4.2.8.2  Beta Gauge
23          The Andersen FH 62I-N beta attenuation  sampler was described by Merrifield (1989)
24     and uses a 30 mCi Krypton-85 source and detector to determine the attenuation caused by
25     deposited aerosols on a filter  (see diagram in Figure 4-15).  To improve the stability over
26     time, a reference reading is periodically made of a foil with an  attenuation similar to that of
27     the filter and collected aerosol.  The Wedding beta attenuation sampler  was described by
28     Wedding and Weigand (1993) and uses a  100 mCi 14C source.  Both samplers have inlets
29     with a PM10 cutpoint, with the Andersen sampler operating at 16.67 1pm and the Wedding at
30     18.9 1pm.  The filter material is contained on a roll and advances automatically on a time
31     sequence, or when a preset aerosol loading is reached.  An automatic beta gauge sampler

       April 1995                               4.50      DRAFT-DO NOT QUOTE OR CITE

-------
\s

?
Tl
H
I
O
O
o
H
O
cj
O
H
W
O
                                             MC -
                                             CC -
                                             C  -
                                              Measuring Chamber
                                              Compensation Chamber
                                              Chamber for Dust Precipitation
                                              and Measurement
                                              30 m Ci KR-85 Source
                                              Filter Feed Spool
                                              Filter Takeup
                                              High-Voltage Power Supply
Bit
I/O


50-Pin
Connector
                        jmperature / Pressure


                         Rotary Vane Pump
                                                                V24/RS232
Figure 4-15. Beta gauge.
n

-------
 1     was also described by Spagnolo (1989), using a  15 /j,m inlet and a 14C source.  The
 2     calibration of a beta gauge is site specific, and a calibration regression must be processed
 3     electronically to provide accurate mass readings.  Rupprecht et al. (1992) suggested that the
 4     closer link between deposited mass and frequency shift for the TEOM principle should
 5     provide less  site-specific response, compared to the aerosol compositional sensitivity of the
 6     beta gauge technique.
 7           Arnold et al. (1992) provided data over a 2 year period in Denver, CO for the mass
 8     concentration regression data from a Wedding beta gauge, showing a  range of correlations
 9     (r2 from 0.72 to 0.86),  varying by sampler and season.  The authors suggested that
10     installation of a newer technology beta gauge accounted for the higher correlations, but noted
11     that unexplained outliers resulted in poorer than  expected results.  The regression slopes
12     between the two sampler types showed that the beta gauge averaged 19% lower than a
13     collocated Wedding PM10 gravimetric sampler.  Field data from Wedding and Weigand
14     (1993) at two sites (Fort Collins, CO and Cleveland, OH) using the same samplers produced
15     regressions exhibiting strong correlations (r2  = 0.99) with no apparent outliers and a
16     composite slope of 1.00.  Arnold et al.  (1992) operated the PM10 high volume samplers on
17     the required  every-6th-day schedule and the beta attenuation monitors continuously, and
18     noted that only 22.5% of the exceedance days, as measured by the beta monitor,  were
19     operational days for the high volume samplers.
20
21     4.2.8.3  Nephelometer
22           The integrating nephelometer is commonly used as a visibility monitor, measuring the
23     light scattered by aerosols, integrated over as  wide a range of angles as possible.   A
24     schematic diagram of the integrating nephelometer is shown in Figure 4-16 (from Hinds,
25     1982). The  measured scattering coefficient of particles, bsp, can be summed with the
26     absorption coefficient, b  , and the comparable coefficients for the gas phase, to compute  the
27     overall atmospheric extinction coefficient, bext.  The atmospheric extinction has been related
28     to visibility as visual range.  The particle scattering coefficient is particle size dependent,  as
29     shown by Charlson et al. (1968) in Figure 4-17, while the absorption coefficient is relatively
30     independent  of size.  The field calibration of nephelometers has historically been based on
31     the refractive index of Freon-12 (and occasionally carbon dioxide), but newer calibration

       April 1995                                4-52       DRAFT-DO  NOT QUOTE OR CITE

-------
VO
O
O
Z
O
H
O

g
H
W
O
90
n
»—i
3
         Power
        Supply
                                         Flash Tube
                                       Power Supply
          Clean Air
            Purge
                                      Aerosol
                                       Outlet
                                         A
     Photomultiplier
          Tube
                                     Scattering
                                      Volume
                         Collimating
                            Disks
                                                   Aerosol
                                                     Inlet
Clean Air
  Purge
Amplifier


Recorder
Figure 4-16. Integrating nephelometer.

-------
               10
o
tl
H
6
o
as
s
     co
     CO
  E  «
  o    £
       o


  I"E
 *•"•»   "*^
T-     "•"»«,  C
 »    T-    t)
     "o
     "b
        S
                                                                                     10
                                                     Diameter (urn)
4.00
s
w
 Figure 4-17. Particle-scattering coefficient as a function of particle size.
O
>— I
H
W

-------
  1      procedures using atomized sugar aerosols have been proposed (Horvath and Kaller, 1994) as
  2      more environmentally conscious.  Another class  of optical instruments which do not resolve
  3      the particle size spectrum, measure the laser light scattered from a volume of aerosol
  4      containing a number of particles.  Gebhart (1993) described devices such as the MIE,
  5      Inc.11.  MINIRAM, often used in portable applications to estimate real-time aerosol
  6      concentrations. Woskie et al. (1993) described the calibration and performance of a
  7      MINIRAM (using  the manufacturer's calibration) against gravimetric borate concentrations,
  8      and found significant biases  (a regression slope =  4.48).
  9           The relative insensitivity of the nephelometer to particles above  ~ 2  /xm provides for
10      poor correlations with PM10 mass.  Larson et  al. (1992) showed strong correlations (r2 =
11      0.945)  between bsp and Fine fraction mass (see Figure 4-19) for a woodsmoke impacted
12      neighborhood near Seattle, WA, with a slope of 4.89 m2/g.  They noted that this slope fell
13      within  the range of values reported by  others and was predicted by Mie scattering theory.
14      The slope of the Larson et al. (1992) data could  be compared with other site-specific
15      calibrations, such as the data of Waggoner and Weiss (1980), which gave  a composite slope
16      of 3.13 m2/g, characterized  by the authors as representative of a "wide range" of sites.
17      Lewis (1981) provided an analysis of the relationships of the features  of the ambient size
18      distribution to bsp.   The inlet air stream to the nephelometers for the latter data was heated
19      from 5 to  15 °C above background.  Rood et al. (1987) conducted a controlled comparison of
20      the influence of aerosol properties  on bsp in Riverside, CA, and reported a regression  slope
21      against fine mass (defined as less than 2.0 /mi) of 2.1 m2/g with an r2 value of 0.92.  In this
22      experiment the relative humidity for bsp determinations was controlled to less than 35% and
23      the gravimetric filter substrate was nylon.  They attributed the smaller than normal slope
24      reading to possible nitrate evaporation from the filtered aerosol and artifact reactions with the
25      nylon substrate material.
26           The data scatter in Figure 4-18 (if assumed to be typical of such comparisons) would
27      suggest that fine particle mass concentration estimates from bsp values were typically within 5
28      to 7 /ig/m3 of the gravimetrically determined values.  To be useful as  a surrogate measure
29      for mass concentration, the site-specific nephelometer calibration should be valid for a wide
30      range of situations,  especially during episodes  where the concentration levels approach or
31      "Bedford, MA.
        April 1995                                4.55       DRAFT-DO NOT QUOTE OR CITE

-------
       2.2
       2.0
       1.8
       1.6
       1.4   -
       1.2   -
       1.0   -
       0.8   -
       0.6   -
       0.4   ~
       0.2   ~
          0
         Lake Forest Park
        Weekly Average Values
January 17,1991 to December 19,1991
             0
                                            Slope = 4.89 m  /g
                                            R2= 0.945
               5       10      15      20      25      30      35      40
                                    PM25(ngm3)
Figure 4-18.  Correlation of bsp and fine fraction mass.
                                                                      45
1     exceed an action limit.  The scattergram of bsp versus fine particle mass provided by Rood
2     et al. (1987), showed much greater variability, with a given bsp value providing an estimated
3     20 to 25 /xg/m3 concentration range.  They noted that metastable H2O contributed 5 to 20%
4     of the total particle light scattering coefficient, especially during the late afternoon and early
5     evening.  The precisions and biases of the dependent and  independent variables between bsp
6     and Fine mass concentration are not constants, since at  least one factor - moisture content of
7     the aerosol - affects both measures.  The gravimetric sample filters are typically equilibrated
8     to a specific relative humidity range (e.g., 40 to 60%) to  normalize the tare weighings.
      April 1995
                               4-56
                                                  DRAFT-DO NOT QUOTE OR CITE

-------
 1           Sloane (1986) and others have noted that light scattering from particles is not solely a
 2      function of mass, but very dependent on a summation of the scattering coefficients of each
 3      specie.  The scattering cross section of a particle is dependent on the water content, and
 4      hence the relative humidity in situ.. Pre-heating of the inlet air of the nephelometer
 5      normalizes  the response to water content, but biases the reading relative to the in situ case.
 6      Sloane (1986) also gave the computed and measured scattering coefficients for ammonium
 7      sulfate, and noted that chemical interactions can cause  a two-fold variation in scattering
 8      response to a change in the mass of hygroscopic constituents.  It was  also observed that the
 9      light scattering efficiency of an aerosol such  as ammonium acid sulfate is not a constant, but
10      varies with the overall aerosol composition.  Eldering et al. (1994) developed and validated a
11      predictive model for bsp in Southern California.   This model used composite size
12      distributions constructed from a TSI, Inc.12 EAA, a PMS,  Inc.13 LAS-X and a Climet,
13      Inc.14 multi-channel OPC, and filter-based estimates of refractive indices for ammonium
14      sulfate, ammonium nitrate, organic carbon, elemental carbon and residual aerosol mass
15      concentrations  as independent variables.  The quality of their comparisons with nephelometer
16      data, suggested that this approach could be used to test models that predict visual range from
17      source emissions. Further research is needed to determine the role of the integrating
18      nephelometer as a predictor of Fine particle mass concentrations.
19
20      4.2.9  Specialized Sampling
21      4.2.9.1  Personal Exposure Sampling
22           The application of aerosol measurement technologies  to smaller and less obtrusive
23      samplers have  resulted in devices used as fixed-location indoor aerosol samplers and personal
24      exposure monitors (PEMs) worn on the body to estimate exposure.  The reduction in
25      physical size of personal aerosol sampling systems to reduce participant burden sometimes
26      results in poorer  aerosol collection performance  as compared to the outdoor counterparts.
27      Wiener and Rodes (1993) noted that personal sampling systems generally have poorer
28      precisions than outdoor aerosol samplers, due to the smaller sampler collections (from lower
29      ^Minneapolis, MN.
30      "Boulder, CO.
31      14Redlands, CA.
        April 1995                                4.57       DRAFT-DO NOT QUOTE OR CITE

-------
 1     flowrates) and poorer flow controllers.  Wallace (1994) reported biases for the Particle Total
 2     Exposure Assessment Methodology study averaging a factor of two between personal
 3     exposure measurements and fixed location PM10 concentrations.  He was unable to
 4     completely account for the biases, but attributed portions to proximity to indoor sources, a
 5     difference in inlet cutpoints (11.7 /urn vs 10.0 jum) and the collection of aerosols from the
 6     "personal cloud" caused by body dander. Rodes et al.  (1991) showed that the ratio of
 7     personal to indoor aerosol measurements for the EPA PTEAM study appeared to be log-
 8     normally distributed with a median value of 1.98 and an unexpectedly high value of 3.7 at
 9     the 90th ("most exposed") percentile.  Ingham and Yan (1994) suggested that the
10     performance of a personal aerosol sampling inlet in an isolated mode (without mounting on a
11     representative humanoid bluff body) can result in substantial under-sampling for larger
12     particles.  The relationship between measured aerosol exposure at some external location on
13     the body and actual uptake through oral and nasal entry is very complex.
14          Buckley et al. (1991) described the collection efficiency of an MSP, Inc.15 personal
15     aerosol sampler at 4.0 1pm as shown in Figure 4-19.  They evaluated this sampler in a field
16     comparison study with collocated PM10 high volume and dichotomous samplers. The
17     precision for the personal  sampler was found to be very good (CV = ±3.2%) with strong
18     correlations (r2 =  0.970) with the dichotomous samplers.  Lioy  et al. (1988) described a
19     similar comparison for a 10 1pm Air Diagnostics and Engineering, Inc.16 indoor air
20     sampler, with a PM10 inlet characterized by Marple et al. (1987). Correlations against the
21     PM10 dichotomous sampler were also described as very strong (r2 > 0.970), but noted a
22     substantial bias caused by  the loss of fragments from indoor air  sampler's glass fiber filters.
23     They recommended that exposure studies using samplers that collect small total volumes
24     should utilize filters with greater integrity, such as Teflon. Colome et al. (1992) describe an
25     indoor/outdoor sampling study using  an impactor [characterized  by Marple et al. (1987)] with
26     a PM10 cutpoint that had duplicate impactors with the same cutpoint in series.  This
27     sequential arrangement, in combination with a coating of 100 /xl of light oil, was used to
28     minimize particle bounce at 4.0 1pm for 24  h period.
29      15Minneapolis, MN.
30      16Naples, ME.
        April 1995                                4-58      DRAFT-DO NOT QUOTE OR CITE

-------
             100
              80
          'o
          it
          LJU
          "o
          o
              60
              20
                 1
10
                                     Aerodynamic Particle Diameter (jim)
        Figure 4-19.  Collection efficiency of the MSP personal aerosol sampler.
 1           Personal aerosol sampler systems have typically been characterized as burdensome
 2      (excessive weight, size, noise).  The success of passive detector badges for gaseous
 3      contaminants has recently prompted research into passive aerosol samplers. Brown et al.
 4      (1994) described a prototype aerosol sampler utilizing electrostatic charge to move the
 5      particles to a collection substrate.  They noted that preliminary results are  encouraging, but
 6      the effective sampling rate and size-selectivity  of the sampler was dependent on the electrical
 7      mobility of the aerosol.  This posed calibration problems for real aerosols with a distribution
 8      of electrical mobility's.  Hollander (1992) described a passive pulsed-corona sampler that has
 9      similar collection characteristics  as a PM10 inlet, with only modest wind speed dependence.
10           The performance characterization of PEMs has been considered for occupational
11      settings by Kenny and Liden (1989), who  reviewed the ACGIH, National Institute for
12      Occupational Safety and Health (NIOSH), and EPA PM10 aerosol sampler  performance
13      programs.  They proposed that an international consensus be reached on the basic principles
        April 1995                               4.59      DRAFT-DO NOT  QUOTE OR CITE

-------
 1     underlying the experimental protocols for testing personal samplers, as an essential
 2     prerequisite to the setting of standards.  An ISO working group has made progress in
 3     developing such a consensus (Kenny, 1992). As EPA becomes more  focused on exposure
 4     assessment and personal exposure sampling, it will become even more important for the
 5     agency to consider establishing performance specifications for personal aerosol samplers.
 6          Models have become powerful tools in understanding aerosol behavior in the vicinity of
 7     personal exposure samplers.  This is demonstrated by particle trajectory models that can
 8     predict the influences of the geometries and flow field on aerosol capture and losses (e.g.,
 9     Okazaki and Willeke, 1987, Ingham and Yan, 1994, and Tsai and Vincent,  1993).  These
10     models  have not only permitted more rapid design changes to accommodate new cutpoints
11     and flowrates, but have added insights as to the influence of air flow  obstructions on
12     sampling efficiencies.  Vincent and Mark (1982) suggested that there  is a critical particle
13     trajectory that determines whether a particle is  sampled or rejected by an inlet worn on the
14     body.  An extension of this model applicable to personal exposure sampling by Ingham and
15     Yan (1994) suggested that testing the performance of a personal aerosol sampling inlet in an
16     isolated mode (without mounting the inlet on a representative bluff body) can result in under-
17     sampling for larger particles by a factor of two.  Validation of this model may explain a
18     portion of the bias reported by Wallace et al. (1994) between personal and indoor sampler
19     measurements.
20
21     4.2.9.2  Receptor Model Sampling
22          Receptor modeling has become an established tool to relate  ambient concentrations of
23     pollutants to major source categories, by apportioning the components in collected ambient
24     aerosol samples using complimentary source "signatures".   Various approaches developed for
25     constructing source/receptor relationships were described by Henry et al.  (1984), who also
26     provided a review of modeling fundamentals.  They listed the advantages and disadvantages
27     of multivariate models and discussed multi-collinearity problems associated with the presence
28     of two or more sources with  nearly identical signatures. Javitz et al.  (1988) described the
29     basic Chemical Mass Balance (CMB) approach and showed the influence of the variance in
30     identifying a component in the source signature sample on the projected apportionment.
31     Dzubay et al. (1984) described aerosol source and receptor collection schemes that permitted

       April 1995                               4-60      DRAFT-DO  NOT QUOTE OR CITE

-------
  1      the separation of ambient samples into Fine and Coarse fractions for mass, elemental and
  2      volatile carbon, and metals analyses.  Stevens and Pace (1984) suggested the addition of
  3      Scanning Electron Microscopy to permit additional categorization using x-ray diffraction
  4      analysis.  The most widely used aerosol receptor model is the EPA CMB 7.0 model
  5      described by Watson et al.  (1990).  This paper  describes the structure of the model and
  6      computer code and the data requirements to evaluate the validity of the estimates.  Numerous
  7      papers have been published describing the applications of receptor models to the
  8      apportionment of the sources of aerosols, with the receptor modeling conference summary by
  9      Watson et al. (1989) descriptive of the state-of-the-art.
 10           Stevens et al. (1993) described (see Figure 4-20) a modified dichotomous sampler with
 11      a PM10 inlet, two  Fine channels operating at  15 1pm and one  Coarse channel operating at 2.0
 12      1pm, designated as the Versatile Air Pollution Sampler (VAPS).  The additional Fine fraction
 13      channel permitted  sampling on a 47 mm Teflon filter for elemental analysis and a 47 mm
 14      quartz filter for carbon speciation (elemental and volatile).  A Nuclepore filter was used on
 15      the Coarse  channel for Scanning Electron Microscopy (SEM) evaluation and energy
 16      dispersive x-ray diffraction analysis for selected particles.
 17
 18      4.2.9.3 Particle Acidity
 19           An emphasis was placed on sampling  sulfuric acidic aerosols in the  1982 Criteria
 20      Document.   This was followed by a number of  research efforts (e.g., Ferm, 1986; Koutrakis
 21      et al.,  1988) to identify and study the in situ rate reactions, develop sampling strategies to
 22      representatively remove the acid particle from the air, identify the co-existing reactive
 23      species (e.g., ammonia, nitric acid aerosol, aerosol sulfates and nitrates),  and protect the
 24      collected aerosol prior to analysis.  A "Standard"  and an "Enhanced" method were
25      subsequently described (U.S.  Environmental Protection Agency, 1992) for the determination
26      of aerosol acidity using annular denuder technology.  The "Standard" method did not account
27      for potential interferences from  nitric acid,  ammonium nitrate aerosol, or other ammonium
28      salts.  The  "Enhanced" method added an additional denuder prior to filtration, with nylon
29      and treated  glass fiber backup filters to account  for these species.  These sampling
30      technologies utilized either  an inlet impactor or  a cyclone with 2.5 /^m cutpoints to sample
        April 1995                                4.51       DRAFT-DO NOT QUOTE OR CITE

-------
                                                -   o-
                                                Q> _^ *^ E

                                                S,2tS§
                                                CO 3 0)0-
                                                          E
                                                II
                                                N w
                                                ll
                                                o"-
                                                     Q. CD

                                                     if
                        o
                        00
                                                co
                                                CO
                                                0)
                                                I
                                                             i
                                                             Of)
April 1995
4-62
DRAFT-DO NOT QUOTE OR CITE

-------
 1     the Fine fraction.  This technology has recently been extended to other reactive aerosol
 2     systems, including semi-volatile organics (e.g., Vossler et al., 1988).  Bennett et al. (1994)
 3     describe a PM2 5 cyclone-based, filter pack sampling system designed fine particle network
 4     sampling and acidity measurements, as part of the Acid MODES program.  The sampler
 5     operated at 8.8 1pm, and was designed to selectively remove ammonia, speciate gas and
 6     particle phase sulfur compounds, as well as collect gas phase nitric acid.  An intercomparison
 7     of 18 nitric acid measurement methods was reported by Hering et al. (1988), who noted that
 8     measurements differed by as much as a factor of four and biases increased as nitric acid
 9     loadings increased. In general the filter pack systems reported the highest acidity
10     measurements, while the denuder-difference techniques reported significantly lower
11     measurements.  Benner et al. (1991) in a comparison of the SCENES filter pack sampler
12     with a denuder-based sampler found excellent agreement between sampler types for both
13     nitric acid and total nitrates.  They attributed the close agreement to limited positive artifact
14     formations, since the test field site had high nitric acid gas to paniculate nitrate ratios.  John
15     et al. (1988) noted that internal aluminum  sampler surface denude nitric acid, and describe
16     the design of an aluminum denuder for the inlet of a commercially available dichotomous
17     sampler to quantitatively remove nitric acid for extended periods.
18          Brauer et al.  (1989) describe the design of a miniature personal sampler to collect acid
19     aerosols and gases.  A significant finding was the lower than expected personal acidity
20     levels, attributed to the "personal cloud" production of ammonia by the body. Personal
21     exposure levels of acid aerosols were reported to be lower than indoor measurements.
22
23
24     4.3   ANALYSIS OF PARTICIPATE MATTER
25          The interest in the composition of aerosol particles  lies in the areas of:  (1) explaining
26     and inventorying the observed mass, (2) establishing the  effect of aerosols on health and
27     welfare, and (3) attributing ambient aerosols to pollution sources. While any compositional
28     measurement will address one or more of these goals, certain methods excel for specific
29     tasks.  In general,  no single method can measure all chemical species, and comprehensive
30     aerosol characterization programs use a combination of methods to address complex needs.
31     This allows each method to be optimized for its objective, rather than be compromised to

       April 1995                               4-63      DRAFT-DO NOT QUOTE OR CITE

-------
 1     achieve goals unsuitable to the technique.  Such programs also greatly aid quality assurance
 2     objectives, since confidence may be placed in the accuracy of a result when it is obtained by
 3     two or more methods on different substrates and independent samplers.
 4           In the sections that follow, some of the more commonly used methods that address the
 5     goals state above are described.  The  sections are designed to be illustrative rather than
 6     exhaustive, since new methods are constantly appearing as old methods are being improved.
 7     These chemical analysis methods for the following section are  divided into four categories:
 8     1) mass, 2) elements, 3) water-soluble ions, and 4) organics. Material balance comparing
 9     the sum of the chemical species to the PM mass concentrations show that elements,  water
10     soluble ions, and organic and elemental  carbon typically explain 65 to 85% of the measured
11     mass and are adequate to characterized the chemical composition of measured mass  for filter
12     samples collected in most urban and non-urban areas.   Some of these chemical analysis
13     methods are non-destructive, and these are preferred because they preserve the filter for
14     other uses.  Methods which require destruction of the filter are best performed on a section
15     of the filter to save a portion of the filter of other analyses or as a quality control check on
16     the same analysis method.  Table 4-2 identifies the elements and chemical compounds
17     commonly found in air using these methods  with typical detection limits.
18           Less common analytical methods,  which are applied to a small number of specially-
19     taken samples,  include isotopic abundances (Jackson,  1981; Currie, 1982; Hirose and
20     Sugimura, 1984); mineral compounds (Davis,  1978, 1980; Schipper et al., 1993); and
21     function groups (Mylonas et al., 1991; Palen et al., 1992; 1993; Allen et al., 1994). Recent
22     advances  in infrared optics and detectors have resulted in the quantitative determination of
23     the major functional groups (e.g., sulfate,  nitrate, aliphatic carbons, carbonyl carbons,
24     organonitrates, and alcohols) in the atmospheric aerosol (Allen et al.,  1994).  The advantages
25     of functional analysis in source apportionment are that the number of functional groups is
26     much less than the number of organic compounds to be classified.  The cited references
27     provide information on sampling and  analysis methods for these highly-specialized methods.
28           The following  section focuses on:
29           •    Physical analysis of elements and single particle size, shape, and composition,
30
31           •    Wet chemical analysis of anions and cations, and
        April 1995                               4-64       DRAFT-DO NOT QUOTE OR CITE

-------
          TABLE 4.2. INSTRUMENTAL DETECTION LIMITS FOR
                      PARTICLES ON FILTERS
Minimum Detection Limit in ng/m3a
Species
Be
Na
Mg
Al
Si
P
S
Cl
K
Ca
Sc
Ti
V
Cr
Mn
Fe
Co
Ni
Cu
Zn
Ga
As
Se
Br
Rb
Sr
Y
Zr
Mo
Pd
ICP/
AESM
0.06
NA
0.02
20
3
50
10
NA
NA
0.04
0.06
0.3
0.7
2
0.1
0.5
1
2
0.3
1
42
50
25
NA
NA
0.03
0.1
0.6
5
42
AA
FlameM
2d
0.2d
0.3
30
85
100,000
NA
NA
2d
ld
50
95
52
2
1
4
6d
5
4
1
52
100
100
NA
NA
4
300
1000
31
10
AA
Furnaceb
0.05
< 0.05
0.004
0.01
0.1
40
NA
NA
0.02
0.05
NA
NA
0.2
0.01
0.01
0.02
0.02
0.1
0.02
0.001
NA
0.2
0.5
NA
NA
0.2
NA
NA
0.02
NA
INAAb'f
NAh
2
300
24
NA
NA
6,000
5
24
94
0.001
65
0.6
0.2
0.12
4
0.02
NA
30
3
0.5
0.2
0.06
0.4
6
18
NA
NA
NA
NA
FIXES
NA
60
20
12
9
8
8
8
5
4
NA
3
3
2
2
2
NA
1
1
1
1
1
1
1
2
2
NA
3
5
NA
XRFC
NA
NA
NA
5
3
3
2
5
3
2
NA
2
1
1
0.8
0.7
0.4
0.4
0.5
0.5
0.9
0.8
0.6
0.5
0.5
0.5
0.6
0.8
1
5
ICb
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
ACb
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
TORb
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
April 1995
4-65
DRAFT-DO NOT QUOTE OR CITE

-------
       TABLE 4.2 (cont'd). INSTRUMENTAL DETECTION LIMITS FOR
                      PARTICLES ON FILTERS
Minimum Detection Limit in ng/m3a
Species
Ag
Cd
In
Sn
Sb
I
Cs
Ba
La
Au
Hg
Tl
Pb
Ce
Sm
Eu
Hf
Ta
W
Th
U
Cl-
NO3-
S04=
NH4+
ICP/
AESM
1
0.4
63
21
31
NA
NA
0.05
10
2.1
26
42
10
52
52
0.08
16
26
31
63
21
NA
NA
NA
NA
AA
FlameM
4
1
31
31
31
NA
NA
8d
2,000
21
500
21
10
NA
2,000
21
2,000
2,000
1,000
NA
25,000
NA
NA
NA
NA
AA
Furnaceb
0.005
0.003
NA
0.2
0.2
NA
NA
0.04
NA
0.1
21
0.1
0.05
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
INAAb-f
0.12
4
0.006
NA
0.06
1
0.03
6
0.05
NA
NA
NA
NA
0.06
0.01
0.006
0.01
0.02
0.2
0.01
NA
NA
NA
NA
NA
PIXEg
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
3
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
XRFC
6
6
6
8
9
NA
NA
25
30
2
1
1
1
NA
NA
NA
NA
NA
NA
NA
1
NA
NA
NA
NA
ICb
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
50
50
50
NA
ACb
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
50
TORb
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
April 1995
4-66
DRAFT-DO NOT QUOTE OR CITE

-------
                 TABLE 4.2 (cont'd).  INSTRUMENTAL DETECTION LIMITS FOR
                                      PARTICLES ON FILTERS
                                  Minimum Detection Limit in ng/m
                                                                  ,3a
                   ICP/      AA       AA
       Species   AESb-d  FlameM  Furnace5   INAAb-f   PIXR8   XRFC    ICb   ACb  TORb
oc
EC
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
NA
100
100
        "Minimum detection limit is three times the standard deviation of the blank for a filter of 1 mg/cm2 areal
        density.
        ICP/AES = Inductively Coupled Plasma with Atomic Emission Spectroscopy.
        AA = Atomic Absorption Spectrophotometry.
        PIXE = Proton Induced X-ray Emissions Spectrometry.
        XRF = Non-Dispersive X-ray Fluorescence Spectrometry.
        INAA = Instrumental Neutron Activation Analysis.
        1C = Ion Chromatography.
        AC = Automated Colorimetry.
        TOR = Thermal Optical Reflectance.
        bConcentration is based on the extraction of 1/2 of a 47 mm quartz-fiber filter in 15 ml of deionized-distilled
        water, with a nominal flow rate of 20 L/min for 24-h samples.
        cConcentration is based on 13.8 cm2 deposit area for a 47 mm ringed teflon-membrane filter, with a nominal
        flow rate of 20 L/min for 24-h samples with 100 sec radiation time.
        dHarman (1989).
        Ternandez (1989).
        f01mez (1989).
        SEldred (1993).
        hNot Available.
 1           •   Organic analysis of organic compounds and elemental/organic carbon.
 2

 3      4.3.1   Mass Measurement Methods

 4           Paniculate mass concentration is the most commonly made measurement on aerosol
 5      samples. It is used to determine compliance with PM10 standards and to select certain
 6      samples for more detailed, and more expensive, chemical analyses.  As noted in Section 2,
 7      the beta attenuation and inertial microbalance methods have  been incorporated into in situ

 8      measurement systems which acquire real-time mass measurements.  Gravimetric analysis is
 9      used almost exclusively to obtain mass measurements of filters in a laboratory environment.
10      U.S.  Environmental Protection Agency (1976) and Watson et al.  (1989a) have published
11      detailed procedures for mass analyses associated with 20.32  cm x 25.40 cm fiber filters, but
12      the guidance for other types of filters used for chemical analyses  is less well documented.


        April 1995                                4.57       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Gravimetry measures the net mass on a filter by weighing the filter before and after
 2      sampling with a balance in a temperature- and relative humidity-controlled environment.
 3      PM10 reference methods require that filters be equilibrated for 24 h at a constant (within
 4      +5%) relative humidity between 20 and 40% and at a constant (within +3 °C) temperature
 5      between 15 and  30 °C. These  are intended to minimize the liquid water associated with
 6      soluble compounds and to minimize the loss of volatile species.  Nominal values of 30% RH
 7      and 15  to 20 °C best conserve the particle deposits during sample weighing.
 8           Balances used to weigh 20.32 cm  x 25.40 cm filters from high volume PM10 samples
 9      must have  a sensitivity of at least 100 /ig.  Balances used for medium volume PM10 samples
10      should have a sensitivity of at least 10 ;wg, and those used for low-volume  PM10 samples
11      should have a sensitivity of at least 1 /*g.  Modifications to the balance chamber are
12      sometimes needed to accommodate filters of different sizes.  All filters,  even those from
13      high-volume PM10 samplers, should be  handled with gloved hands when subsequent chemical
14      analyses are a possibility.
15           Balance calibrations  should be established before and after each weighing session using
16      Class M and Class S  standards,  and they should  be verified with a standard mass every ten
17      filters.  Approximately one out of ten filters should be re-weighed by a different person at a
18      later time.   These re-weights should be  used to calculate the precision of the measurement as
19      outlined by Watson et al.  (1989b).
20           Feeney et al. (1984) examined the gravimetric measurement of lightly loaded membrane
21      filters and  obtained excellent precision and accuracy.  The sensitivity of the electrobalance is
22      about ± 0.001 mg, though tolerances on re-weights of Teflon-membrane filters are typically
23      ± 0.010 mg.  The main interference in gravimetric analysis of filters results from
24      electrostatic effects.  Engelbrecht et al.  (1980) found that residual charge on a filter could
25      produce an electrostatic interaction between the filter on the pan and the metal casing of the
26      electrobalance.  This  charge can be removed by  exposing the filter to a radioactive polonium
27      source before and during sample weighing.
28           Beta  attenuation methods have been applied in the laboratory as well as in the field, and
29      the results are comparable to those of gravimetric measurements. The precision of
30      beta-gauge measurements  has been shown to be  ± 5 /xg/m3 or better for counting intervals of
31      one minute per sample, which translates into ±  32 /xg/filter for 37 mm  diameter substrates.

        April 1995                               4-68      DRAFT-DO NOT QUOTE OR CITE

-------
 1     This is substantially higher than the  ± 6 jig/filter precision determined by gravimetric
 2     analysis using an electrobalance (Feeney et al., 1984).  Jaklevic et al. (1981) found
 3     equivalent accuracy and precision for both techniques as they were used  in that study.
 4     Courtney et al. (1982) found beta attenuation and gravimetric mass measurements to differ
 5     by less than + 5%.  Patashnick and Rupprecht (1991) examine results from TEOM samplers
 6     operated alongside filter-based PM10 samplers and Barnes et al. (1988) and Shimp (1988)
 7     report comparisons with beta attenuation field monitors; these comparisons all show good
 8     agreement for mass measurements.
 9
10     4.3.2  Physical Analysis
11           The most common interest in elemental composition derives from concerns about health
12     effects and the utility of these elements to trace the sources of suspended particles.
13     Instrumental neutron activation analysis (INAA), photon-induced x-ray fluorescence (XRF),
14     particle-induced x-ray emission (PIXE), atomic absorption spectrophotometry (AAS),
15     inductively-coupled plasma with atomic emission spectroscopy (ICP/AES), and scanning
16     electron microscopy  with x-ray fluorescence (SEM/XRF) have all been applied to elemental
17     measurements of aerosol samples. AAS and ICP/AES are also appropriate for ion
18     measurements when  the particles are extracted in deionized-distilled water (DDW). Since air
19     filters contain very small particle deposits (20 to 100 jug/cm2), preference is given to methods
20     that can accommodate small sample  sizes.  XRF and PIXE leave the sample intact after
21     analysis so that it can be submitted to additional examinations by other methods.  Excellent
22     agreement was found for the intercomparison  of elements acquired  form the XRF and PIXE
23     analyses (Cahill, 1980). The  analytical measurement specifications of air filter samples for
24     the different elemental analysis is shown in Table 4.2.
25
26     4.3.2.1  X-Ray Fluorescence (XRF) of Trace Elements
27           In x-ray fluorescence (XRF) (Dzubay and Stevens, 1975; Jaklevic et al., 1977; Torok
28     and Van Grieken,  1994), the filter deposit is irradiated by high energy x-rays that eject inner
29     shell  electrons from the atoms of each element in the sample.  When a higher energy electron
30     drops into the vacant lower energy orbital, a fluorescent x-ray photon is  released.  The
31     energy of this photon is unique to each element, and the number of photons is proportional to

       April 1995                               4-69      DRAFT-DO NOT QUOTE OR CITE

-------
 1     the concentration of the element.  Concentrations are quantified by comparing photon counts
 2     for a sample with those obtained from thin-film standards of known concentration.
 3           Emitted x-rays with energies less than =4 KeV (affecting the elements sodium,
 4     magnesium, aluminum, silicon, phosphorus, sulfur, chlorine, and potassium) can be  absorbed
 5     in the filter,  in a thick particle deposit, or even by large particles in which these elements are
 6     contained.  Very thick filters also scatter much of the excitation radiation or protons, thereby
 7     lowering the signal-to-noise ratio.  For these reasons, thin membrane  filters with deposits in
 8     the range of 10 to 50 pig/cm2 provide the best  accuracy and precision for XRF.
 9           XRF  methods can be  broadly divided into two categories:  wavelength dispersive x-ray
10     fluorescence (WDXRF), which utilizes crystal  diffraction for observation of fluorescent
11     x-rays,  and energy dispersive x-ray fluorescence (EDXRF), which uses a silicon
12     semiconductor  detector. The WDXRF method is characterized by high spectral resolution,
13     which minimizes peak overlaps. It  requires high power excitation to  overcome low
14     sensitivity, resulting in excessive sample heating and potential degradation.  Conversely,
15     EDXRF features high sensitivity but less spectral resolution, requiring complex spectral
16     deconvolution procedures.
17           XRF  methods can be  further categorized  as direct/filtered excitation, where the x-ray
18     beam from the tube is optionally filtered and then focused directly on the sample, or
19     secondary  target excitation, where the beam is focused  on a target of  material selected to
20     produce x-rays of the desired energy. The secondary fluorescent radiation is then used to
21     excite the samples.  The direct/filtered approach has the advantage of delivering higher
22     incident radiation flux to the sample for a given x-ray tube power, since about 99% of the
23     incident energy is lost in a  secondary fluorescence.  However, the secondary fluorescence
24     approach, produces a more nearly monochromatic excitation that reduces unwanted scatter
25     from the filter, thereby yielding better detection limits.
26           XRF  is usually performed on Teflon-membrane filters for sodium, magnesium,
27     aluminum, silicon, phosphorus, sulfur, chlorine, potassium, calcium,  titanium, vanadium,
28     chromium, manganese, iron, cobalt, nickel, copper, zinc, gallium, arsenic,  selenium,
29     bromine, rubidium, strontium, yttrium,  zirconium, molybdenum, palladium, silver,
30     cadmium,  indium, tin, antimony, barium, lanthanum, gold, mercury,  thallium, lead, and
31     uranium.

       April 1995                              4-70        DRAFT-DO  NOT QUOTE OR CITE

-------
  1           A typical XRF system is schematically illustrated in Figure 4-21.  The x-ray output
  2      stability should be within ±0.25% for any 8-h period within a 24-h duration.  Typically,
  3      analyses are controlled, spectra are acquired, and elemental concentrations are calculated by
  4      software on a computer that is interfaced to the analyzer.
  5           Separate XRF analyses are conducted on each sample to optimize detection limits for
  6      the specified elements.  A comparison of the minimum detectable limits of Teflon-membrane
  7      and quartz-fiber filters is listed in Table 4-3.  Figure 4-22 shows an example of an XRF
  8      spectrum.
  9           Three types of XRF standards are  used for calibration, performance testing, and
10      auditing:   1) vacuum-deposited thin-film elements and compounds (Micromatter); 2) polymer
11      films (Dzubay et al., 1981); and 3) National Institute of Science and Technology (NIST,
12      formerly NBS) thin-glass films.  The thin film standards cover the largest number of
13      elements and are used to establish calibration curves, while the polymer film standards are
14      used to verify the accuracy of the thin film standards.  The NIST standards are used to
15      validate the accuracy of the calibration curves.  NIST produces the definitive standard
16      reference materials, but these are only available for the species of aluminum, silicon,
17      calcium, iron, cobalt, copper,  manganese, and uranium (SRM 1832), and silicon, potassium,
18      titanium, iron, zinc, and lead (SRM 1833).  One or more separate Micromatter thin-film
19      standards are used to calibrate the system for each element.
20           Sensitivity factors (number of x-ray counts per jug/cm2 of the element)  are determined
21      for each excitation condition.  These factors  are then adjusted for absorption of the incident
22      and emitted radiation in the thin film.  These sensitivity factors are plotted as a function of
23      atomic number and a smooth curve is  fitted to the experimental values.  The calibration
24      sensitivities are then read from these curves for the atomic numbers  of each element in each
25      excitation condition.  NIST standards are analyzed on a periodic basis to verify the sensitivity
26      factors.  A multi-layer thin  film standard prepared  by Micromatter is analyzed  with each set
27      of samples to check the stability of the instrument response.  When deviations from specified
28      values are  greater than ±5%, the system should be re-calibrated.
29          The sensitivity factors are multiplied by the net peak intensities yielded by ambient
30      samples to obtain the /*g/cm2 deposit for each element. The net peak intensity  is obtained
        April 1995                                4.7!       DRAFT-DO NOT QUOTE OR CITE

-------
  Sample
              ^Characteristic  Silicon detector
              _x-rays       /   FET
 X-ray excitation
Secondary
target
                  Be
                  window
                                    ' - ,/
                                        Analog-to-
                                        digital
                                        converter
                               Anode

                             Electron beam
                      X-ray tube
                Data output
       Video
       display
                                                    Signal
                                                    processing
                                                Data
                                                handling
Figure 4-21. Schematic of a Typical X-Ray Fluorescence (XRF) System.
April 1995
                           4-72
DRAFT-DO NOT QUOTE OR CITE

-------
     TABLE 4-3. MINIMUM DETECTABLE LIMITS3 FOR XRF ANALYSIS
                        OF AIR FILTERS
Element
Al
Si
P
S
Cl
K
Ca
Ti
V
Cr
Mn
Fe
Co
Ni
Cu
Zn
Ga
As
Se
Br
Rb
Sr
Y
Zr
Mo
Pd
Ag
Cd
In
Condition
Numberd
5
5
5
5
4
4
4
3
3
3
3
3
3
3
3
3
2
2
2
2
2
2
2
2
4
1
1
1
1
Quartz-Fiber
Filter13
Protocol QA-
A ng/cm2 e
NAf
NA
NA
40g
30
40
100
50
20
8
7
15
5
4
4
6
8
9
5
5
5
8
8
10
20
20
20
25
30
Teflon Membrane Filter0
Protocol A
ng/cm2 d
10
6.3
5.6
5.0
10
6.1
4.5
2.9
2.5
1.9
1.6
1.5
0.88
0.89
1.1
1.1
1.9
1.6
1.2
1.0
1.0
1.1
1.3
1.7
2.7
11
12
12
13
Protocol B
ng/cm2
7.2
4.4
4.0
3.5
7.4
4.3
3.2
2.1
1.7
1.4
1.1
1.1
0.62
0.63
0.76
0.76
1.4
1.1
0.86
0.72
0.68
0.78
0.92
1.2
1.9
7.6
8.6
8.6
9.5
Protocol C
ng/cm2
3.6
2.2
2.0
1.8
3.7
2.2
1.6
1.0
0.87
0.67
0.56
0.54
0.31
0.31
0.38
0.38
0.68
0.56
0.43
0.36
0.34
0.39
0.46
0.59
0.95
3.8
4.3
4.3
4.8
Protocol D
ng/cm2
2.5
1.4
1.4
1.2
2.6
1.5
1.1
0.73
0.62
0.48
0.40
0.38
0.22
0.22
0.27
0.27
0.48
0.39
0.31
0.25
0.24
0.28
0.33
0.42
0.67
2.7
3.0
3.0
3.4
April 1995
4-73
DRAFT-DO NOT QUOTE OR CITE

-------
   TABLE 4-3.  (cont'd)  MINIMUM DETECTABLE LIMITS3 FOR XRF ANALYSIS
                                       OF AIR FILTERS
Element
Sn
Sb
Ba
La
Au
Hg
Tl
Pb
U
Condition
Numberd
1
1
1
1
2
2
2
2
2
Quartz-Fiber
Filterb
Protocol QA-
A ng/cm2 e
40
50
170
190
NA
20
NA
14
NA
Teflon Membrane Filter0
Protocol A
ng/cm2 d
17
18
52
62
3.1
2.6
2.5
3.0
2.3
Protocol B
ng/cm2
12
13
37
44
2.2
1.8
1.8
2.2
1.7
Protocol C
ng/cm2
6.2
6.4
18
22
1.1
0.91
0.88
1.1
0.83
Protocol D
ng/cm2
4.4
4.5
13
16
0.77
0.65
0.62
0.76
0.59
aMDL defined as three times the standard deviation of the blank for a filter of 1 mg/cm2
 areal density.
bAnalysis times are 100 sec. for Conditions 1 and 4, and 400 sec. for Conditions 2 and 3.
 Actual MDL's for quartz filters vary from batch to batch due to elemental contamination
 variability.
Standard protocol, developed at the Desert Research Institute, University and Community
 College System of Nevada, Reno, NV, analysis times are 100 sec. for Conditions 1, 4 and
 5, and 400 sec. for Conditions 2  and 3 for Protocol A; 200 sec. for Conditions 1, 4 and
 5 and 800 sec. for Conditions 2 and 3 for Protocol B; 800 sec. for Conditions 1,4 and 5
 and 3,200 sec. for Conditions 2 and 3 for Protocol C; and 1600 sec. for Conditions 1, 4
 and 5 and 6400 sec. for Conditions 2 and 3 for Protocol D.
dCondition 1 is direct mode excitation with a primary excitation filter of 0.15 mm thick Mo.
 Tube voltage is 50 KV and tube current is 0.6 mA.  Condition 2 is direct mode excitation
 with a primary excitation filter of 0.13 mm thick Rh. Tube voltage is 35 KV and tube
 voltage is 2.0 mA.  Condition 3 uses Ge secondary target excitation with the secondary
 excitation filtered  by a Whatman  41 filter.  Tube voltage is 30 KV and tube current is
 3.3  mA.  Condition 4 uses Ti secondary target excitation with the secondary excitation
 filtered by  3.8 /tm thick mylar film. Tube voltage is 30 KV and tube current is 3.3 mA.
 Condition 5 uses direct mode excitation with a primary excitation filter consisting of
 3 layers of Whatman 41 filters. Tube voltage is 8 KV and tube current  os 0.6 mA.
 Multi-channel analyzer energy range is 0 to 40 KeV for condition 1, 0 - 20 KeV for
 condition 2, and 0 to 10 KeV for conditions 3,4, and 5.
eTypical exposed area is 406 cm2  for standard high-volume filters; 6.4 cm2 for 37 mm ringed
 Teflon-membrane  filters; and 13.8 cm2 for 47 mm ringed Teflon-membrane filters.
Information not available.
gFor condition 4.
April  1995                                    4.74        DRAFT-DO NOT QUOTE OR CITE

-------
                J_
_L
           26-Oct-1992 18:09:56
           SJTT046
           Vert=    2000 counts Disp= 1
                                   Preset=     100 sees
                      Comp=  2   Elapsed=    400 sees
           San Jose, 1/21/92,  PM 10
           18:01  - 06:00
           Excitation Condition 3
                         Fe
                                                                      Ge i sec.
                                                                      target
                                                                      scatter
                     Si
                        S      K  Ca
                         &  Cl  II  ;!
                                           Zn
                  Al
                          ! /  I
                Ti
                                 Fe
                                       Cu
                   V  Cr
                                                Mn
                                                           Ni
                  0.320    Range=    10.230 keV
                                                      Integral 0  -
                                              10.230  >
                                                243425
                 I      I       I      I       |      I       I       I
                                          5

      Figure 4-22. Example of an X-Ray Fluorescence  (XRF) Spectrum.
      Source:  Chow et al. (1990).
                                                     I
                                                    10
1     by:  (1) subtracting background radiation; (2) subtracting spectral interferences; and
2     (3) adjusting for x-ray absorption.
3          The elemental x-ray peaks reside on a background of radiation scattered from the
4     sampling substrate.  A model background is formed by averaging spectra obtained from
5     several blank filters of the same type used in ambient sampling.  It is important to retain
6     blank filters for this purpose when XRF analysis is anticipated. This model background has
7     the same shape and features of the sample spectra (minus the elemental peaks) if the deposit
8     mass is small relative to the substrate  mass  (Russ, 1977). This model background is
      April 1995
                     4-75
DRAFT-DO NOT QUOTE OR CITE

-------
  1      normalized to an excitation radiation scatter peak, or the background areas of the spectrum
  2      that have no elemental peaks,  in each sample spectrum to account for the difference in scatter
  3      intensity due  to different masses.
  4           The number and spacing of the characteristic x-ray lines relative to detector resolution
  5      are such that  the peaks from one  element can interfere with a peak from another element
  6      (Dzubay, 1986).  A variety of methods has been used to subtract these peak overlaps (Arinc
  7      et al., 1977; Parkes et al., 1979;  Drane et al., 1983), including least squares fitting to library
  8      spectra, Gaussian and other mathematical functions, and the use of peak overlap coefficients.
  9           Peak overlap coefficients are applied to aerosol deposits.  The most important of these
10      overlaps are the K-beta to K-alpha overlaps  of elements that increase in atomic number from
11      potassium to zirconium, the lead L-alpha to  arsenic K-alpha interference, and  the lead M line
12      to sulfur K line interference.   The ratios of overlap peaks to the primary peak are determined
13      from the thin film standards for each element for the spectral regions of the remaining
14      elements.  These ratios are multiplied by the net peak intensity of the primary peak and
15      subtracted from the spectral regions of other elements.
16           The ability of an x-ray to penetrate matter depends on the energy of the  x-ray and the
17      composition and thickness of the material.  In general, lower energy x-rays, characteristic of
18      light elements, are absorbed in matter to a much greater degree than higher energy x-rays.
19      XRF analysis of air paniculate samples has had widest application to samples  collected on
20      membrane-type filters such as  Teflon- or polycarbonate-membrane filter substrates. These
21      membrane filters collect the deposit on their surfaces, which eliminates biases  due to
22      absorption of x-rays by the filter material.  These filters also have a  low areal density which
23      minimizes the scatter of incident x-rays, and their inherent trace element content is very  low.
24           Quartz-fiber filters used for  high-volume aerosol sampling do not exhibit these features.
25      As noted earlier, blank elemental  concentrations in quartz-fiber filters that have not
26      undergone acceptance testing can  be several  orders of magnitude higher than the
27      concentrations in the paniculate deposits.  The concentrations  vary substantially among the
28      different types of quartz-fiber  filters, and even within the same filter type and  manufacturing
29      lot. Blank impurity concentrations and their variabilities decrease  the precision of
30      background subtraction from the XRF spectral data, resulting in higher detection limits.
31      Impurities observed in various types of glass- and quartz-fiber filters include aluminum,

        April 1995                                4.75       DRAFT-DO NOT QUOTE OR CITE

-------
  1      silicon, sulfur, chlorine, potassium, calcium, iron, nickel, copper, zinc, rubidium, strontium,
  2      molybdenum, barium, and lead.  Concentrations for aluminum, silicon, phosphorus, sulfur,
  3      and chlorine cannot be determined for quartz-fiber filters because of the large silicon content
  4      of the filters.
  5           Quartz-fiber filters also trap particles within the filter matrix,  rather than on the
  6      surface.  This causes absorption of x-rays within the filter fibers yielding  lower
  7      concentrations than would otherwise be measured.  The magnitude  of this absorption
  8      increases exponentially as the atomic  number of the measured element decreases, and varies
  9      from sample to sample. Absorption factors generally are  "1.2" or  less for iron and  heavier
10      elements, but can be from  "2" to "5" for sulfur.
11           Quartz-fiber filters are much thicker than membrane filters resulting in an increased
12      scattering of x-rays and a  consequent increase in background and degradation  of detection
13      limits.  The increased x-ray scatter also overloads the x-ray detector which requires  samples
14      to be analyzed at a lowered x-ray intensity. These effects alone can result in degradation of
15      detection limits by up to a factor of 10 with respect to Teflon-membrane substrates.
16           Larger particles collected during aerosol sampling have sufficient size  to cause
17      absorption of x-rays within the particles. Attenuation factors for fine particles (PM2 5,
18      particles with aerodynamic diameters  equal to or less than 2.5 jum)  are generally  negligible
19      (Criss, 1976), even for the lightest elements, but these attenuations  can be significant for
20      coarse fraction particles (particles with aerodynamic diameters from 2.5 to 10 /mi).
21      Correction factors for XRF have been derived using the theory of Dzubay and Nelson (1975)
22      and should  be applied to coarse particle measurements.
23           During XRF analysis,  filters are removed from their Petri slides  and placed with their
24      deposit sides down into filter cassettes.  These cassettes are loaded  into a  mechanism that
25      exposes the filter deposits  to x-rays. The sample chamber  is evacuated and a computer
26      program controls the positioning of the samples and the excitation conditions.  The vacuum in
27      the x-ray chamber and the heat induced by the absorption  of x-rays can cause  certain
28      materials to volatilize.  Therefore, labile species such as nitrate and organic carbon are better
29      measured on a quartz-fiber filter that  samples simultaneously with the Teflon-membrane
30      filter.
        April 1995                                4.77       DRAFT-DO NOT QUOTE OR CITE

-------
 1           Quality control standards and replicates from previous batches should be analyzed for
 2      every 10 to 20 samples.  When quality control results differ from specifications by more than
 3      +5%, or if the replicate concentrations differ from the original values (assuming they are at
 4      least 10  times detection limits) by more than ±10%, the samples should be re-analyzed.
 5
 6      4.3.2.2   Particle Induced X-Ray Emission (PIXE) of Trace Elements
 7           Particle Induced X-Ray Emission (PIXE) is another form of elemental analysis based on
 8      the characteristics of x-rays and the nature of x-ray detection (Cahill et al., 1975; 1987;
 9      1989; 1993; Cahill, 1990).  PIXE uses beams of energetic  ions,  consisting of protons at an
10      energy level of 2 to 5 MeV, to create inner electron shell vacancies.  As inner electron shell
11      atomic vacancies are filled by outer electrons, the emitted characteristics of x-rays can  be
12      detected by wavelength dispersion, which is scattering from a crystal, or by energy
13      dispersion, which involves direct conversion of x-rays.  The development  of focusing
14      energetic proton beams (proton microprobes) has expanded the application of PIXE from
15      environmental and biological sciences to geology and material sciences.  Figure 4-23
16      illustrates a typical PIXE setup in a thin target mode (Cahill,  1989).  PIXE analysis is  often
17      used for impactor samples  or small filter substrates, since proton beams can be focused to a
18      small area with no loss of sensitivity (Cahill, 1993).
19           Very thick filters or thick particle deposits on filter substrates scatter the excitation
20      protons and lower the signal-to-noise ratio for PIXE.  X-ray analysis methods, such as  PIXE
21      and XRF, require particle size diameter corrections (for low atomic number targets)
22      associated with a spherical particle of a given diameter (typically particles with aerodynamic
23      diameters >2.5 ^m), and compositions typical in ambient aerosol studies.  These analyses
24      also require correction for  sample loadings that reflect the passage of x-rays through a
25      uniform deposit layer.  Procedures for instrument calibration,  spectrum process,  and quality
26      assurance are similar to those documented in Section 4.3.1.2  for XRF.
27           PIXE analysis can provide one of the widest range of elements in a single analysis,
28      since x-ray results require two or three  separate anodes.  Attempts to improve sensitivity of
29      PIXE analysis may result in damage to  Teflon-membrane filters, however. Recent
30      developments (Malm et al., 1994) using PIXE analysis at moderate  sensitivity plus single
        April 1995                                4-78       DRAFT-DO NOT QUOTE OR CITE

-------
                                                           OJ
                                                           53
                                                           f
                                                           •§
                                                           C5
                                                           §
                                                           •g
                                                           C/l
                                                           SJ
                                                           £
April 1995
4-79
DRAFT-DO NOT QUOTE OR CITE

-------
 1     anode XRF analysis at high sensitivity for transition/heavy metals have achieved the
 2     minimum detectable limits of less than 0.01 ng/m3.  With the addition of hydrogen analysis
 3     (a surrogate for organic matter), almost all gravimetric mass concentrations can be explained
 4     (Cahill, 1987).
 5           XRF and PIXE are the most commonly used elemental analysis methods owing to its
 6     nondestructive multi-element capabilities, relatively low cost, high detection limits, and
 7     preservation of the  filter for additional analyses.  XRF sometimes needs to be supplemented
 8     with INAA when extremely low detection limits are needed, but the high cost of INAA
 9     precludes this method from being applied to large numbers of samples.  AAS is  a good
10     alternative for water-soluble species, especially for low atomic number .  ICP/AES analysis
11     is a viable alternative, but it is less desirable because of the sample extraction elements such
12     as sodium and magnesium, but it requires  large dilution factors to measure many different
13     elements expense and the destruction of the filter.
14
15     4.3.2.3   Instrumental Neutron Activation Analysis of Trace Elements
16           Instrumental neutron activation analysis (INAA) (Dams et al., 1970; Zoller and
17     Gordon, 1970; Olmez, 1989; Ondov and  Divita, 1990), irradiated the thin membrane  filter
18     sample  in the  core of a nuclear  reactor for periods  ranging from a few minutes to several
19     hours.  Bombardment of the sample with neutrons induces a nuclear reaction of the stable
20     isotopes in the sample.  The energies of the gamma rays emitted by the decay of this induced
21     radioactivity are used to identify them, and therefore, their parents. With the use of
22     prepared elemental  standards, the amount of parent element in the  sample can be determined
23     since the intensity of these gamma rays are proportional to their number.
24           The gamma-ray  spectra  of radioactive species are usually collected with a high
25     resolution germanium detector utilizing commercially available amplifiers and multi-channel
26     analyzers.  Typical  detector efficiencies range from 10 to 40% relative to a 3 x  3 in. sodium
27     iodide detector.  Detector system resolution, measured as the full-width at half-maximum for
28     Table 4-4, the 1,332 KeV gamma-ray peak of 60Co, should be less than 2.3 KeV in order to
29     provide adequate resolution between isotopes of neighboring energies.
       April 1995                               4-80      DRAFT-DO NOT QUOTE OR CITE

-------
             TABLE 4-4.  INAA COUNTING SCHEME AND ELEMENTS MEASURED
        Counting
        Period
Irradiation
Time
Cooling
Time
Counting
Time
Elements Measured
        Short-Lived 1    10 min
        Short-Lived 2
        Long-Lived 1   4-6 h
        Long-Lived 2
                5 min
                20 min
                30 days
             5 min
             20 min
                3-4 days     6-8 h
             12-24 h
              Mg, Al, S, Ca, Ti, V, Cu
              Na, Mg, Cl, K, Ca, Mn,
              Zn, Ga, Br, Sr, In, I, Ba
              Na, K, Ga, As, Br, Mo,
              Cd, Sb, La, Nd, Sm, Yb,
              Lu, W, Au, U
              Sc, Cr, Fe, Co, Zn, Se,
              Sr, Ag, Sb, Cs, Ba, Ce,
              Nd, Eu, Gd, Tb, Lu, Hf,
              Ta, Th
 1          In order to obtain a full suite of elemental analysis results (often over 40 elements),
 2     multiple counting periods and irradiations are performed on the same sample (e.g., two
 3     irradiations would produce elements separated into short- and long-lived decay products).
 4     An example of the elements determined from multiple irradiations and counting periods, and
 5     the irradiation, cooling, and counting  times used for ambient particulate samples collected on
 6     Teflon-membrane filter material is summarized in Table 4-4 (Divita, 1993).   These
 7     irradiations were performed at the 20-MW NIST Research  Reactor operated at 15-MW
 8     (neutron flux of 7.7 x  1013 and 2.7 x 1013 neutron/cm2 x s). Typical gamma-ray spectra
 9     resulting from the counting scheme described in Table 4-4 are shown in Figures 4-24
10     and 4-25.
11          The power of INAA is that it is  not generally subject  to interferences like XRF or
12     PIXE due to a much better ratio of gamma ray peak widths to total spectral  width, by a
13     factor of about 20. INAA does  not quantify some of the abundant species in ambient
14     particulate matter such as silicon, nickel, tin, cadmin, mercury, and lead.  While INAA  is
15     technically nondestructive, sample preparation involves folding the samples tightly and
16     sealing it in plastic, and the irradiation process makes the filter membrane brittle and
17     radioactive.  These factors limit  the use of the sample for subsequent analyses by other
18     methods.   The technique also suffers from the fact that a nuclear reactor is usually used  as a
       April 1995
                       4-81
                  DRAFT-DO NOT QUOTE OR CITE

-------
3113 >IO 31OAO ION
5661
Figure 4-24. Typical Gamma-Ray Spectra Observed for Long Counts.
Count Count
Thousands Thousands
— * ~* IN) 10 CD — * -* r\i FO tj
oenoenoeno ocnocnoenc
§
o
O
03
CD.J-*
c 8
1
i
O
IN)
O
O
O
PC
8
o
ro
en
o
0°
13
cp_co
i§
o-
CD
co
en
o
o
o
fe 	
r-

•

Q
(Q
ro
>
Count
Thousands
[
	

(O
ro
CD
en
8
o
1
CD -*
= 8
1°
1
_*
Oi
8
ro
8
IN)
en
0
0°
Q)
CD.CO
I8
CT
CD
CO
0
1 1 1 f 1
I
i.
" cT
_  ->j

/_
I

—
^^__M "7
(O
i
CD

-------
aio HO aiono ION
             £661 I
                               Count
                              Thousands
                            
3
                                            o

                                            ro
                                                  O

                                                  0>
                                                 -c
                                                  3
                                                    CO

                               Count
                               Thousands
                                                                Count

                   en
                   00
                      r
                            i    i    i   r
                                            03
                                                   Oi
                                                 O
                                                   O)
                                                 CT
                                                 CD
               co

               o
               3.


                1

               00

-------
 1     source of neutrons. However, since the advent of high-resolution gamma-ray detectors,
 2     individual samples can be analyzed for numerous elements simultaneously, most at
 3     remarkably trace levels without the need for chemical separation. This greatly diminishes the
 4     danger of contamination due to excessive sample handling and introduction of chemical
 5     reagents used for separation procedures.
 6
 7     4.3.2.4  Microscopy Analysis of Particle Size, Shape, and Composition
 8          Morphological and chemical features of particles can be used to identify the sources and
 9     transport mechanism of airborne particles.  The chemical analysis of individual particles
10     allows the attribution  of specific pollution sources more straightforward while the abundance
11     of a specific group is  a representative  of the source strength. Both light (optical) and
12     scanning electron microscopy have been applied in environmental studies to examine the
13     single particles (e.g.,  Casuccio et al.,  1983; Bruynseels et al., 1988; Van Borm and Adams,
14     1988; Javitz and Watson, 1989; Van Borm  et al., 1989; CorniUe et al., 1990; Hopke and
15     Casuccio, 1991; Hoffer et al., 1991; Cheng et al., 1992; Turpin et al., 1993; 1994; 1995).
16          Light microscopy has been used  for providing particle size information regarding the
17     morphology of microscopic features (Crutcher, 1982). The practical resolution of optical
18     microscopes is limited by the wavelengths associated with light of the visible spectrum.
19     When features  of interest occur in micron and submicron size ranges, detailed resolution
20     cannot be obtained.  The practical resolution of light microscopy is typically 1 to 2 /mi
21     (Meyer-Arendt, 1972).
22          The use of accelerated electrons  in electron microscopy allows for the formation of
23     magnified images  and an increased depth of field, and provides  the resolution of a few
24     angstroms (10"4 ^m).   Electron microscopy has now evolved to  include:  1) the transmission
25     electron microscope (TEM); 2) the scanning electron microscope (SEM), and; 3) the
26     scanning transmission electron microscope (STEM) (Hearle et al. 1972; Lee et al., 1979;
27     Scott and Chatfield, 1979; Lee and Fisher, 1980; Lee and Kelly, 1980; Lee et al., 1981;
28     Johnson et al., 1981;  Mclntyre and  Johnson,  1982; Casuccio et  al., 1983; Wernisch, 1985,
29     1986; Kim et al.,  1987, 1988; Dzubay and Mamane, 1989; Henderson et al., 1989;
30     Schamber, 1993).
        April 1995                                4-84      DRAFT-DO NOT QUOTE OR CITE

-------
  *
 1          The SEM and STEM use accelerated electrons to strike the sample.  As the electron
 2     beam strikes the samples, various signals (e.g., secondary, backscattered, and Anger
 3     electrons, characteristic x-rays, photons, and cathodoluminescence) are generated.  These
 4     signals can be collected to provide highly  detailed information on a point-by-point basis.  The
 5     secondary electron signal yields a sample  image with three-dimensional prospective, high
 6     depth of field, and illuminated appearance.  Back scattered electron images are used to
 7     separate phases containing elements of different atomic number.
 8          The information obtained from light and scanning microscopy analyses are  usually
 9     considered to be qualitative, due to the  limited number of particles counted. To  achieve a
10     quantitative analysis, a sufficient number of particles must be properly sized and identified by
11     morphology and/or chemistry  to represent the entire sample.  The selection of filter media,
12     optimal particle loadings, and  sample handling methods are also of importance.   In this
13     manner, the microscopic characteristics can be directly and reliably related to the bulk or
14     macroscopic properties of the  sample (Casuccio et al., 1983).
15          Microscopic analysis requires a high degree of skill  and extensive quality assurance to
16     provide quantitative information. The techniques is complex and expensive when quantitative
17     analysis is required.  The evolution of computer technology has allowed for quantitative
18     analysis of particle samples of an entire population of features. With advanced pattern
19     recognition methods, data from individual particle features can be sorted and summarized by
20     size and composition, permitting improved quantitative source apportionment (Bruynseels
21     et al.,  1988; Hopke and  Casuccio, 1991).  Casuccio et al. (1983) summarized the pros and
22     cons of automatic scanning electron microscopy.
23          Recent development of the SEM/XRF allows analysis of elemental compositions and
24     morphological information on small quantities of material (Bruynseels et al., 1988).  Coupled
25     with statistical data analysis, computer controlled scanning electron microscopy shows great
26     promise for identifying and quantifying complex pollution sources in the field of receptor
27     modeling source apportionment (e.g., Griffin and Goldberg, 1979; Janocko et al., 1982;
28     Johnson et al, 1982; Massart and Kaufman, 1983; Hopke, 1985; Derde et al.,  1987, Saucy
29     et al.,  1987; Mamane, 1988a,  1988b; Cheng and Hopke,  1989; Dzubay and Mamane, 1989).
        April 1995                                4-85       DRAFT-DO NOT QUOTE OR CITE

-------
 1     4.3.3   Wet Chemical Analysis
 2           Aerosol  ions refer to chemical compounds that are soluble in water.  The water-soluble
 3     portion of suspended particles associates itself with liquid water in the atmosphere when
 4     relative humidity increases, thereby changing the light scattering properties of these particles.
 5     Different emissions sources may also be distinguished by their soluble and non-soluble
 6     fractions. Gaseous precursors can also be converted to their ionic counterparts when they
 7     interact with chemicals impregnated on the filter material.
 8           Several simple ions, such as soluble  sodium,  magnesium, potassium, and calcium are
 9     best quantified by atomic absorption spectrometry (AAS) as described above. In practice,
10     AAS has been very useful for measuring water-soluble potassium and sodium, which are
11     important in apportioning sources of vegetative burning and sea salt, respectively.
12     Polyatomic  ions such as  sulfate, nitrate, ammonium, and phosphate must be quantified by
13     other methods such as ion chromatography (1C) and automated colorimetry (AC).  Simple
14     ions, such as chloride, chromium III, and chromium IV, may also be measured by these
15     methods along with the polyatomic ions.
16           All ion analysis methods require filters to be extracted in DDW and then filtered to
17     remove the  insoluble residue.  The extraction volume needs to be as small as possible, lest
18     the solution become too  dilute to detect the desired constituents.   Each square centimeter of
19     filter should be extracted in no more than 2 ml of solvent for typical sampler flow rates of
20     20 to 30 L/min and sample durations of 24 h. This often results  in no more than 20 ml of
21     extract that  can be submitted to the different analytical methods, thereby giving preference to
22     those methods which require only  a small sample volume.  Sufficient sample deposit must be
23     acquired to  account for the dilution volume required by each method.
24           When other analyses are to be performed on the same filter, the filter must first be
25     sectioned using a precision positioning jig attached to a paper cutter.  For rectangular filters
26     (typically 20.32 cm by 25.40 cm), a 2.0 cm by 20.32 cm wide strip is cut from the center
27     two-thirds of  the filter.   Circular filters of 25-, 37-, and 47-mm diameters are usually cut in
28     half for these analyses,  so the results need to be multiplied by two to obtain the deposit on
29     the entire filter.  Filter materials that can  be easily sectioned without damage to the filter or
30     the deposit  must be chosen for these analyses.
31

        April 1995                                4-86       DRAFT-DO NOT QUOTE OR CITE

-------
 i»    4.3.3.1   Ion Chromatographic Analysis for Chloride, Nitrate, and Sulfate
 2          Ion Chromatography (1C) can be used for both anions (fluoride [F~] chloride [Cl~],
 3     nitrite  [NO^ bromide [Br~], nitrate [NO^], phosphate [PO43], sulfate [SO4=]) and cations
 4     (soluble potassium [K+], ammonium [NH4+], soluble sodium [Na+]) with separate columns.
 5     Applied to aerosol samples, the anions are most commonly analyzed by 1C with the cations
 6     being analyzed by a combination of atomic absorption spectrophotometry (AAS) and
 7     automated colorimetry (AC) (Chow and Watson, 1994).  In 1C (Small et al., 1975; Mulik
 8     et al.,  1976, 1977; Butler et al., 1978; Mueller et al., 1978; Rich et al., 1978; Small, 1978),
 9     the sample extract passes through an ion-exchange column that separates the ions in time for
10     individual quantification, usually by a electroconductivity detector.  Figure 4-26 shows a
11     schematic representation of the 1C system.  Prior to detection, the column effluent enters a
12     suppressor column where the chemical composition of the eluent is altered, resulting in a
13     lower background conductivity.  The ions are identified by their elution/retention times and
14     are quantified by the conductivity peak area or peak height.  1C is especially desirable for
15     particle samples because it provides results for several ions with a single analysis and it uses
16     a small portion of the filter extract with low detection limits.  Water-soluble chloride (Cl~),
17     nitrate (NOp, and sulfate (SO4=) are the most commonly measured anions in aerosol
18     samples.  Figure 4-27 shows  an example of an 1C anion chromatogram.  1C analyses can be
19     automated by interfacing to an automatic sampler that can conduct unattended analysis of as
20     many as 400 samples (Tejada et al.,  1978).
21          In 1C, approximately 2 to 3  ml of the filter extract are injected into the 1C system.  The
22     resulting peaks are integrated and the peak integrals are converted to concentrations using
23     calibration curves derived from standard solutions.  For instance, the Dionex system
24     (Sunnyvale, CA) for the analysis of Cl", NO^, NO~3 PO4=, and SO4  contains a guard column
25     (AG4A column, Cat. No. #37042) and an anion separator column (AS4A column, Cat. No.
26     #37041) with a strong-basic anion-exchange resin, and an anion micro-membrane suppressor
27     column (250 x 6 mm ID) with a strong-acid ion-exchange resin.  The 4  x 250 mm
28     analytical column is composed of 16 micron polystyrene/divinylbenzene  substrate
29     agglomerated with anion exchange latex that has been completely aminated. The 0.5%
30     crosslinked  latex particles have a diameter of approximately 0.175 ^im and carry the ion
31     exchange  sites.  The ion exchange capacity of the 4  X 250 mm analytical column is

       April 1995                               4-87       DRAFT-DO NOT QUOTE OR CITE

-------
           Delivery Module—
   Chromatography Module —
           Detector Module —
                                                   Eluent
                                                 Reservoir
                                                   Pump
                                                  Sample
                                                  Injector
                                                   Guard
                                                  Column
                                                 Separator
                                                  Column
                                                 Suppressor
                                                   Device
                                                Conductivity
                                                    Cell
                                       \Waste J
Figure 4-26.  Schematic Representation of an Ion Chromatography (1C) System.
April 1995
4-i
DRAFT-DO NOT QUOTE OR CITE

-------
18,000-
15,500-
13,000-
10,500-
8,000-
5,500-
3,000-
500-
-2,000-
0.






Fluoride
I Chloride
I I

-^

I

30
Nitrite
lj

Nitrate
\ A Dk u . Sulfate
\ \ Phosphate /x
\ / \_ /\ / \




5.00 10.00
Minutes
       Figure 4-27. Example of an Ion Chromatogram Showing the Separation of Fluoride,
                    Chloride, Nitrite, Nitrate, Phosphate, and Sulfate Ions.
 1     20 /ueq/column.  The column is stable between pH 0 and 14.  The anion eluent consists of
 2     sodium carbonate (Na2CO3) and sodium bicarbonate (NaHCO3) prepared in DDW.  The
 3     DDW  is verified to have a conductivity of less than 1.8  x 10"5 ohm"1 cm"1 prior to
 4     preparation of the eluent. For quantitative determinations, the 1C is operated at a flow rate
 5     of 2.0  L/min.  The system can also analyze fluoride with an eluent concentration of 1.8 M
 6     Na2CO3/1.7MNaHCO3.
 7          The primary standard solution is prepared annually and stored in a refrigerator.  It is
 8     prepared from the reagent grade sodium salts (e.g., NaF, NaCl, NaNO2, NaBr, NaNO3,
 9     Na2HPO, and  Na2SO4.  These anhydrous  salts are dried in an oven at 105 °C for 1 h and
10     then cooled to room temperature in a dessicator.  They are weighed to the nearest 0.10 mg
11     on a routinely  calibrated analytical balance under controlled temperature ( — 20° C) and
12     relative humidity (± 30%) conditions. These  salts are diluted in precise volumes of DDW.
13     Calibration standards are prepared at  least once within each month by diluting the primary
14     standard to concentrations covering the range of concentrations expected in the filter extracts.
       April 1995
4-89
DRAFT-DO NOT QUOTE OR CITE

-------
 1     The normal calibration concentrations prepared are 0.1, 0.2, 0.5,  1.0, and 2.0 ptg/ml for
 2     each of the analysis species.
 3          Several  independent quality assurance (QA) standards should be used to check the
 4     calibration curve.  The standards that are traceable to NIST simulated rainwater standards
 5     are:  Environmental Resource Associates (ERA, Arvada, CA) custom standards  containing
 6     the anions measured at a concentration of 100 /xg/ml, ERA Waste Water Nutrient Standard,
 7     ERA Waste Water Mineral Standard, and Alltech individual standards at 200 /ig/ml. The
 8     QA standards are diluted in DDW to concentrations that are within the range of the
 9     calibration curve.
10          Calibration curves are performed weekly.  Chemical compounds are identified by
11     matching the  retention time of each peak in the unknown sample with the retention times of
12     peaks in the chromatograms of the standards.  The QA standards are analyzed at the
13     beginning of  each sample run to check calibrations.  A DDW blank is analyzed  after every
14     20 samples and a calibrations standard is analyzed  after every 10 samples. These quality
15     control (QC)  checks verify the baseline and calibration respectively.
16
17     4.3.3.2   Automated Colorimetric  Analysis for Ammonium, Nitrate, and Sulfate
18          Automated Colorimetry  (AC) applies different colorimetric analyses to small sample
19     volumes with automatic sample throughput. The most common ions measured are
20     ammonium, chloride,  nitrate, and sulfate (Butler et al.,  1978; Mueller et al., 1978; Fung
21     et al.,  1979;  Pyen and Fishman,  1979). Since 1C provides multi-species analysis for the
22     anions, ammonium is  most commonly measured by AC.
23          The  AC system  is illustrated schematically in Figure 4-28.  The heart of the automated
24     colorimetric system is a peristaltic pump,  which introduces air bubbles into the sample
25     stream at  known intervals.  These bubbles separate samples in the continuous stream. Each
26     sample is  mixed with  reagents and subjected to appropriate reaction periods before
27     submission to a colorimeter.  The ion being measured usually reacts to form a colored liquid.
28     The liquid absorbance is related to the amount of the ion in the sample by Beer's Law. This
29     absorbance is measured by a photomultiplier tube through an interference filter  specific to the
30     species being measured.
        April 1995                               4-90       DRAFT-DO NOT QUOTE OR CITE

-------
18,000-
15,500-
13,000-
10,500-
8,000-
5,500-
3,000-
500-
-2,000-
0.




Fluoride
Chloride
I 1

-^
Nitrite
111

)0
Nitrate
\ |\ Phosphate /^e
\ / V s\ / \




5.00 10.00
Minutes
       Figure 4-28. Schematic of a Typical Automated Colorimetric (AC) System.
 1          The standard AC technique can analyze »60 samples per hour per channel, with
 2     minimal operator attention and relatively low maintenance and material costs.  Several
 3     channels can be set up to simultaneously analyze several ions.  The methylthymol-blue
 4     (MTB) method is applied to analyze sulfate.  The reaction of sulfate with MTB-barium
 5     complex results in free ligand, which is measured colorimetrically at 460 nm.  Nitrate is
 6     reduced to nitrite that reacts with sulfanilamide to form a diazo compound.  This compound
 7     is then reacted to an azo dye for colorimetric determination at 520 nm. Ammonium is
 8     measured with the indophenol method.  The sample is mixed sequentially with potassium
 9     sodium tartrate, sodium phenolate, sodium hypochlorite, sodium hydroxide,  and sodium
10     nitroprusside. The reaction results in a blue-colored solution with an absorbance measured at
11     630 nm.
12          The system determines carry-over by analysis of a low concentration standard following
13     a high concentration.  The percent carry-over is then automatically calculated and can be
14     applied to the samples analyzed during the run.
15          Formaldehyde has been found to interfere with ammonium measurements when present
16     in an amount exceeding 20% of the ammonium content, and hydrogen sulfide interferes in
       April 1995
4-91
DRAFT-DO NOT QUOTE OR CITE

-------
 1     concentrations exceeding 1 mg/ml.  Nitrate and sulfate are also potential interferents when
 2     present at levels exceeding 100 times the ammonium concentration.  These levels are rarely
 3     exceeded in ambient samples.  The precipitation of hydroxides of heavy metals such as
 4     magnesium and calcium is prevented by the addition of disodium ethylenediamine-tetracetate
 5     (EDTA) to the sample stream (Chow et al., 1980; Chow, 1981-).  It was learned in the
 6     Sulfate Regional Experiment (SURE) (Mueller et al.,  1983) that the auto-sampler should be
 7     enclosed in an atmosphere that is purged of ammonia by bubbling air through a phosphoric
 8     acid solution.
 9          The automated colorimetric system requires a periodic standard calibration with the
10     daily prepared reagents flowing through the system. Lower quantifiable limits of AC for
11     sulfate and nitrate are an order of magnitude higher than those obtained with 1C.
12          Intercomparison studies between AC and 1C have been conducted by Butler et al.
13     (1978); Mueller  et al. (1978); Fung et al. (1979); and Pyen and Fishman (1979).  Butler
14     et al. (1978) found excellent agreement between sulfate and nitrate measurements by AC and
15     1C.  The accuracy of both methods is within the experimental errors, with higher blank
16     values observed  for AC techniques.  Comparable results  were also obtained between the two
17     methods by  Fung et al. (1979). The choice between the  two methods  for sample analysis is
18     dictated by sensitivity, scheduling, and  cost constraints.
19          Two milliliters of extract in  sample vials are placed in an autosampler that is controlled
20     by a computer. Five standard concentrations (e.g.,  (NH4)2SO4, Na2SO4, NaNO3) are
21     prepared from American Chemical Society reagent-grade chemicals  following the same
22     procedure as that for 1C standards.  Each set of samples  consists of two DDW blanks to
23     establish a baseline, five calibration standards and a blank, then sets of ten samples followed
24     by analysis of one of the standards and a replicate from a previous batch.  The computer
25     control allows additional analysis of any filter extract to be repeated without the necessity of
26     loading the extract into more than one vial.
27
28     4.3.3.3  Atomic Absorption Spectrophotometric (AAS) and Inductive Coupled Plasma
29              Atomic Emission Spectro (ICP/AES) Photometry Analyses for Trace Elements
30          In AAS (Ranweiler and Moyers, 1974; Fernandez,  1989), the sample is first extracted
31     in a strong solvent to dissolve the solid material; the filter or a portion thereof is also
32     dissolved during this process.  A few milliliters of this extract are introduced into a flame
       April 1995                               4.92       DRAFT-DO NOT QUOTE OR CITE

-------
 1      where the elements are vaporized.  Most elements absorb light at certain wavelengths in the
 2      visible spectrum, and a light beam with wavelengths specific to the elements being measured
 3      is directed through the flame to be detected by a monochrometer.  The light absorbed by the
 4      flame containing the extract is compared with the absorption from known standards to
 5      quantify the elemental concentrations.  AAS requires an individual analysis for each element,
 6      and a large filter or several filters are needed to obtain concentrations for a large number of
 7      the elements  specified in Table 4.3.1.  AAS is a useful complement to other methods, such
 8      as XRF  and PIXE, for species such as beryllium, sodium, and magnesium that are not well-
 9      quantified by XRF and PIXE.  Airborne particles are chemically complex and do not
10      dissolve easily into complete solution, regardless of the strength of the solvent.  There is
11      always a possibility that insoluble residues are left behind and soluble species may
12      co-precipitate on them or on container walls.
13           In  ICP/AES  (Fassel and Kniseley, 1974; McQuaker et al., 1979; Lynch et al.,  1980;
14      Harman, 1989), the dissolved sample is introduced into an atmosphere of argon gas seeded
15      with free electrons induced by high voltage from a surrounding Tesla coil.  The high
16      temperatures in  the induced plasma raise valence electrons above their normally stable states.
17      When these electrons return to their stable states, a photon of light is emitted which is unique
18      to the element which was excited.  This light is detected at specified wavelengths to identify
19      the elements  in the sample.  ICP/AES acquires a large number of elemental concentrations
20      using small sample volumes with acceptable detection limits for atmospheric samples.
21      As with  AAS, this method requires complete extraction and destruction of the sample.
22
23      4.3.4   Organic Analysis
24      4.3.4.1   Analysis of Organic Compounds
25           Organic compounds comprise a major portion of airborne particles in the atmosphere,
26      thus contributing to visibility degradation, and affecting the properties of clouds into  which
27      these particles are scavenged.  Specific groups of organic compounds (e.g., polycyclic
28      aromatic hydrocarbons, PAHs) have also been implicated in human health effects.  However,
29      due to the very complex composition of the organic fraction of atmospheric aerosols, the
30      detailed composition and atmospheric distributions of organic aerosol constituents are still not
31      well understood.  .

        April 1995                               4.93       DRAFT-DO NOT QUOTE OR CITE

-------
 1          Sampling techniques for atmospheric participate matter have been extensively
 2     investigated, resulting in the development of collection methods for ambient aerosol particles
 3     in a wide range of particle sizes.  Particles are most frequently collected on glass or quartz-
 4     fiber filters that have been specially treated to achieve low "carbon blanks".  Ambient
 5     organic paniculate matter has also been collected on a variety of particle sizing devices, such
 6     as low pressure impactors and Micro Oriface Uniforms Deposit Impactors("MOUDI").  Very
 7     recently, diffusion based samplers have been used as well (i.e., Eatough et al., 1994; Tang
 8     et al.,  1994). However, the task of sampling organic compounds in airborne particles is
 9     complicated by the fact that many of these compounds have equilibrium vapor pressures
10     (gaseous concentrations) that are considerably larger than their normal ambient
11     concentrations. This implies a temperature- and concentration-dependent distribution of such
12     organics between paniculate and vapor phases.  It also suggests that artifacts may occur due
13     to volatilization during the sampling process (Coutant et al., 1988).  Such volatilization
14     would cause the under-estimation of the particle-phase concentrations of organics.
15     Conversely, the adsorption of gaseous substances on deposited particles,  or on the filter
16     material itself,  a process driven by the lowered  vapor pressure over the sorbed material,
17     would lead to over-estimation of the particle-phase fraction (Bidleman et al., 1986;  Ligocki
18     and Pankow, 1989; McDow and Huntzicker, 1990).  In addition, several studies have
19     suggested that chemical degradation of some organics may occur during  the sampling
20     procedure  (Lindskog et al., 1985; Arey et al.,  1988; Parmar and Grosjean,  1990).
21          The partitioning of semi-volatile organic compounds (SOC) between vapor and particle
22     phases has received  much attention recently (Ligocki and Pankow,  1989; Gotham and
23     Bidleman, 1992; Lane et al., 1992; Kaupp and  Umlauf, 1992; Pankow,  1992; Turpin et al.,
24     1993).  Most estimates  of partition have relied on high-volume (hi-vol) sampling, using a
25     filter to collect particles followed by a solid adsorbent trap to collect the gaseous portion of
26     SOC (e.g., Kaupp and Umlauf, 1992, Foreman and Bidleman, 1990). It has been shown
27     recently (Kaupp and Umlauf, 1992) that this approach, although not absolutely free from
28     sorption and desorption artifacts,  produces reliable results.  The maximum differences
29     observed between hi-vol filter-solid adsorbent sampling and impactor sampling (the latter
30     believed to be less susceptible to  these sampling artifacts) did not exceed a factor of two.
        April 1995                                4-94      DRAFT-DO NOT QUOTE OR CITE

-------
 1           There is good theoretical and experimental evidence that use of a diffusion denuder
 2      technique significantly improves measurements of vapor-particle phase partitioning (Coutant
 3      et al., 1988, 1989, 1992; Lane et al.,  1988).  However, at the present state of their
 4      technological development, the reliability of denuders for investigation of atmospheric
 5      partitioning of non-polar SOC needs to be improved, as suggested by contradictions in
 6      published field data (e.g.,  Kaupp and Umlauf,  1992). A new, improved sampler has recently
 7      been introduced (Gundel et al., 1992)  which uses a proprietary XAD-4-coated tube for vapor
 8      collection, followed by filter collection of organic aerosol particles and a sorbent bed to
 9      quantitatively retain desorbed (volatilized) organic vapors.  Preliminary results from the use
10      of this device look very promising for direct measurements of the phase distribution of semi-
11      volatile organic aerosol constituents.  Another promising application of denuder technology
12      has been their use to  remove ozone from an air sampling stream before it reaches the filter,
13      reducing the potential for artifact formation on the captured paniculate material during the
14      sampling period (Williams and Grosjean, 1990).
15           Since the organic fraction of airborne particulate matter is typically a complex mixture
16      of hundreds  to thousands of compounds distributed over many organic functional groups, its
17      chemical analysis  is an extremely difficult task (Appel et al., 1977; Simoneit,  1984; Flessel
18      et al., 1991; Hildemann et al., 1991; Li and Kamens, 1993; Rogge et al., 1993a, 1993b,
19      1993c).  Analyses of organics generally begin with solvent extraction of the particulate
20      sample.  A variety of solvents and extraction techniques have been used in the past.  One
21      common method is sequential extraction with increasingly polar solvents, which typically
22      separates the organic material into nonpolar, moderately polar, and polar fractions (Daisey
23      et al., 1982, 1987). This step is usually followed by further  fractionation using open-column
24      liquid chromatography and/or high-performance liquid chromatography (HPLC) in order to
25      obtain several less complicated fractions (e.g.,  Schuetzle and Lewtas, 1986; Atkinson et al.,
26      1988).  These fractions can then be analyzed further with high resolution capillary-column
27      gas chromatography (GC), combined with mass spectrometry (GC/MS), Fourier transform
28      infrared (GC/FTIR/MS) or other selective detectors.
29           Much of the recent work on the identification of nonpolar and semi-polar organics in
30      airborne samples has  used  bioassay-directed chemical analysis (Scheutzle and Lewtas, 1986),
31      and has focused on identification of fractions and compounds that are most likely to be of

        April 1995                                4.95       DRAFT-DO NOT QUOTE OR CITE

-------
 1      significance to human health.  In particular, PAHs and their nitro-derivatives (nitroarenes)
 2      attracted considerable attention due to their mutagenic and, in some cases, carcinogenic
 3      properties.  More than 100 PAHs have been identified in the PM2 5 fraction of ambient
 4      particulate matter (Lee et al.,  1981).  While most of the nitroarenes found in ambient
 5      particles are also present in primary combustion-generated emissions,  some are formed from
 6      their parent PAH in the atmospheric nitration reactions (e.g., Arey et al., 1986;  Zielinska
 7      et al., 1989, Ramdahl et al.,  1986).
 8           Little work has been done to date to chemically characterize the polar fraction in detail,
 9      even though polar material accounts for up to half the mass and mutagenicity of soluble
10      ambient particulate organic matter (Atherholt et al., 1985; Gundel et al., 1994).  Until
11      recently the polar fraction has remained analytically intractable, since very polar and labile
12      species interact with conventional fractionation column packing materials and cannot be
13      recovered  quantitatively.  Recently very polar particulate organic matter has been
14      successfully fractionated using cyanopropyl-bonded silica (Gundel et al.,  1994), with good
15      recovery of mass and mutagenicity (Kado et al., 1991). However, ambient particulate polar
16      organic material  cannot be analyzed with conventional GC/MS because of GC column losses
17      resulting from adsorption, thermal decomposition,  and  chemical interactions.  New analytical
18      techniques, such as HPLC/MS and MS/MS, need to be applied if the chemical constituents
19      of polar particulate organic matter are to be identified and quantified.
20           Most of the recent work on the identification of particulate organic matter has focused
21      on mutagenic  and carcinogenic compounds that are of significance to human health.
22      Relatively  little work has  been done to characterize individual compounds or classes of
23      compounds that are specific to certain sources of organic aerosol. In urban and  rural
24      atmospheres,  as well as in the remote troposphere, organic composition corresponding to
25      fingerprints of plant waxes, resin residues, and long-chain hydrocarbons  from petroleum
26      residues have  been found (e.g., Gagosian et al., 1981;  Simoneit, 1984; Mazurek et al., 1987,
27      1989,  1991; Simoneit et al., 1991; Rogge et al., 1994). However, a variety of smaller,
28      multi-functional compounds characteristic of gas-to-particle conversion have also been
29      observed (e.g., F inlay son-Pitts and Pitts,  1986). These compounds  tend to be present in the
30      polar fraction of ambient  organic aerosol particles, having been formed from atmospheric
        April 1995                               4.96       DRAFT-DO NOT QUOTE OR CITE

-------
  1      chemical reactions of less polar precursors.  Little is currently known about the chemical
  2      composition of this polar fraction due to the serious analytical difficulties mentioned above.
  3
  4      4.3.4.2  Analysis of Organic and Elemental Carbon
  5           Three classes  of carbon are commonly measured in aerosol samples collected on
  6      quartz-fiber filters:  1) organic, volatilized, or non-light absorbing carbon; 2) elemental or
  7      light-absorbing carbon; and 3) carbonate carbon.  Carbonate carbon (i.e., K2CO3, Na2CO3,
  8      MgCO3, CaCO3) can be determined on a separate filter section by measurement of the
  9      carbon dioxide (C02) evolved upon acidification (Johnson et al., 1981).  Though progress
 10      has been made in the  quantification of specific organic chemical compounds in suspended
 11      particles (e.g., Rogge et al.,  1991), sampling and analysis methods have not yet evolved for
 12      use in practical monitoring situations.
 13           Many methods have  been applied to the separation of organic and elemental carbon in
 14      ambient  and source particulate samples (McCarthy and  Moore, 1952; Mueller et al., 1971;
 15      Lin et al.,  1973;  Patterson, 1973; Gordon,  1974; Grosjean,  1975; Smith et al.,  1975; Appel
 16      et al., 1976, 1979;  Kukreja and Bove, 1976; Merz,  1978; Rosen et al., 1978; Dod et al.,
 17      1979; Johnson and Huntzicker, 1979;  Macias et al., 1979; Malissa, 1979; Weiss et al., 1979;
 18      Cadle et al., 1980a, 1980b; Heisler et al., 1980a, 1980b; Johnson et al., 1980,  1981;
 19      Pimenta  and Wood, 1980; Daisey et al., 1981; Mueller et al., 1981; Novakov,  1981, 1982;
20      Cadle and  Groblicki,  1982; Gerber, 1982; Heintzenberg, 1982; Huntzicker et al., 1982;
21      Muhlbaier  and Williams, 1982; Rosen et al.,  1982; Tanner et al., 1982; Stevens et al., 1982;
22      Wolff et al., 1982;  Japar et al., 1984). Comparisons among the results of the majority of
23      these methods show that they yield comparable quantities of total carbon in aerosol samples,
24      but the distinctions between organic and elemental carbon are quite different (Countess,
25      1990; Heringetal., 1990).
26           The definitions of organic and elemental carbon are operational and reflect the method
27      and purpose of measurement.  Elemental carbon  is sometimes termed "soot", "graphitic
28      carbon," or "black carbon."  For studying visibility reduction, light-absorbing carbon is a
29      more useful concept than elemental carbon.  For source apportionment by receptor models,
30      several consistent but  distinct fractions of carbon in both source  and receptor samples are
31      desired, regardless of their light-absorbing or chemical properties.  Differences in ratios of

        April 1995                               4.97      DRAFT-DO NOT QUOTE  OR CITE

-------
 1     the carbon concentrations in these fractions form part of the source profile that distinguishes
 2     the contribution of one source from the contributions of other sources.
 3          Light-absorbing carbon is not entirely constituted by graphitic carbon, since there are
 4     many organic materials that absorb light (e.g., tar, motor oil, asphalt, coffee).  Even the
 5     "graphitic" black carbon in the atmosphere has only a poorly developed graphitic structure
 6     with abundant surface chemical groups.  "Elemental carbon" is a poor but common
 7     description of what is measured.  For example, a substance of three-bond carbon molecules
 8     (e.g., pencil  lead) is black and completely absorbs light, but four-bond carbon in a diamond
 9     is completely transparent and absorbs very little light. Both are pure, elemental carbon.
10          Chow et al. (1993) document several variations of the thermal (T), thermal/optical
11     reflectance (TOR), thermal/optical transmission (TOT), and thermal manganese oxidation
12     (TMO) methods for  organic and elemental carbon. The TOR and TMO methods have been
13     most commonly applied in aerosol studies in the United States.
14          The TOR method of carbon analysis developed by Huntzicker et al. (1982) has been
15     adapted by several laboratories for the quantification of organic and elemental carbon on
16     quartz-fiber filter deposits.  While the principle used by these laboratories  is identical to that
17     of Huntzicker et al.  (1982), the details differ with respect to calibration standards, analysis
18     time, temperature ramping,  and volatilization/combustion temperatures.  In the TOR method
19     (Chow et al., 1993), a filter is submitted to volatilization at temperatures ranging from
20     ambient to 550°C in a pure helium atmosphere, then to combustion at temperatures between
21     550 to 800°C in a 2% oxygen and 98% helium atmosphere with several temperature ramping
22     steps.  The carbon that evolves at each temperature is converted to methane and quantified
23     with a flame ionization detector.  The reflectance from the deposit side of the filter punch is
24     monitored throughout the analysis.  This reflectance usually decreases during volatilization in
25     the helium atmosphere owing to the pyrolysis of organic material.  When oxygen is added,
26     the reflectance increases as the light-absorbing carbon is combusted and removed.  Organic
27     carbon is defined as that which evolves prior to re-attainment of the original reflectance, and
28     elemental carbon is  defined as that which  evolves after the original reflectance has  been re-
29     attained.  By this definition,  "organic carbon" is actually organic carbon that does not absorb
30     light at the wavelength (632.8 nm) used, and "elemental carbon" is light-absorbing carbon
31     (Chow et al., 1993). The TOT method applies the same thermal/optical carbon analysis

       April  1995                               4-98      DRAFT-DO NOT QUOTE OR CITE

-------
  1      method except that transmission instead of reflectance of the filter punch is measured.
  2      Thermal methods apply no optical correction and define elemental carbon as that which
  3      evolves after the oxidizing atmosphere is introduced.
  4           The TMO method (Mueller et al., 1982; Fung, 1990) uses manganese dioxide (MnO2),
  5      present and in contact with the sample throughout the analysis,  as the oxidizing agent,  and
  6      temperature is relied upon to distinguish between organic and elemental carbon.  Carbon
  7      evolving at 525°C is classified as organic carbon, and carbon evolving at 850°C  is classified
  8      as elemental carbon.
  9           Carbon analysis methods require a uniform filter deposit because only a small portion
10      of each filter is submitted to chemical analysis.  The blank filter should be white for light
11      reflection methods, and at least partially transparent for light transmission methods.  The
12      filter must also withstand very high temperatures without melting during combustion.
13           Since all organic matter contains hydrogen as the most common elemental species,
14      analysis of hydrogen by proton elastic scattering analysis (PESA) has been developed by
15      Cahill (1987).  A correction must be made for hydrogen in sulfates and nitrates, but since the
16      analysis is done in a vacuum, water is largely absent. The method has excellent sensitivity
17      which is approximately 20 times better than combustion techniques.   This method requires
18      knowledge of the chemical state of sulfates, nevertheless, reasonable  agreement was found as
19      compared to the combustion techniques.
20
21      4.3.5   Quality Assurance
22           The use  of multiple methods and parallel samplers achieves both optimum performance
23      and quality assurance.  While this has been a part of major research studies since the 1970's,
24      its extension to routine monitoring of aerosols was most extensively introduces in the
25      SCENES  and  IMPROVE visibility programs (Eldred, 1989).  The concept was labeled,
26      "Integral Redundancy,"  and was recently adopted by the United Nation's Global Atmospheric
27      Watch Program.
28           The internal consistency checks applied to the IMPROVE  network are listed as follows:
29           1)   Mass (gravimetric) is compared to the sum of all elements on the Teflon-membrane
30               filter of Channel A (PIXE, PESA, XRF analysis; Internally XRF and PIXE are
31               compared for elements around iron).  This  was pioneered in the SCENES program
32               and  is now the standard practice for many aerosol studies.

        April 1995                               4.99      DRAFT-DO NOT QUOTE OR CITE

-------
 1          2)  Sulfate, by ion chromatography on Channel B's nylon filter, after an acidic vapor
 2              denuder, is compared to sulfur (X3)  from Channel A's Teflon-membrane filter by
 3              PIXE.  Agreement is excellent, except for summer.
 4
 5          3)  Organic matter, by combustion on Channel C's quartz-fiber filter stack, is
 6              compared to  organic matter via PESA analysis of hydrogen on Channel A's Teflon-
 7              membrane filter.  This is an exceptionally sever test due to the nature of organics.
 8              These comparisons are made for every IMPROVE analysis, yielding about 25,000
 9              comparisons  to date  (Malm et al., 1994).

10          These types of data validation checks should be carried out in every PM measurement

11     program to ensure the  accuracy, precision, and validity of the chemical  analysis data.
       April 1995                              4-100     DRAFT-DO NOT QUOTE OR CITE

-------
  1      REFERENCES
  2
  3      Allen, D. T.; Palen, E. J.; Haimov, M. I.; Hering, S. V.; Young, J. R. (1994) Fourier transform infrared
  4             spectroscopy of aerosol collected in a low pressure impactor (LPI/FTIR): method development and field
  5             calibration. Aerosol Sci. Technol.  21: 325-342.
  6
  7      American Conference  of Governmental Industrial Hygienists. (1985) Particle size-selective sampling in the
  8             workplace. Cincinnati, OH: American Conference of Governmental Industrial Hygienists.
  9
10      Appel, B. R.; Hoffer,  E.  M.; Haik, M.; Wall, S. M.; Kothny, E. L. (1977) Characterization of organic
11             paniculate matter. Sacramento, CA: California Air Resources Board; document no. ARB-R-5-682-77-72.
12
13      Appel, B. R.; Hoffer,  E.  M.; Kothny, E. L.; Wall, S. M.; Haik, M.; Knights, R. L. (1979) Analysis of
14             carbonaceous material in southern California atmospheric aerosols. 2.  Environ. Sci. Technol. 13: 98-104.
15
16      Appel, B. R.; Povard, V.; Kothny, E. L. (1988) Loss of nitric acid within inlet devices intended to  exclude
17             coarse particles during atmospheric sampling. Atmos. Environ. 22: 2535-2540.
18
19      Appel, B. R.; Cheng,  W.; Salaymeh, F. (1989) Sampling of carbonaceous particles in the atmosphere—II.
20             Atmos.  Environ. 23: 2167-2175.
21
22      Arey, J.; Zielinska, B.; Atkinson,  R.; Winer, A. M.; Ramdahl, T.; Pitts, J. N., Jr. (1986) The formation of
23             nitro-PAH from the gas-phase reactions of fluoranthene and pyrene with the OH radical in the presence
24             of NOX. Atmos. Environ. 20: 2339-2345.
25
26      Arey, J.; Zielinska, B.; Atkinson,  R.; Winer, A. M. (1988) Formation of nitroarenes during ambient
27             high-volume sampling. Environ. Sci. Technol. 22: 457-462.
28
29      Arinc, F.; Wielopolski, L.; Gardner, R. P. (1977) The linear least-squares analysis of X-ray fluorescence spectra
30             of aerosol samples using pure element library standards and photon excitation. In:  Dzubay, T. G., ed.
31             X-ray fluorescence analysis of environmental samples. Ann Arbor, MI: Ann Arbor Science Publishers;
32             p. 227.
33
34      Arnold, S.; Hague, W.; Pierce, G.; Sheetz, R. (1992) The use of beta gauge monitors for PSI and every day SIP
35             monitoring: an overview of the Denver experience.  In:  Chow, J. C.; Ono, D. M., eds. PM10 standards
36             and non-traditional paniculate source controls: proceedings  of AWMA/EPA international specialty
37             conference. Pittsburgh,  PA: Air & Waste Management Association; pp. 13-23.
38
39      Askne,  C.; Brosset, C.; Fern, M.  (1973) Determination of the proton-donating property of airborne  particles.
40             Goteborg, Sweden; NL publication B157.
41
42      Atherholt, T. B.; McGarrity, T. B.; Louis, J. B.; McGeorge, L. J.; Lioy, P.  J.; Daisey, J. M.;  Greenberg, A.;
43             Darack, F. (1985) In: Waters, M.  D.; Sandhu, S. S.; Lewtas, J.; Claxton, L.; Strauss, G.;
44             Nesnow, S., eds.  Short-term bioassays  in the analysis of complex environmental mixtures IV. New York,
45             NY: Plenum Press; pp.  211-231.
46
47      Atkinson, R.; Arey, J.; Winer,  A.  M.; Zielinska, B. (1988) A survey of ambient concentrations  of selected
48             polycyclic aromatic hydrocarbons (PAH) at various  locations in California [final report].  Sacramento,
49             CA: California Air Resources Board; contract no. A5-185-32.
50
51      Baron, P. A. (1986) Calibration and use of the aerodynamic particle sizer (APS-3300). Aerosol Sci.  Technol.
52             5: 55-67.
53
         April 1995                                    4-101       DRAFT-DO NOT QUOTE OR CITE

-------
 1     Baron, P. A. (1993) Measurement of asbestos and other fibers. In: Willeke, K.; Baron, P. A., eds. Aerosol
 2            measurement principles, techniques and applications. New York, NY: Van Nostrand Reinhold; p. 574.
 3
 4     Baron, P. A.; Willeke, K. (1993) Aerosol fundamentals. In: Willeke, K.; Baron, P.  A., eds.  Aerosol
 5            measurement principles, techniques and applications. New York, NY: Van Nostrand Reinhold; p. 16.
 6
 7     Baron, P. A.; Mazumder, M. K.; Cheng, Y. S. (1993) Direct reading techniques using optical particle detection.
 8            In: Willeke, K.; Baron, P. A., eds. Aerosol measurement principles, techniques and applications. New
 9            York, NY: Van Nostrand Reinhold; p. 381.
10
11     Bartley, D. L.;  Breuer, G. M. (1982) Analysis and optimization of the performance  of the 10 mm cyclone. Am.
12            Ind. Hyg. Assoc. J. 43: 520-528.
13
14     Benner, C. L.;  Eatough, D. J.; Eatough, N. L.; Bhardwaja, P. (1991) Comparison of annular denuder and filter
15            pack collection of HNO3 (g), SO2 (g), and particulate-phase nitrate, nitrite and sulfate in the south west
16            desert. Atmos.  Environ. Part A 25: 1537-1545.
17
18     Bidleman, T. F.; Billings, W. N.; Foreman, W. T. (1986) Vapor-particle partitioning of semivolatile organic
19            compounds: estimates from field collections. Environ. Sci. Technol. 20: 1038-1043.
20
21     Botham,  R.  A.; Hughson, G. W.; Vincent, J. H.; Mark, D. (1991) The development of a test system for
22            investigating the performances of personal aerosol samplers under actual workplace conditions. Am. Ind.
23            Hyg. Assoc. J. 52: 423-427.
24
25     Brauer, M.; Koutrakis, P.; Spengler, J. D. (1989) Personal exposures to acidic aerosols and gases. Environ. Sci.
26            Technol. 23: 1408-1412.
27
28     Brockmann, J. E.;  Yamano, N.; Lucero,  D. (1988) Calibration of the aerodynamic particle sizer (AOS-3310)
29            with polystyrene latex monodisperse spheres and oleic acid monodisperse particles. Aerosol Sci.  Technol.
30            8: 279-281.
31
32     Broddin, G.; Cautreels, W.; Van Cauwenberghe, K. (1980) On the aliphatic and polyaromatic hydrocarbon levels
33            in urban and background aerosols from Belgium and the Netherlands. Atmos. Environ. 14: 895-910.
34
35     Brosset, C.; Perm, M.  (1978) Man-made airborne acidity and its determination. Atmos. Environ. 12: 909-916.
36
37     Brown, R. C.; Wake, D.; Thorpe, A.; Hemingway, M. A.; Roff, M. W. (1994) Preliminary assessment of a
38            device for passive sampling  of airborne paniculate.  Ann.  Occup. Hyg. 38: 303-318.
39
40     Bruynseels,  F.; Storms, H.; Van Grieken, R.  (1988) Atmos. Environ. 22: 2593-2602.
41
42     Buckley, T. J.;  Waldman, J. M.; Freeman, N. C. G.; Lioy, P. J.; Marple, V. A.; Turner, W. A. (1991)
43            Calibration, intersampler comparison, and field application of a new PM-10  personal  air-sampling
44            impactor. Aerosol Sci. Technol.  14: 380-387.
45
46     Buettner, H. (1990) Measurement of the size distribution of fine nonspherical particles with a light-scattering
47            particle counter. Aerosol Sci. Technol. 12: 413-421.
48
49     Butler, F. E.; Jungers, R. H.; Porter, L. F.; Riley, A. E.;  Toth, F. J. (1978) Analysis of air particulates by ion
50            chromatography: comparison with accepted methods. In:  Sawicki, E.; Mulik, J. D.; Wittgenstein, E.,
51            eds. Ion chromatographic analysis of environmental pollutants. Ann Arbor, MI: Ann  Arbor Science
52            Publishers, Inc.; pp. 65-76.
53
         April 1995                                    4-102       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Cadle, S. H.; Groblicki, P. J. (1982) An evaluation of methods for the determination of organic and elemental
  2            carbon in paniculate samples. In: Wolff, G. T.; Klimisch, R. L., eds. Paniculate carbon: atmospheric
  3            life cycles. New York, NY: Plenum Press; pp. 89-109.
  4
  5     Cadle, S. H.; Groblicki, P. J.; Stroup, D. P. (1980) An automated carbon analyzer for paniculate samples. Anal.
  6            Chem. 52: 2201-2206.
  7
  8     Cadle, S. H.; Nebel, G. J.; Williams, R. L. (1980) Measurements of unregulated emissions from General
  9            Motors' light duty vehicles. Soc. Automotive Engineers Trans. 87:  2381-2401.
10
11     Cadle, S. H.; Groblicki, P. J.; Mulawa, P. A. (1983) Problems in the sampling and analysis of carbon particles.
12            Atmos. Environ. 17: 593-600.
13
14     Cahill, T. A. (1975) Ion-excited X-ray analysis of environmental samples. In: Ziegler, J., ed. New uses for ion
15            accelerators. New York, NY: Plenum Press.
16
17     Cahill, T. A. (1980) Proton microbes and particle-induced X-ray analytical systems. Ann. Rev. Nucl, Particle
18            Sci. 30: 211-252.
19
20     Cahill, T. A. (1990) Particle-induced X-ray emission. In: Mills, K., ed. Metals handbook. 9th ed. American
21            Society for metals; pp. 102-108.
22
23     Cahill, T. A.; Wakabayashi, P. (1993) Compositional analysis of size-segregated aerosol samples.  In:
24            Newman, L., ed. Measurement challenges in atmospheric chemistry. Washington, DC: American
25            Chemical Society; pp. 211 -228.
26
27     Cahill, T. A.; Eldred, R. A.; Kusko, B. H.; Feeney, P. J.; Malm,  W. C. (1987) Concentrations of natural
28            hydrocarbon particles at National Park Service  sites as derived from mass/hydrogen/sulfur correlations.
29            In: Bhardwaja, P.  S., ed. Visibility protection: research and policy  aspects. Pittsburgh, PA: Air Pollution
30            Control Association; pp. 407-417.
31
32     Cahill, T. A.; Eldred, R. A.; Motallebi, N.; Malm, W. C. (1989) Indirect  measurement of hydrocarbon aerosols
33            across the United States by nonsulfate hydrogen-remaining gravimetric mass correlations. Aerosol Sci.
34            Technol.  10: 421-429.
35
36     Cahill, T. A.; Eldred, R. A.; Feeney, P. J.; Beveridge, P. J.; Wilkinson, L. K.  (1990) The stacked  filter unit
37            revisited. In: Mathai, C. V., ed. Visibility and fine particles: transactions. Pittsburgh, PA: Air & Waste
38            Management Association; pp. 213-218.
39
40     Cahill, T,; Gill, T.; Gillette, D.; Gearhart, E.; Reid, J.; Yau, M. (1994) Generation, characterization and
41            transport of Owens (dry) Lake dust. Sacramento, CA: California Environmental Protection Agency, Air
42            Resources Board; contract report no. A132105.
43
44     Casuccio, G. S.; Janocko, P. B.; Lee, R. J.;  Kelly, J.  F.; Dattner,  S. L.; Mgebroff, J. S.  (1983) The use of
45            computer controlled scanning electron microscopy  in environmental studies. J.  Air Pollut. Control Assoc.
46            33: 937-943.
47
48     Charlson, R. J.; Ahlquist, N. C.; Horvath, H. (1968) On the generality of correlation  of atmospheric aerosol
49            mass concentration and light scatter. Atmos. Environ. 2:  455-464.
50
51     Chen, B. T.; Cheng, Y. S.; Yeh, H. C.  (1985) Performance of a TSI aerodynamic particle sizer. Aerosol Sci.
52            Technol. 4: 89-97.
53
         April 1995                                    4403       DRAFT-DO NOT QUOTE OR CITE

-------
 1     Code of Federal Regulations. (1979) National primary and secondary ambient air quality standards; appendix
 2             B—reference method for the determination of suspended particulates in the atmosphere (high volume
 3             method). C.F.R. 40: § 50; pp. 12-17.
 4
 5     Commins, B. T. (1962) Interim report on the study of techniques  for determination of polycyclic aromatic
 6             hydrocarbons in air. In: Sawicki, E.; Cassel,  K., Jr., eds. Analysis of carcinogenic air pollutants:
 7             [proceedings of a symposium]; August 1961;  Cincinnati, OH. Bethesda,  MD: National Cancer Institute;
 8             pp. 225-233. (National Cancer Institute monograph no.  9).
 9
10     Conner, T. L.; Miller, J. L.; Willis, R. D.; Kellogg, R. D.; Dann, T. F. (1993) Source apportionment of fine
11             and coarse particles in southern Ontario, Canada.  Presented at: 86th annual meeting and exhibition of the
12             Air & Waste Management Association; June; Denver, CO. Pittsburgh, PA: Air & Waste Management
13             Association; paper no. 93-TP-58.05.
14
15     Cooper, D. W.; Guttrich, G. L. (1981) A study of the cut diameter concept for  interpreting particle sizing data.
16             Atmos. Environ. 15: 1699-1707.
17
18     Cornille, P.;  Maenhaut, W. (1990) Atmos.  Environ. Part  A 24: 1083-1093.
19
20     Gotham, W.  E.; Bidleman, T. F. (1992)  Laboratory investigations of the partitioning of organochlorine
21             compounds between the gas phase and atmospheric aerosols on glass  fiber filters. Environ. Sci. Technol.
22             26: 469-478.
23
24     Countess, R.  J. (1990) Inter-laboratory analyses of carbonaceous aerosol samples. Aerosol Sci. Technol.
25             12: 114-121.
26
27     Courtney, W. J.;  Shaw, R. W.; Dzubay, T. G. (1982) Precision and accuracy of a beta-gauge for aerosol mass
28             determinations. Environ.  Sci. Technol. 16: 236-239.
29
30     Coutant, R. W.; Brown, L.; Chuang, J. C.; Riggin, R. M.; Lewis, R. G. (1988) Phase distribution and artifact
31             formation in ambient air sampling for polynuclear aromatic hydrocarbons. Atmos. Environ. 22: 403-409.
32
33     Coutant, R. W.; Callahan, P. J.; Kuhlman, M. R.; Lewis, R. G.  (1989) Design and performance of a
34             high-volume compound annular denuder. Atmos. Environ. 23: 2205-2211.
35
36     Coutant, R. W.; Callahan, P. J.; Chuang, J. C.; Lewis, R.  G.  (1992) Efficiency of silicone-grease-coated
37             denuders for collection of polynuclear aromatic  hydrocarbons. Atmos. Environ. Part A 26: 2831-2834.
38
39     Criss, J. W.  (1976) Particle size and composition effects in X-ray fluorescence analysis of pollution samples.
40             Anal. Chem. 48:  179-186.
41
42     Crutcher, E.  R. (1982) Light microscopy as an analytical  approach to receptor modeling. In:  Dattner,  S. L.;
43             Hopke, P. K., eds. Receptor models applied to  contemporary pollution problems. Pittsburgh, PA: Air
44             Pollution Control Association; pp. 266-284.
45
46     Currie, L. A. (1982) Contemporary paniculate carbon.  In: Wolff, G. T.; Klimisch, R. L., eds. Paniculate
47             carbon: atmospheric  life cycle. Chelsea, MI:  Lewis  Publishers, Inc.; pp. 47-65.
48
49     Daisey, J. M. (1987)  Chemical composition of inhalable paniculate matter—seasonal and intersite comparisons.
50             In: Lioy, P. J.; Daisey, J. M., eds. Toxic air pollution: a comprehensive study of non-criteria air
51             pollutants. Chelsea, MI: Lewis Publishers, Inc.; pp. 45-65.
52
53
         April  1995                                     4-104      DRAFT-DO NOT QUOTE OR CITE

-------
  1     Daisey, J. M.; Kneip, T. J. (1981) Atmospheric paniculate organic matter: multivariate models for identifying
  2             sources and estimating their contributions to the ambient aerosol. In: Macias, E. S.; Hopke, P. K., eds.
  3             Atmospheric aerosol: source/air quality relationships; based on a symposium jointly sponsored by the
  4             Divisions of Nuclear Chemistry and Technology and Environmental Chemistry  at the 180th national
  5             meeting of the American Chemical Society; August 1980; Las Vegas, NV. Washington, DC: American
  6             Chemical Society; pp. 197-221. (Comstock, M. J., ed. ACS symposium series: 167).
  7
  8     Daisey, J. M.; McCaffrey, R. J.; Gallagher, R. A. (1981) Polycyclic aromatic hydrocarbons and total extractable
  9             paniculate organic matter in the Arctic aerosol. Atmos. Environ. 15:  1352-1363.
 10
 11     Daisey, J. M.; Herschman, R. J.;  Kneip, T.  J. (1982) Ambient levels of paniculate organic matter in New York
 12             City in winter and summer. Atmos. Environ. 16: 2161-2168.
 13
 14     Dams, R.; Robbins, J. A.; Rahn, K. A.; Winchester, J. W. (1970) Non-destructive neutron activation analysis of
 15             air pollution particulates. Anal. Chem. 42: 861-867.
 16
 17     Davis, B. L. (1978) Additional suggestions for X-ray quantitative analysis of high-volume filters.  Atmos.
 18             Environ. 12: 2403-2406.
 19
 20     Davis, B. L. (1980) Standardless' X-ray diffraction quantitative analysis. Atmos. Environ. 14: 217-220.  -
 21
 22     Delia Fiorentina, H.; De Wiest, F.; De Graeve, J. (1975) Determination par spectrometrie infrarouge de la
 23             matiere organique non volatile associee aux particules en suspension dans 1'air—II. Facteurs influencant
 24             1'indice aliphatique. [Determination of the nonvolatile organic matter associated with suspended panicles
 25             by infrared spectrometry—II. Factors influencing the aliphatic index]. Atmos. Environ. 9: 517-522.
 26
 27     Derde, M. P.; Buydens,  L.; Guns, C.;  Massart, D. L.; Hopke, P. K. (1987)  Comparison of rule-building expert
 28             systems with pattern recognition for the classification of analytical data. Anal. Chem. 59: 1868-1871.
 29
 30     Divita, F., Jr. (1993) Size distributions and sources od  submicrometer atmospheric particles in Washington, D.C.
 31            and Philadelphia,  PA [dissertation]. College Park,  MD: University of Maryland. Available from: Xerox
 32             University Microfilms, Ann Arbor, MI; publication no. AAD94-25018.
 33
 34     Dod, R. L.;  Rosen,  H.; Novakov,  T. (1979) Atmospheric aerosol research  annual report for 1977-1978.
 35            Berkeley, CA: Lawrence Berkeley Laboratories; document no. 8696.
 36
 37     Drane, E. A.; Branton, D. G.; Tysinger, S. H.; Courtney, W. J. (1983) Data processing  procedures for
 38            elemental analysis of atmospheric aerosols by X-ray fluorescence. Research Triangle Park, NC: Northrop
 39            Services, Inc.; document no. TR-83-01.
 40
 41     Durham, J.  L.; Wilson, W. E.; Bailey, E. B. (1978) Application of an SO2 denuder for continuous measurement
 42            of sulfur in submicrometric aerosols.  Atmos. Environ. 12: 883-886.
 43
 44     Dzubay, T.  G.  (1986) Analysis of aerosol samples by X-ray fluorescence. Research Triangle Park, NC:  U.S.
 45            Environmental Protection Agency.
46
47     Dzubay, T.  G.; Clubb, K. W. (1981) Comparison of telephotometer measurements of extinction coefficients with
 48            scattering and absorption coefficients. In: White, W. H.; Moore, D. J.;  Lodge, J. P., Jr., eds. Plumes
 49            and  visibility: measurements and model components, proceedings of the symposium; November 1980;
 50            Grand Canyon National Park, AZ. Atmos. Environ. 15: 2617-2624.
 51
52     Dzubay, T.  G.; Mamane,  Y. (1989) Use of electron microscopy data in  receptor models for PM-10. Atmos.
53             Environ. 23: 467-476.
54


         April  1995                                    4.105       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Dzubay, T. G.; Nelson, R. O. (1975) Self absorption corrections for X-ray fluorescence analysis of aerosols.
  2             Adv. X-ray Anal.  18: 619-631.
  3
  4      Dzubay, T. G.; Stevens, R. K. (1975) Ambient air analysis with dichotomous sampler and X-ray fluorescence
  5             spectrometer.  Environ. Sci. Technol. 9: 663-668.
  6
  7      Dzubay, T. G.; Morosoff, N.; Whitaker, G. L.; Yasuda, H. (1981) Evaluation of polymer films as standards for
  8             X-ray fluorescence spectrometers.  In: Electron microscopy and X-ray applications to environmental and
  9             occupational health analysis. Ann Arbor, MI: Ann Arbor Science Publishers, Inc.
10
11      Dzubay, T. G.; Stevens, R. K.; Balfour, W. D.; Williamson, H. J.;  Cooper, J.  A.; Core, J. E.;
12             De Cesar,  R.  T.; Crutcher, E. R.; Datiner, S. L.; Davis, B.  L.; Heisler, S. L.; Shah,'J. J.;
13             Hopke, P.  K.; Johnson, D. L. (1984) Interlaboratory comparison of receptor model results for Houston
14             aerosol. Atmos. Environ. 18: 1555-1566.
15
16      Eatough, D. J. (1995) BOSS, the Brigham Young University Organic Sampling System: determination of
17             paniculate carbonaceous material using diffision denuder sampling  technology. In: Lane,  D.,
18             ed.  Gas/phase partition measurements of atmospheric organic compounds. New York, NY: Gordon and
19             Breach Science Publishers;  in press.
20
21      Eatough, D. J.; Sedar, B.; Lewis, L.; Hansen, L. D.; Lewis, E. A.; Farber, R. J. (1989) Determination of
22             semivolatile organic compounds in particles in the Grand Canyon area. Aerosol Sci. Technol.
23             10: 438-449.
24
25      Eatough, D. J.; Wadsworth, A.; Eatough, D. A.; Crawford, J. W.; Hansen, L. D.; Lewis, E. A. (1993)
26             A multiple-system, multi-channel diffusion denuder sampler for the determination of fine-particulate
27             organic material in the atmosphere. Atmos. Environ. Part A 27:  1213-1219.
28
29      Eatougrs D.J.; Tang,  H.; Cui, W.; Machir, J. (1995)  Determination of the size  distribution and chemical
30             composition of fine paniculate semi-volatile organic material  in urban environments using diffusion
31             denuder technology.  Inhalation Toxicol. 7: 691-710.
32
33      Eichmann, R.; Neuling, P.; Ketseridis, G.; Hahn, J.; Jaenicke, R.; Junge, C.  (1979) n-alkane studies in the
34             troposphere—I. Gas and paniculate concentrations in North Atlantic air.  Atmos. Environ. 13: 587-599.
35
36      Eldering, A.;  Cass, G. R.; Moon, K. C. (1994) An air monitoring network using continuous particle size
37             distribution monitors: connecting pollutant properties to visibility via Mie scattering calculations. Atmos.
38             Environ. 28: 2733-2749.
39
40      Eldred, R. A.; Cahill, T. A. (1994) Trends in elemental concentrations of fine particles at remote sites in  the
41             United States  of America.  Atmos. Environ. 28: 1009-1019.
42
43      Eldred, R. A.; Cahill, T. A.; Feeney, P.  (1989) Regional patterns of selenium and other trace elements in the
44             IMPROVE network.
45
46      Eldred, R. A.; Cahill, T. A.; Malm, W. C.  (1989) Composition of fine aerosols during the summer of 1987
47             determined by the IMPROVE and National Service networks. In: Proceedings of the 8th annual AAAR
48             meeting; Reno, NV.
49
50      Eldred, R. A.; Cahill, T. A.; Feeney, P. J. (1993) Comparison of independent measurements of sulfur and
51             sulfate in the IMPROVE network. Presented at: 86th annual meeting and exhibition of the Air & Waste
52             Management Association; June;  Denver, CO. Pittsburgh, PA: Air  & Waste Management Association;
53             paper no. 93-RA-l 10.02.
54


         April 1995                                    4-106        DRAFT-DO NOT QUOTE OR CITE

-------
   1      Eldred, R. A.; Cahill, T. A.; Flocchini, R. G. (1994) Composition of PM10 and PM2 5 aerosols in the
   2             IMPROVE network. Proceedings of the international specialty conference on aerosol and atmospheric
   3             optics: radiative balance and visual air quality, volume A. Air Waste: submitted.
   4
   5      Engelbrecht, D. R.; Cahill, T. A.; Feeney, P. J. (1980) Electrostatic effects on gravimetric analysis of
   6             membrane filters. J.  Air Pollut. Control Assoc. 30: 391-392.
j   7
;   8      Federal Register. (1978) Reference method for the determination of lead in suspended paniculate matter collected
|   9             from ambient air. F. R. 43: 46258.
!  10
  11      Federal Register. (1987) Procedures for testing performance characteristic of methods for PM10. F. R.
  12             52: 24724.
  13
  14      Federal Register. (1987) Revisions to the national ambient air quality standards for paniculate matter.  F. R. (July
  15             1) 52: 24634-24669.
  16
  17      Feeney, P. J.; Cahill, T. A.; Olivera, J.; Guidara, R. (1984) Gravimetric determination of mass on lightly loaded
  18             membrane filters. J.  Air Pollut. Control Assoc.  34: 367-378.
  19
  20      Ferek, R. J.; Lazrus, A. L.; Haagenson, P. L.; Winchester, J. W. (1983) Strong and weak acidity of aerosols
  21             collected over the northeastern United States. Environ. Sci. Technol. 17: 315-324.
  22
  23      Ferek, R. J.; Hegg, D.  A.;  Hobbs, P. V.; Orphan, M.; Farber, R. J.; Bhardwaja, P. S. (1987)  Airborne
  24             measurements of atmospheric  trace constituents in the  Grand Canyon region. In: Mathai, C. V.,  ed.
  25             Visibility and fine particles: transactions. Pittsburgh, PA: Air & Waste Management Association; p. 328.
  26
  27     Perm, M. (1986) Concentration measurements and equilibrium studies of ammonium, nitrate and sulphur species
  28             in air and precipitation [Ph.D. thesis]. Goteborg, Sweden: Goteborg University, Department of Inorganic
  29             Chemistry.
  30
  31      Fernandez, F. J. (1989) Atomic absorption spectroscopy. In: Lodge, J. P., Jr., ed. Methods of air sampling and
  32             analysis. 3rd ed. Chelsea, MI: Lewis Publishers; pp.  143-150.
  33
  34     Finlayson-Pitts, B.  J.; Pitts, J. N., Jr. (1986) Atmospheric chemistry: fundamentals and experimental  techniques.
  35              New York, NY: John Wiley & Sons.
  36
  37      Fitz, D. R. (1990)  Reduction of the positive organic  artifact on quartz filters. Aerosol. Sci. Technol.
  38              12: 142-148.
  39
 40      Flessel,  P.; Wang,  Y. Y.; Chang, K. I.; Wesolowski, J. J.; Guirguis, G.  N.; Kim, I. S.; Levaggi, D.; Wayman,
 41             S. (1991) Seasonal variations and trends in concentrations of filter-collected polycyclic aromatic
 42             hydrocarbons (PAH) and mutagenic activity in the San Francisco Bay area. J. Air Waste  Manage. Assoc.
 43             41: 276-281.
 44
 45      Fletcher, R. A.; Small,  J. A. (1993) Analysis of individual collected particles. In: Willeke, K.; Baron, P. A.,
 46             eds. Aerosol measurement principles, techniques and applications. New York, NY:  Van Nostrand
 47             Reinhold; pp. 260-295.
 48
 49      Foreman, W.  T.; Bidleman, T. F. (1990) Semivolatile organic compounds in the ambient air of Denver,
 50             Colorado. Atmos. Environ.  Part A 2405-2416.
 51
 52      Forrest,  J.; Spandau, D. J.; Tanner, R. L.;  Newman, L.  (1982) Determination of atmospheric nitrate and nitric
 53             acid employing a diffusion denuder with a filter pack.  Atmos. Environ.  16: 1473-1485
 54


          April  1995                                    4.107       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Friedlander, S. K. (1977) Smoke, dust and haze: fundamentals of aerosol behavior.  New York, NY: John Wiley
 2             & Sons, Inc.
 3
 4      Fung, K. K. (1990) Paniculate carbon speciation by MnO2 oxidation. Aerosol Sci. Technol.  12: 122-127.
 5
 6     .Fung, K. K.; Heisler, S. L.; Price, A.; Nuesca, B. V.; Mueller, P. K. (1979) Comparison of ion
 7             chromatography and automated wet chemical methods for analysis of sulfate and nitrate in ambient
 8             paniculate filter samples. In: Mulik, J. D.; Sawicki, E., eds. Ion chromatographic analysis of
 9             environmental pollutants: volume 2. Ann Arbor, MI: Ann Arbor Science Publishers Inc.; pp. 203-209.
10
11      Gagosian, R. B.; Peltzer, E. T.; Zafiriou, O. C. (1981) Atmospheric transport of continentally derived lipids to
12             the tropical North Pacific. Nature (London) 291: 321-324.
13
14      Galloway, J. N.; Cosby, B. S.; Likens, G. E. (1979) Acid precipitation: measurement of pH and acidity.
15             Limnol. Oceanogr. 24: 1161.
16
17      Gebhart, J. (1993) Optical direct-reading techniques: light intensity systems. In: Willeke, K.; Baron, P.  A., eds.
18             Aerosol measurement principles, techniques and applications. New  York,  NY: Van Nostrand Reinhold;
19             pp. 313-344.
20
21      Gerber, H. E. (1982) Optical techniques for the measurement of light absorption by particulates. In:
22             Wolff,  G. T.; Klimisch, B. L., eds.  Particulate carbon: atmospheric life cycles.  New York,  NY: Plenum
23             Publishing Corporation; pp.  1-411.
24
25      Gordon, R. J.  (1974) Solvent selection in entrance of paniculate matter. Atmos. Environ. 8:  189-191.
26
27      Griffin, J. J.; Goldberg, E. D.  (1979) Morphologies and origin of elemental carbon in the environment.  Science
28             (Washington, DC) 206: 563-565.
29
30      Grinshp'un, S.;  Willeke, K.; Kalatoor, S. (1993) A general equation for aerosol aspiration by thin-walled
31             sampling probes in calm and moving air. Atmos. Environ. Part A 27: 1459-1470.
32
33      Grosjean, D. (1975)  Solvent extraction and carbon and organic carbon determination in atmospheric paniculate
34             matter: the organic extraction-organic carbon analyzer technique. Anal. Chem. 47: 797.
35
36      Gundel, L. A.; Stevens, R. K.; Daisey, J. M.; Lee, V.; Mahanama, K. R. R. (1992) Annular denuders  for
37             sampling semi-volatile polycyclic aromatic hydrocarbons and other organic species. Presented at:
38             llth annual meeting of the American Association for Aerosol Research; October; San Francisco, CA.
39
40      Gundel, L. A.; Mahanama, K. R.  R.; Daisey, J. M. (1994) Fractionation of polar organic extracts of airborne
41             paniculate matter using cyanopropyl-bonded silica in solid -phase extraction. J. Chromatogr.: in press.
42
43      Hall, D. J.; Mark, D.; Upton, S. L. (1992) A new large wind tunnel for dust and aerosol  studies. J. Aerosol Sci.
44             23(suppl. 1): S591-S594.
45
46      Harrison, R. M.; Kitto, A.  M. N. (1990) Field intercomparison of filter pack and denuder sampling methods for
47             reactive gaseous and paniculate pollutants.  Atmos. Environ. Part A 24: 2633-2640.
48
49      Hart, K. M.; Pankow, J.  F. (1990) Comparison of n-alkane and PAH concentrarions collected on quartz fiber
50             and Teflon membrane filters in an urban environment. J. Aerosol Sci. 21(suppl. 1): S377-S380.
51
52      Hearle, J. W.  S.; Sparrow, J. R.; Cross, P. M. (1972) The use of the  scanning electron microscope. London,
53             United Kingdom: Pergamon Press.
54


         April 1995                                    4-108       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Heitbrink, W. A.; Baron, P. A.; Willeke, K. (1991) Coincidence in time-of-flight aerosol spectrometers:
  2            phantom particle creation. Aerosol Sci. Technol. 14: 112-126.
  3
  4     Henderson Sellers, A.; McGuffie, K. (1989) Sulphate aerosols and climate: scientific correspondence. Nature
  5            (London) 340: 436-437.
  6
  7     Henry, R. C.; Lewis, C. W.; Hopke, P. K.; Williamson, H. J. (1984) Review of receptor model fundamentals.
  8            Atmos. Environ. 18: 1507-1515.
  9
 10     Hering, S. V., ed. (1989) Air sampling instruments for evaluation of atmospheric contaminants. 7th ed.
 11            Cincinnati, OH: American Conference of Governmental Industrial Hygienists.
 12
 13     Hering, S. V. (1994)  Particle measurements for the children's health study: development of a two-week sampler.
 14            Presented at: International Society for Environmental Epidemiology/International Society for Exposure
 15            Analysis joint conference; September; Research Triangle Park, NC. paper no. 259.
 16
 17     Hering, S. V.; Appel, B. R.; Cheng, W.; Salaymeh, F.; Cadle, S. H.; Mulawa,  P. A.; Cahill, T. A.;
 18            Eldred, R. A.; Surovik, M.; Fitz, D.; Howes,  J. E.; Knapp,  K. T.; Stockburger, L.; Turpin, B. J.;
 19            Huntzicker, J. J.; Zhang, X.-Q.; McMurry, P. H. (1990) Comparison of sampling methods for
 20            carbonaceous  aerosols in ambient air.  Aerosol Sci. Technol.  12: 200-213.
 21
 22     Hildemann, L. M.; Markowski, G. R.; Cass,  G. R. (1991) Chemical composition of emissions from urban
 23            sources of fine organic aerosol. Environ. Sci. Technol. 25: 744-759.
 24
 25     Hillamo, R. E.; Kauppinen,  E. I. (1991) On the performance  of the Berner low pressure impactor. Aerosol Sci.
 26            Technol. 14: 33-47.
 27
 28     Hinds, W. C. (n.d.) Basis for particle size-selective sampling for wood dust. In:  Advances in air sampling.
 29
 30     Hinds,*W. C. (1982) Aerosol technology. New York, NY: John Wiley and Sons.
 31
 32     Hinds, W. C.; Kraske, G. (1986) Performance of PMS model LAS-X optical particle counter. J. Aerosol Sci.
 33            17: 67-72.
 34
 35     Hirose, K.; Sugimura, Y. (1984) Excess  228th in the airborne dust: an indicator  of continental dust from the east
 36            Asian deserts. Earth  Planet. Sci. Lett. 70: 110-114.
 37
 38     Hoffman, A. J.;  Purdue, L. J.; Rehme, K. A.; Holland, D. M. (1988) 1987 PM10 sampler intercomparison
 39            study. In: Mathai, C. V.; Stonefield, D. H., eds. PM10—implementation of standards: transactions.
 40            Pittsburgh, PA:  Air Pollution Control Association; pp. 138-149.
 41
 42     Hollander, W. (1990) Proposed performance criteria for samplers of total suspended paniculate matter. Atmos.
43            Environ. Part  A 24:  173-177.
44
45     Hollander, W. (1992) Ambient aerosol sampling without pumps. J. Aerosol Sci. 23(suppl. 1): S619-S622.
46
47     Hopke, P. K. (1985) Receptor modeling in environmental chemistry. New York,  NY: John Wiley and Sons.
48
49     Hopke, P. K., Jr. (1985) The use of chemometrics in apportionment of air pollution sources.  Trends Anal.
50            Chem. 4: 4.
51
52     Hopke, P. K.; Casuccio, G.  S. (1991) Scanning electron microscopy. In: Hopke, P. K., ed. Receptor modeling
53            for air quality management: v. 7. Amsterdam, The Netherlands: Elsevier; pp. 149-212.
54


         April 1995                                   4_109       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Hornung, R. W.; Reed, L. D. (1990) Estimation of average concentration in the presence of non-detectable
  2             values. Appl. Occup. Environ. Hyg. 5: 46-51.
  3
  4      Ingham, D. B.; Yan, B. (1994) The effect of a cylindrical backing body on the sampling efficiency of a
  5             cylindrical sampler. J. Aerosol Sci. 25: 535-541.
  6
  7      International Standards Organization. (1983) Air quality: particle size fraction definitions for health-related
  8             sampling. Geneva, Switzerland: International Standards Organization; technical report ISO/TR7708-1983.
  9
10      Jackson, M. L. (1981) Oxygen isotopic ratios in quartz as an indicator of provenance of dust. Geol. Soc. Am.
11             Spec. Pap.: 27-36.
12
13      Jaklevic, J. M.; Loo, B. W.; Goulding, F. S. (1977) Photon-induced X-ray fluorescence analysis using
14             energy-dispersive detection and dichotomous sampler.  In: Dzubay, T. G., ed. X-ray fluorescence analysis
15             of environmental samples. 2nd ed. Ann Arbor,  MI: Ann Arbor Science Publishers; pp. 3-18.
16
17      Jaklevic, J. M.; Gatti, R. C.; Goulding, F. S.;  Loo, B. W. (1981) A |8-gauge method applied to aerosol samples.
18             Environ. Sci. Technol. 15: 680-686.
19
20      Janocko, P. B.; Casuccio, G. S.; Dattner, S.  L.; Johnson, D.  L.; Crutcher,  E.  R. (1982) The El Paso Airshed:
21             source apportionment using complementary analyses and receptor models. In: Dattner, S. L.;
22             Hopke, P. K., eds. Receptor models applied to contemporary pollution problems.  Pittsburgh, PA: Air
23             Pollution Control Association.
24
25      Javitz, H.  S.; Watson, J. G.; Robinson, N. (1988) Performance of the chemical mass balance model with
26             simulated local scale aerosols. Atmos.  Environ. 22: 2309-2322.
27
28      Jenson, P. A.; O'Brien, D. (1993) Industrial hygiene. In: Willeke, K.; Baron, P. A., eds. Aerosol measurement
29             principles, techniques and applications.  New York, NY: Van Nostrand Reinhold; p. 540.
30
31      John, W.;  Wall, S. M. (1983) Aerosol testing techniques  for size-selective samplers. J. Aerosol Sci.
32             14: 713-727.
33
34      John, W.;  Wang, H.-C. (1991) Laboratory testing method for PM-10 samplers: lowered effectiveness from
35             particle loading. Aerosol Sci. Technol.  14: 93-101.
36
37      John, W.;  Hering, S.; Reischl, S.;  Sasaki, G.; Goren, S.  (1983) Characterization of nuclepore filters with large
38             pore size: II.  filtration properties.  Atmos. Environ. 17: 373-382.
39
40      John, W.;  Wall, S. M.; Ondo, J. L. (1988) A new method for nitric acid and nitrate aerosol  measurement using
41             the dichotomous sampler. Atmos.  Environ. 22: 1627-1635.
42
43      John, W.;  Wall, S. M.; Ondo, J. L.; Winklmayr, W. (1990) Modes in the size distributions of atmospheric
44             inorganic aerosol. Atmos. Environ. Part A 24: 2349-2359.
45
46      John, W.;  Winklmayr, W.; Wang,  H.-C.  (1991) Particle  deagglomeration and reentrainment  in a PM-10 sampler.
47             Aerosol Sci. Technol.  14: 165-176.
48
49      Johnson, D. L.; Mclntyre, B. L. (1982) A particle class balance receptor model for aerosol apportionment in
50             Syracuse, NY. In: Receptor models applied to contemporary pollution problems: proceedings. Pittsburgh,
51             PA: Air Pollution Control Association;  pp. 238-247.
52
53      Johnson, D. L.; Mclntyre, B. L.; Fortmann, R.; Stevens, R. K.; Hanna, R. B. (1981) A chemical element
54             comparison of individual particle analysis and bulk analysis methods. Scanning Electron Microsc. 1: 469.


         April 1995                                   4-110        DRAFT-DO NOT QUOTE OR CITE

-------
  1     Johnson, R. L.; Shah, J. J.; Gary, R. A.; Huntzicker, J. J. (1981) An automated thermal-optical method for the
  2            analysis of carbonaceous aerosol. In: Macias, E.  S.; Hopke, P. K., eds. Atmospheric aerosol: source/air
  3            quality relationships; based on a symposium jointly sponsored by the  Divisions of Nuclear Chemistry and
  4            Technology at the 180th national meeting of the American Chemical  Society; August 1980; Las Vegas,
  5            NV. Washington, DC: American Chemical Society; pp. 223-233. (Comstock,  M. J., ed.  ACS
  6     '       symposium series: 167).
  7
  8     Junge, C. E. (1977) Basic considerations about trace constituents in the atmosphere as related to the fate of global
  9            pollutants. In: Suffet, I. H., ed.  Fate of pollutants in the air and water environments: part I, mechanism
 10            of interaction between environments and mathematical modeling and the physical fate of pollutants,
 11            papers from the 165th national American Chemical Society meeting;  April 1975; Philadelphia, PA. New
 12            York, NY: John Wiley & Sons;  pp. 7-25. (Advances in environmental science and technology: v.  8).
 13
 14     Kado, N. Y.;  Kuzmicky, P. A.; Hsieh,  D. P. H.; Gundel, L. A.; Daisey, J. M.; Mahanama, K. R. R.;
 15            Schuetzle, D. (1989) Mutagenic activity of the polar organic matter and  its fractions from airborne
 16            particles. Presented at: AAAR meeting; October; Traverse City, MI.
 17
 18     Kaupp, H.; Umlauf, G. (1992) Atmospheric gas-particle  partitioning of organic compounds: comparison of
 19            sampling methods. Atmos. Environ. Part A 26: 2259-2267.
 20
 21     Kim, Y. J.; Boatman, J. F.  (1990) The effects that the optical properties have on atmospheric aerosol
 22            measurements with optical particle spectrometers. J. Aerosol Sci.  21(suppl. 1): S551-S554.
 23
 24     Kim, D.; Hopke, P. K. (1988) Classification of individual particles based on computer-controlled scanning
 25            electron microscopy data. Aerosol Sci. Technol.  9: 133-151.
 26
 27     Kim, D.; Hopke, P. K. (1988) Source apportionment of the El Paso aerosol by particle class balance analysis.
 28            Aerosol Sci. Technol.  9: 221-235.
 29
 30     Kim, D. S.; Hopke, P. K.; Massart, D.  L.; Kaufman, L.; Casuccio, G. S. (1987) Analysis of CCSEM auto
 31            emission data. Sci. Total Environ. 59: 141-155.
 32
 33     Klippel, W.; Warneck, P. (1980) The formaldehyde  content of the atmospheric aerosol.  Atmos.  Environ.
 34            14: 809-818.
 35
 36     Knollenberg, R. G.; Veal, D.  L. (1992) Optical particle  monitors, counters, and spectrometers. J. EIS
 37            (March/April): 64-81.
 38
 39     Konig, J.; Funcke, W.; Balfanz, E.; Grosch, B.; Pott, F. (1980) Testing a high volume  air sampler for
40            quantitative collection of polycyclic aromatic hydrocarbons. Atmos. Environ. 14: 609-613.
41
42     Koutrakis, P.; Wolfson, J. M.; Slater, J. L.; Brauer,  M.; Spengler, J. D.; Stevens, R. K.;  Stone, C. L. (1988)
43            Evaluation of an annular denuder/filter pack  system to collect acidic aerosols and gases.  Environ. Sci.
44            Technol. 22:  1463-1468.
45
46     Koutrakis, P.; Wolfson, J. M.; Brauer, M.; Spengler, J.  D. (1990) Design of a glass impactor for an annular
47            denuder/filter pack system. Aerosol Sci. Technol. 12: 607-612.
48
49     Lamb, S. I.; Petrowski, C.;  Kaplan, I. R.; Simoneit, B.  R. T. (1980) Organic compounds in urban atmospheres:
50            a review of distribution, collection and analysis. J. Air Pollut. Control Assoc.  30: 1098-1115.
51
52     Lane, D. A.; Johnson, N. D.; Hanley, M. J. J.; Schroeder, W. H.; Ord, D.  T. (1992) Gas- and particle-phase
53            concentrations of a-hexachlorocyclohexane, g-hexachlorocyclohexane  and hexachlorobenzene in  Ontario
54            air. Environ.  Sci. Technol. 26: 126-132.


         April  1995                                     4_1H       DRAFT-DO NOT  QUOTE OR CITE

-------
 1      Larson, T. V.;  Yuen,  P. F.; Maykut, N.  N. (1992) Weekly composite sampling of PM10 for total mass and trace
 2             elemental analysis. In: Chow, J. C.; Ono, D. M., eds. PM10 standards and non-traditional paniculate
 3             source controls: proceedings of an AWMA/EPA international specialty conference.  Pittsburgh, PA: Air
 4             & Waste Management Association; pp. 39-50.
 5
 6     -Lee, R. E., Jr.; Duffield, F. V. P. (1979) Sources of environmentally important metals in the atmosphere. In:
 7             Risby, T. H.,  ed. Ultrace metal analysis in biological  sciences and environment: proceedings of the  174th
 8             meeting of the American Chemical Society; August 1977; Chicago, IL. Washington, DC: American
 9             Chemical Society; pp.  146-171. (Advances in chemistry series 172).
10
11      Lee, R. J.; Fisher, R. M. (1980) Quantitative characterization of particulates by scanning and high voltage
12             electron microscopy. Washington, DC: National Bureau of Standards; special publication.
13
14      Lee, R. J.; Kelly, J. F. (1980) Overview  of SEM-based automated image analysis.  Scanning Electron Microsc.
15             1: 303.
16
17      Lee, R. J.; Fasiska, E. J.; Janocko, P.  B.; McFarland, D. G.; Penkala, S.  J. (1979) Electron-beam paniculate
18             analysis. Ind. Res. Dev.
19
20      Lee, P. H.; Hoffer, T. E.; Schorran, D. E.; Ellis, E. C.; Moyer, J. W. (1981) Laser transmissometer—a
21             description. Sci.  Total Environ. 23: 321.
22
23      Lee, R. J.; Spitzig, W. A.; Kelly,  J. F.;  Fisher, R. M. (1981) Quantitative metallography by computer
24             controlled scanning electron microscopy. J.  Metals 33: 3.
25
26      Li, C. K.;  Kamens, R. M. (1993) The use of polycyclic aromatic hydrocarbons as a source signature in receptor
27             modeling. Atmos. Environ. Part A 27: 523-532.
28
29      Liden, G.; Kenny, L.  C. (1991) Comparison of measured respirable dust sampler penetration curves with
30             sampling conventions.  Ann. Occup. Hyg. 35: 485-504.
31
32      Liden, G.; Kenny, L.  C. (1992) The performance of respirable dust samplers: sampler bias, precision and
33             inaccuracy. Ann. Occup. Hyg.  36: 1-22.
34
35      Ligocki, M. P.; Pankow, J. F. (1989) Measurements of the gas/particle distributions of atmospheric organic
36             compounds. Environ. Sci. Technol. 23: 75-83.
37
38      Lindskog, A.; Brorstrom-Lunden, E.; Sjodin,  A. (1985) Transformation of reactive PAH on particles by
39             exposure to oxidized nitrogen compounds and ozone. Environ.  Int.  11: 125-130.
40
41      Lioy, P. J.; Wainman, T.; Turner, W.; Marple, V. A. (1988) An intercomparison of the indoor air sampling
42             impactor and the dichotomous sampler for a 10-^m cut size. JAPCA 38: 668-670.
43
44      Lippmann, M.  (n.d.) Environmental toxicology and exposure limits for ambient air. Appl. Occup. Environ. Hyg.
45             8:  847-858.
46
47      Lippmann, M.; Chan, T.  L. (1979) Cyclone sampler performance.  Staub Reinhalt. Luft 39: 7-11.
48
49      Lippmann, M.; Chan, T.  L. (1980) Experimental measurements and empirical modelling of the regional
50             deposition of inhaled particles in humans. Am. Ind. Hyg. Assoc.  J. 41: 399-409.
51
52      Liu, B. Y. H.,  ed. (1976) Fine particles aerosol generation, measurement, sampling and analysis. New York,
53             NY: Academic Press.
54


         April  1995                                    4-112        DRAFT-DO NOT QUOTE OR CITE

-------
  1     Liu, B. Y. H.; Pui, D. Y. H. (1981) Aerosol sampling inlets and inhalable particles. Atmos. Environ.
  2            15: 589-600.
  3
  4     Loo, B. W.; Cork, C. P. (1988) Development of high efficiency virtual impactors. Aerosol Sci. Technol.
  5            9: 167-176.
  5
  7     Lundgren, D.; Burton, R. M. (1995) The effect of particle size distribution on the cut point between fine and
  3            coarse ambient mass  fractions. Inhalation Toxicol.: accepted.
  9
 10     Lundgren, D. A.; Lippmann, M.; Harris, F. S., Jr.; Clark, W.  E.; Marlow, W. H.; Durham, M. D., eds.
 11            (1979) Aerosol measurement: [papers from a workshop]; March 1976; Gainseville, FL. Gainesville,  FL:
 12            University  Presses of Florida.
 13
 14     Malm, W. C.; Sisler, J.  F.; Huffman, D.; Eldred, R. A.; Cahill, T. A. (1994) Spatial and seasonal trends in
 15            particles concentration and optical extinction in the United States. J. Geophys. Res. 99: 1347-1370.
 16
 17     Mamane, Y. (1988) Estimate of municipal refuse incinerator contribution to Philadelphia aerosol. Atmos.
 18            Environ.  22: 2411-2418.
 19
 20     Mark, D.; Lyons, C. P.; Upton, S.  L.; Hall, D. J. (1992) A review of the rationale of current methods  for
 21            determining the performance of aerosol samplers. J. Aerosol Sci. 23(suppl. 1): S611-S614.
 22
 23     Markowski, G. R.  (1987) On identifying and correcting for reentrainment in cascade impactor measurements.
 24            Aerosol Sci. Technol. 7: 143-159.
 25
 26     Marple, V. A.; Rubow, K. L. (1976) J.  Aerosol Sci.  6: 425-433.
 27
 28     Marple, V. A.; Liu, B. Y. H.; Olson, B. A. (1989) Evaluation of a cleanroom concentrating aerosol sampler. In:
 29            Proceedings of the 35th annual meeting of the Institute of Environmental Sciences. Anaheim, CA:
 30            Institute of Environmental Sciences; pp.  360-363.
 31
 32     Marple, V. A.; Liu, B. Y. H.; Burton, R. M. (1990)  A high volume dichotomous sampler:  design and
 33            evaluation.  Presented at: 83rd annual meeting  of the Air and Waste Management Association; June;
 34            Pittsburgh,  PA. Pittsburgh, PA:  Air and Waste Management Association; paper no. 90-84.2.
 35
 36     Marple, V. A.; Rubow, K. L.; Olson, B. A. (1993) Inertial gravitational, centrifugal, and thermal collection
 37            techniques.  In: Willeke, K.;  Baron, P. A., eds. Aerosol  measurement principles, techniques and
 38            applications. New York, NY: Van Nostrand Reinhold; p. 209.
 39
 40     Marshall, T. A.; Mitchell, J.  P.; Griffiths, W. D. (1991)  The behaviour of regular-shaped, non-spherical
 41            particles in  a TSI aerodynamic particle sizer. J. Aerosol  Sci. 22: 73-89.
 42
 43     Masia,  P.; Di Palo, V.; Possanzini,  M. (1994) Uptake of ammonia by nylon filters in filter pack systems.
 44            Atmos. Environ. 28: 365-366.
 45
 46     Massart, D. L.; Kaufman, L. (1983) The interpretation of analytical chemical data by the use of cluster analysis.
 47            New York,  NY: John Wiley & Sons.
48
 49     Maynard, A. D. (1993) Respirable dust sampler  characterisation: efficiency curve reproducibility. J. Aerosol  Sci.
 50            24(suppl. 1): S457-S458.
51
52     Mazurek, M. A.; Simoneit, B. R. T.; Cass, G. R.; Gray, H. A.  (1987) Quantitative high-resolution gas
53             chromatography and high-resolution gas chromatography/mass spectrometry analyses of carbonaceous fine
54             aerosol particles. Int.  J. Environ. Anal. Chem. 29: 119-139.


         April 1995                                     4-113       DRAFT-DO  NOT QUOTE OR CITE

-------
  1      Mazurek, M. A.; Cass,  G.  R.; Simoneit, B. R. T. (1989) Interpretation of high-resolution gas chromatography /
  2             mass spectrometry data aquired from atmospheric organic aerosol samples. Aerosol Sci. Technol.
  3             10: 408-420.
  4
  5      Mazurek, M. A.; Cass,  G.  R.; Simoneit, B. R. T. (1991) Biological input to visibility-reducing aerosol particles
  6             in the remote arid southwestern United States. Environ. Sci. Technol.  25: 684-694.
  7
  8      McDow, S. R.; Huntzicker, J. J. (1990) Vapor absorption artifact in the sampling of organic aerosol: face
  9             velocity effects.  Atmos.  Environ. Part A 24: 2563-2571.
10
11      McDow, S. R.; Huntzicker, J. J. (1993) Vapor adsorption artifact in the sampling of organic aerosol. In:
12             Winegar,  E. D.; Keith, L. H., eds. Sampling and analysis of airborne pollutants. Boca Raton,  FL:  Lewis
13             Publishers,  Inc.; pp. 191-208.
14
15      McFarland, A. R.;  Ortiz, C. A. (1979) Aerosol characterization of ambient paniculate samplers used in
16             environmental monitoring studies.  College Station, TX: Texas  A&M Research Foundation; contract no.
17             68-02-2720.
18
19      McFarland, A. R.;  Ortiz, C. A. (1983) Characterization of Sierra Anderson PM-10 inlet model 246b.  College
20             Station, TX: Texas  A&M University, Department of Civil Engineering; Air Quality Laboratory report
21             4716/02/02/84/ARM.
22
23      McFarland, A. R.;  Wedding,  J.  B.; Cermak, J. E. (1977) Wind tunnel evaluation of a modified Andersen
24             impactor and an all  weather sampler inlet. Atmos. Environ.  11: 535-539.
25
26      McFarland, A. R.;  Ortiz, C. A.; Rodes, C. E. (1982) Wind tunnel evaluation of the British smoke shade
27             sampler. Atmos. Environ.  16: 325-328.
28
29      McFarlaftd, A. R.;  Ortiz, C. A.; Bertch, R. W. (1984) A 10 /xm cutpoint size selective inlet for hi-vol samplers.
30             J. Air Pollut. Control Assoc. 34: 544-547.
31
32      Mclntyre, B. L.; Johnson, D. L. (1982) A particle class balance receptor model for aerosol apportionment  in
33             Syracuse, NY. In: Receptor models applied to contemporary pollution problems.  Pittsburgh, PA: Air
34             Pollution Control Association.
35
36      Merrifield, T. (1989) The measurement of PM10 by /3-absorption. Presented at: 82nd annual meeting and
37             exhibition of the Air & Waste Management Association; June;  Anaheim, CA. Pittsburgh, PA:
38             Air & Waste Management Association; paper no. 89-66A.2.
39
40      Meyer, M. B.; Lijek, J.; Ono, D. (1992) Continuous PM10 measurements in a woodsmoke environment. In:
41             Chow, J. C.; Ono,  D. M., eds. PM10 standards and non-traditional paniculate source controls:
42             proceedings of an AWMA/EPA international specialty conference. Pittsburgh, PA: Air & Waste
43             Management Association; pp. 24-38.
44
45      Meyer-Arendt, J.  R. (1972) Introduction to classical and modern optics. Prentice Hall, Inc.; p. 133.
46
47      Miller, F. J.; Gardner, D. E.; Graham, J. A.; Lee, R. E.,  Jr.; Wilson, W.  E.; Bachmann, J. D. (1979) Size
48             considerations for establishing a standard for inhalable particles. J. Air Pollut. Control Assoc.
49             29: 610-615.
50
51      Mulholland, G. W.; Pui, D. Y.  H.; Kapadia,  A.;  Liu, B. Y. H. (1980) Aerosol number and mass concentration
52             measurements: a comparison of the electrical aerosol analyzer  with other measurement techniques.
53             J. Colloid Interface  Sci.  77: 57-67.
54


         April 1995                                    4_114       DRAFT-DO NOT QUOTE  OR CITE

-------
   1      Mylonas, D. T.; Allen, D. T.; Ehrman, S. H.; Pratsinis, S. E. (1991) The sources and size distributions of
   2             organonitrates in Los Angeles aerosol. Atmos.  Environ. Part A 25: 2855-2861.
   3
   4      Novick, V. J.; Alvarez, J. L.  (1987) Design of a multistage virtual impactor. Aerosol Sci. Technol. 6: 63-70.
   5
   6      Ogden, T. L. (1992) Inhalable and respirable dust: moving from aerosol science to legislation. J. Aerosol  Sci.
   7             23(suppl. 1): S445-S448.
   8
   9      Ogden, T. L.; Birkett, J. L. (1977) The human head as a dust sampler. In: Walton, W.  H., eds. Inhaled particles
 10             IV. Oxford, United Kingdom: Pergamon Press; pp. 93-105.
 11
 12      Okazaki, K.; Willeke, K.  (1987) Transmission and deposition behavior of aerosols in sampling inlets. Aerosol
 13             Sci. Technol.  7: 275-283.
 14
 15      Olmez, I. (1989) Instrumental  neutron activation analysis of atmospheric paniculate matter. In: Lodge, J. P., Jr.,
 16             ed.  Methods of air sampling and analysis. 3rd ed. Chelsea, MI: Lewis Publishers; pp.  143-150.
 17
 18      Olmez, I. (1989) Trace element signatures in groundwater pollution. In: Watson, J.  G.,  ed. Receptor models in
 19             air resources management:  transactions. Pittsburgh, PA: Air & Waste Management Association;
 20             pp.  3-11.
 21
 22      Palen, E. J.; Allen, D. T. (1992) Fourier transform infrared analysis of aerosol formed in the photo-oxidation of
 23             1-octene. Atmos. Environ.  Part A 27: 1471-1477.
 24
 25      Palen, E. J.; Allen, D. T. (1992) Fourier transform infrared analysis of aerosol formed in the photo-oxidation of
 26             isoprene and b-pinene. Atmos. Environ. Part A 26: 1239-1251.
 27
 28      Pankow, J. F. (1987) Review and comparative analysis of the theories on partitioning between the gas and
 29             aerosol particulate phases in the atmosphere. Atmos. Environ. 21: 2275-2283.
 30
 31      Pankow, J. F. (1992) Application of common Y-intercept regression parameters for log Kp vs 1/T for predicting
 32             gas-particle  partitioning in the urban environment. Atmos. Environ. Part A 26: 2489-2497.
 33
 34      Parkes, J.; Rabbitt,  L. J.;  Hamshire, M. J. (1979) Live peak-stripping during X-ray energy-dispersive analysis.
 35             Anal. Chem. 46:  1830-1831.
 36
 37     Patashnick, H.; Rupprecht, E.  G. (1991) Continuous PM10 measurements using the tapered element oscillating
 38             microbalance.  J. Air Waste Manage.  Assoc. 41: 1079-1083.
 39
 40     Pellizzari, E.; Lioy, P.; Quackenboss, J.; Whitmore, R.; Thomas, K.; Clayton, A.; Freeman, N.; Zelon, H.;
 41            Waldman, J.; Rodes, C. (1994) NHEXAS: a pilot study in EPA region  V. Presented at: International
 42            Society for Epidemiology/International Society for Exposure Analysis joint conference;  September;
 43            Research Triangle  Park, NC. paper 415.
 44
 45     Pickle, T.; Allen, D. T.; Pratsinis,  S. E. (1990) The sources and size distributions of aliphatic and carbonyl
 46            carbon in Los Angeles  aerosol. Atmos. Environ. Part A 24: 2221-2228
47
48     Pupp, C.; Lao, R. C.; Murray, J. J.; Pottie, R. F. (1974) Equilibrium vapour concentrations of some polycyclic
49            aromatic  hydrocarbons, As4O6 and SeO2 and the collection efficiencies of these air pollutants. Atmos
50            Environ.  8: 915-925.
51
52     Purdue, L. J.; Rodes, C. E.; Rehme, K.  A.; Holland, D. M.; Bond, A. E. (1986) Intercomparison of
53            high-volume PM10 samplers at a site with high particulate concentrations. J. Air Pollut. Control Assoc
54            36: 917-920.


         April 1995                                    4_H5       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Rader, D. J.; O'Hera, T. J. (1993) Optical direct-reading techniques: in situ sensing. In: Willeke, K.;
  2            Baron, P. A., eds. Aerosol measurement principles, techniques  and applications. New York, NY: Van
  3            Nostrand Reinhold; pp. 345-380.
  4
  5     Ramdahl, T.; Zielinska,  B.; Arey, J.; Atkinson, R.; Winer, A.; Pitts, J. N., Jr. (1986) Ubiquitous occurrence of
  6            2-nitrofluoranthene and 2-nitropyrene in air. Nature (London) 231: 425-427.
  7
  8     Ranade, M. B. (Arun); Woods, M. C.; Chen, F.-L.; Purdue,  L. J.; Rehme, K. A. (1990) Wind tunnel
  9            evaluation of PM10 samplers. Aerosol Sci.  Technol. 13: 54-71.
 10
 11     Reist, P. C. (1984) Introduction to aerosol science. New York, NY: Macmillan Publishing.
 12
 13     Rodes, C.; Holland, D.;  Purdue, L.; Rehme, K. (1985) A field comparison of PM10 inlets at four locations.
 14            J. Air Pollut. Control Assoc. 35: 345-354.
 15
 16     Rodes, C. E.; Kamens, R. M.; Wiener, R. W. (1991) The significance  and characteristics of the  personal
 17            activity cloud on  exposure  assessment measurements for indoor contaminants. Indoor Air  2: 123-145.
 18
 19     Rogge, W. F.; Hildemann, L. M.; Mazurek, M. A.; Cass, G. R.; Simoneit, B. R. T. (1993a) Sources of fine
 20            organic aerosol. 2. Noncatalyst and catalyst-equipped automobiles  and heavy-duty diesel trucks. Environ.
 21            Sci. Technol. 27: 636-651.
 22
 23     Rogge, W. F.; Mazurek, M. A.; Hildemann, L. M.; Cass, G. R.; Simoneit, B. R. T. (1993b) Quantification of
 24            urban organic aerosols at a molecular level: identification, abundance and seasonal  variation. Atmos.
 25            Environ.  Part A 27: 1309-1330.
 26
 27     Rogge, W. F.; Hildemann, L. M.; Mazurek, M. A.; Cass, G. R.; Simoneit, B. R. T. (1993c) Sources of fine
 28            organic aerosol. 3. Road dust, tire debris, and organometallic brake lining dust: roads as sources and
 29            sinks. Environ. Sci. Technol. 27:  1892-1904.
 30
 31     Rondia, D. (1965) Sur la volatilite des hydrocarbures polycycliques [Volatility of polycyclic hydrocarbons]. Int.
 32            J. Air Water Pollut. 9:  113-121.
 33
 34     Rood, M. J.; Covert, D.  S.; Larson, T. V. (1987) Hygroscopic properties of atmospheric aerosol in Riverside,
 35            California. Tellus 39B:  383-397.
 36
 37     Rubow, K. L.; Marple, V. A.; Olin, J.; McCawley, M.A. (1987) A personal cascade impactor: design,
 38            evaluation and calibration.  Am. Ind. Hyg. Assoc. J. 48: 532-538.
 39
40     Rupprecht, E.; Meyers, M.; Patashnick, H. (1992) The tapered element oscillating microbalance as a tool for
41            measuring ambient paniculate concentrations in real time. J. Aerosol Sci. 23(suppl. 1):  S635-S638
42
43     Russ, J. C. (1977) Processing of energy-dispersive X-ray spectra. X-ray Spectrom. 6:  37-55.
44
45     Saltzman, B. E. (1984) Generalized performance characteristics of miniature cyclones for atmospheric paniculate
46            sampling. Am. Ind. Hyg. Assoc. J. 45: 671-680.
47
48     Sass-Kortsak,  A.  M.; O'Brien,  C. R.; Bozek, P. R.; Purdham, J. T. (1993) Comparison of the 10 mm nylon
49            cyclone, horizontal elutriator, and aluminum cyclone for silica and wood dust measurements.  Appl.
50            Occup. Environ.  Hyg. 8: 31-37.
51
52     Saucy,  D. A.; Anderson, J. R.; Buseck, P. R. (1987) Cluster  analysis applied to atmospheric aerosol samples
53            from the Norwegian Arctic. Atmos. Environ. 21:  1649-1657.
54


         April 1995                                    4_116       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Schamber,  F. H. (1993) SEM for nontraditional users. Am. Lab. (June): 34-38.
  2
  3     Schipper, L. B., Ill; Chow, J. C.; Frazier, C. A. (1993) Particulate air toxic emission estimation of the PM10
  4             fraction in natural aggregate processing facilities. Presented at: 86th annual meeting and exhibition of the
  5             Air & Waste Management Association; June; Denver, CO. Pittsburgh,  PA: Air & Waste Management
  6             Association; paper no. 93-MP-6.03.
  7
  8     Schuetzle, D.; Lewtas, J. (1986) Bioassay-directed chemical analysis in environmental research. Anal. Chem.
  9             58: 1060A-1076A.
 10
 11     Schwartz, G. P.; Daisey, J. M.; Lioy, P. J. (1981) Effect of sampling duration on the concentration of
 12             paniculate organics collected on glass fiber filters. Am. Ind. Hyg. Assoc. J. 42: 258-263.
 13
 14     Scott, W. R.; Chatfield, E. J. (1979) A precision SEM image analysis system with full feature EDXA
 15             characterization. Scanning  Electron Microsc. 2:  53.
 16
 17     Shimp,  D.  R. (1988) Field comparison of beta attenuation PM10 sampler and high-volume PM10 sampler. In:
 18             Mathai, C. V.; Stonefield, D. H.,  eds. PM10: implementation of standards, transactions. Pittsburgh, PA:
 19             Air & Waste Management  Association; pp. 171-178.
 20
 21     Sickles, J. E., II;  Hodson, L. L. (1989) Fate of nitrous  acid on selected collection surfaces. Atmos.  Environ.
 22             23: 2321-2324.
 23
 24     Sickles, J. E., II;  Hodson, L. L.; McClenny, W. A.; Paur, R. J.; Ellestad,  T.  G.; Mulik, J. D.; Anlauf, K. G.;
 25             Wiebe, H. A.; Mackay, G. I.; Schiff, H. I.; Bubacz, D. K. (1988)  Field comparison of methods for the
 26             measurement of gaseous and particulate contributors to acidic dry deposition. Atmos. Environ. Part A
 27             24: 155-165.
 28
 29     Sides, J. O.; Saiger, H. F. (1976)  Effects of prolonged static exposures of filters to ambient air on high volume
 30             sampling results. Topeka, KS: Department of Health and Environment.
 31
 32     Simoneit, B. R. T. (1984) Organic matter of the troposphere—III. characterization and sources of petroleum and
 33            pyrogenic residues in aerosols over the western United States. Atmos. Environ. 18: 51-67.
 34
 35     Simoneit, B. R. T.; Sheng, G.; Chen, X.;  Fu, J.; Zhang, J.;  Xu, Y. (1991) Molecular marker study of
 36            extractable organic matter in aerosols from urban areas of China. Atmos. Environ. Part A
 37            25:2111-2129.
 38
 39     Sioutas, C.; Koutrakis, P.; Olson, B. A. (1994) Development and evaluation of a low cutpoint virtual impactor.
 40            Aerosol Sci. Technol. 21: 223-235.
 41
 42     Sloane, C. S. (1986) Effect of composition on aerosol light scattering efficiencies. Atmos. Environ.
 43             20:  1025-1037.
 44
 45      Spagnolo, G. S. (1989) Automatic instrument for aerosol sampler using the beta-particle attenuation.  J. Aerosol
 46            Sci. 20: 19-27.
47
48      Spagnolo, G. S.; Paoletti, D. (1994) Automatic system for three fractions sampling of aerosol particles in
49             outdoor environments. J. Air Pollut. Control Assoc. 44: 702-706
 50
51      Spicer, C. W.; Schumacher, P. M. (1979)  Particulate nitrate: laboratory and field studies of major sampling
52             interferences. Atmos. Environ.  13: 543-552.
53
         April 1995                                    4_H7       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Stevens, R. K.; Pace, T. G. (1984) Overview of the mathematical and empirical receptor models workshop
  2            (Quail Roost II). Atmos. Environ. 18: 1499-1506.
  3
  4     Stevens, R. K.; Dzubay, T. G.; Shaw, R. W., Jr.; McClenny, W.  A.; Lewis, C. W.; Wilson, W.  E. (1980)
  5            Characterization of the aerosol in the Great Smoky Mountains. Environ. Sci. Technol.  14:  1491-1498.
  6
  7     Stevens, R. K.; Pinto, J.; Mamane, Y.;  Ondov, J.; Abdulraheem, M.; Al-Majed, N.; Sadek, M.; Cofer, W.;
  8            Ellenson, W.; Kellogg, R. (1993) Chemical and physical properties of emissions from Kuwaiti oil fires.
  9            Water Sci.  Technol. 27: 223-233.
 10
 11     Sweitzer, T. (1985) A field evaluation of two  PM10 inlets  in an industrialized area of Illinois. J. Air Pollut.
 12            Control Assoc.  35: 744-746.
 13
 14     Tang, H.; Lewis, E. A.; Eatough, D. J.; Burton, R. M.; Farber, R. J. (1994) Determination of the particle size
 15            distribution and chemical composition of semi-volatile organic compounds in atmospheric fine particles
 16            with a diffusion denuder sampling system.  Atmos. Environ. 28: 939-947.
 17
 18     Teague, S. V.; Raabe, O. G.; Elred, R. A. (1992) A bench  top wind tunnel for testing PM10 sampling inlets. In:
 19            Chow, J. C.; Ono, D. M., eds.  PM10 standards and  non-traditional particulate source controls:
 20            proceedings of an AWMA/EPA  international specialty conference.  Pittsburgh, PA: Air & Waste
 21            Management Association;  pp. 2-12.
 22
 23     Thanukos, L. C.; Miller, T.; Mathai, C. V.; Reinholt, D.; Bennett, J. (1992) Intercomparison of PM10 samplers
 24            and source  apportionment of ambient PM10 concentrations at Rillito, Arizona.  In: Chow, J.  C.;
 25            Ono, D. M., eds.  PM10 standards and non-traditional particulate source controls: proceedings of an
 26            AWMA/EPA international specialty conference. Pittsburgh, PA: Air &  Waste Management Association;
 27            pp. 244-261.
 28
 29     Torok, S..B.; Van  Grieken,  R. E.  (1994) X-ray spectrometry. Anal. Chem. 66:  186R-206R.
 30
 31     Tsai, P. J.; Vincent, J.  H. (1993) Impaction model for the aspiration efficiencies of aerosol samplers at large
 32            angles with respect to the wind.  J. Aerosol Sci. 24: 919-928.
 33
 34     Tufto, P. A.; Willeke, K. (1982) Dynamic evaluation of aerosol sampling inlets. Environ. Sci.  Technol.
 35            16: 607-609.
 36
 37     Turner, J. R.; Hering,  S. V. (1987) Greased and oiled substrates as bounce-free impaction surfaces. J. Aerosol
 38            Sci. 18: 215-224.
 39
40     Turpin, B. J., et al. (1993) Elemental analysis of single atmospheric particles influencing visibility at the Grand
41            Canyon. In: Proceedings of the 51st annual meeting of the Microscopy Society of America. San
42            Francisco, CA:  San Francisco Press.
43
44     Turpin, B. J.; Liu,  S.-P.; Podolske, K. S.; Gomes, M. S.  P.; Eisenreich,  S. J.;  McMurry, P. H. (1993) Design
45            and evaluation of a novel diffusion separator for measuring gas/particle  distributions of semivolatile
46            organic compounds.  Environ. Sci. Technol. 27: 2441-2449.
47
48     Turpin, B. J., et al. (1994) Single particle analysis of atmospheric aerosols by transmission electron microscopy:
49            technique development and application to the Mohave visibility study. To be submitted.
50
51     Turpin, B. J., et al. (1994) Description of a cluster analysis technique to categorize single particle compositions
52            and application to Grand Canyon aerosols.  Aerosol Sci. Technol.:  submitted.
53
         April  1995                                     4-118       DRAFT-DO NOT QUOTE OR CITE

-------
    1     Turpin, B. J.; Huntzicker, J. J.; Hering, S. V. (1994) Investigation of organic aerosol sampling artifacts in the
    2             Los Angeles basin. Atmos. Environ. 28: 3061-3071.
    3
    4     U.S. Environmental Protection Agency. (1976) Laboratory procedures for the analysis of ammonia in particulates
    5             collected by means of HI-VOL samplers-Technicon autoanalyzer II procedures. Research Triangle Park,
    6             NC: Analytical Chemistry Branch.
    7
    8     U.S. Environmental Protection Agency. (1982a) Air quality criteria for particulate matter and sulfur oxides: v. II.
    9             Research Triangle Park, NC: Environmental Criteria and Assessment Office; EPA report no.
   10             EPA-600/8-82-029b. Available from: NTIS, Springfield, VA; PB84-120419.
   11
   12     U.S. Environmental Protection Agency. (1982b) Air quality criteria for particulate matter and sulfur oxides:
   13             v. II. Research Triangle Park, NC: Environmental Criteria and Assessment Office; pp.  3-40; EPA report
   14             no. EPA-600/8-82-029b. Available from: NTIS, Springfield, VA; PB84-120419.
   15
   16     U.S. Environmental Protection Agency. (1982c) Air quality criteria for particulate matter and sulfur oxides: v. II.
   17             Research Triangle Park, NC: Environmental Criteria and Assessment Office; pp.  14-112; EPA report no.
   18             EPA-600/8-82-029b. Available from: NTIS, Springfield, VA; PB84-120419.
   19
   20     U.S. Environmental Protection Agency. (1992) Determination of the strong acidity of atmospheric fine-particles
   21             (<2.5 urn) using annular denuder technology. Washington, DC:  Atmospheric Research and Exposure
   22             Assessment Laboratory; EPA report no. EPA/600/R-93/037.
   23
   24     Upton, S. L.; Mark, D.; Hall,  D.  J. (1992) Improvements to the design and performance of an ambient inhalable
   25             aerosol sampler. J. Aerosol Sci. 23(suppl.  1): S595-S598.
   26
   27     Van Borm, W. (1989) Source apportionment of atmospheric particles by electron probe X-ray microanalysis and
'  28             receptor models. Belgium: University of Antwerp.
   29
   30     Van Borm, W. A.; Adams, F.  C. (1988) Cluster analysis of electron microscopy analysis data of individual
   31            particles for the source  apportionment of air paniculate matter. Atmos. Environ. 22: 2297-2307.
   32
 .  33     Van Vaeck, L.; Van Cauwenberghe, K.; Janssens, J. (1984) The gas-particle distribution of organic aerosol
 34              constituents: measurements of the volatilisation artefact in hi-vol cascade impactor sampling. Atmos.
 35              Environ. 18: 417-430.
 36
 37       Vanderpool, R. W.; Lundgren, D. A.; Marple, V. A.; Rubow,  K.  L. (1987) Cocalibration of four large-panicle
 38              impactors. Aerosol Sci. Technol.  7: 177-185.
 39
 40      Vaughan, N. P. (1989) The Andersen impactor: calibration, wall losses and numerical simulation. J. Aerosol Sci.
 41             20: 67-90.
  42
  43      Vossler, T. L.; Stevens, R. K.; Paur, R. J.; Baumgardner, R. E.; Bell, J. P. (1988) Evaluation of improved
   44            inlets and annular denuder systems to measure inorganic air pollutants. Atmos. Environ. 22: 1729-1736.
  45
  46     Waggoner, A. P.; Weiss, R.  E. (1980) Comparison of fine particle mass  concentration and light scattering in
  47            ambient aerosol. Atmos. Environ.  14: 623-626.
  48
  49     Wake, D. (1989) Anomalous  effects in filter penetration measurements using the aerodynamic particle sizer (APS
  50            3300). J. Aerosol Sci. 20:  13-17.
  51
  52      Wang, H.-C.; John,  W. (1987)  Comparative bounce properties of particle materials.  Aerosol  Sci. Technol.
  53             7: 285-299.
  54


           April  1995                                    4_119       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Wang, H.-C.; John, W. (1988) Characteristics of the Berner impactor for sampling inorganic ions. Aerosol Sci.
 2             Technol. 8: 157-172.
 3
 4      Wang, H.; John, W. (1989) A simple iteration procedure to correct for the density effect in the aerodynamic
 5             particle size. Aerosol Sci. Technol. 10: 501-505.
 6
 7      Watson, J. G.; Chow, J. C. (1993) Ambient air sampling.  In:  Willeke, K.; Baron, P. A., eds. Aerosol
 8             measurement principles, techniques and applications. New York, NY: Van Nostrand Reinhold; p. 627.
 9
10      Watson, J. G.; Chow, J. C.; Shah, J. J.; Pace, T. G. (1983) The effect of sampling inlets on the PM-10 and
11             PM-15 to TSP concentration ratios. J. Air Pollut. Control Assoc.  33:  114-119.
12
13      Watson, J. G.; Bowen, J. L.; Chow, J. C.; Rogers, C. F.; Ruby, M. G.; Rood, M. J.; Egami, R. T. (1989a)
14             Method 501: high volume measurement of size classified suspended paniculate matter. In: Lodge, J. P.,
15             ed. Methods of air sampling and analysis. 3rd ed. Chelsea, MI: Lewis Publishers, Inc.; pp. 427-439.
16
17      Watson, J. G.; Lioy, P. J.; Mueller, P. K. (1989b) The  measurement process: precision, accuracy, and validity.
18             In: Hering, S. V., ed. Air sampling instruments  for evaluation of atmospheric contaminants. 7th ed.
19             Cincinnati, OH: American Conference of Governmental Industrial Hygienists; pp. 51-57.
20
21      Wedding, J. B. (1980) Ambient aerosol sampling: history,  present thinking and a proposed inlet for inhalable
22             paniculate matter. Presented at: 73rd  annual meeting of the Air Pollution Control Association; June;
23             Montreal, Canada. Pittsburgh,  PA: Air Pollution Control Association; paper no.  80-38.1.
24
25      Wedding, J. B.; Carney, T.  C. (1983) A quantitative technique for determining the impact  of non-ideal ambient
26             sampler inlets on the collected  mass. Atmos.  Environ.  17: 873-882.
27
28      Wedding, J. B.; Weigand, M. A. (1983) A thoracic panicle inlet (D50 =  6 ^m, D0  =  10 /*m) for high volume
29             sampler. Atmos. Environ. 17:  1203-1204.
30
31      Wedding, J. B.; Weigand, M. A. (1993) An automatic particle sampler with beta gauging.  J. Air Waste Manage.
32             Assoc. 43: 475-479.
33
34      Wedding, J. B.; Weigand, M. A.; Carney, T. C. (1982) 10 m Cutpoint inlet for the dichotomous sampler.
35             Environ. Sci. Technol. 16: 602-606.
36
37      Wernisch, J. (1985) Quantitative electron microprobe analysis  without standard.  X-ray  Spectrom.  14:  109-119.
38
39      Wernisch, J. (1986) Application of different (p-z) distributions in quantitative electron probe microanalysis with
40             standard samples. Radex-Rundschau 2/3: 110-113.
41
42      Wiener, R. W.; Rodes,  C. E. (1993) Indoor aerosols and aerosol exposure. In: Willeke, K.; Baron, P. A., eds.
43             Aerosol measurement principles, techniques and  applications.  New York, NY:  Van Nostrand Reinhold;
44             pp. 659-689.
45
46      Willeke, K.; Baron, P. A., eds. (1993) Aerosol measurement: principles, techniques and applications. New
47             York, NY: Van Nostrand Reinhold Publishers.
48
49      Willeke, K.; Degarmo,  S. J. (1988) Passive versus active aerosol  monitoring. Appl. Ind. Hyg. 3: 110-114.
50
51      Willeke, K.; Grinshpun, S. A.;  Chang, C.; Juozaitis, A.; Liebhaber,  F.;  Nevalainen, A.; Thompson, M. (1992)
52             Inlet sampling efficiency of bioaerosol samplers. J. Aerosol Sci. 23(suppl. 1): S651-S654.
53
         April  1995                                    4-120       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Williams, E. L., II; Grosjean, D. (1990) Removal of atmospheric oxidants with annular denuders.  Environ. Sci.
  2            Technol. 24: 811-814.
  3
  4     Wilson, J. C.;  Gupta, A.; Whitby, K. T.; Wilson, W. E. (1988) Measured aerosol light scattering  coefficients
  5            compared with values calculated from EAA and optical particle counter measurements: improving the
  6            utility of the comparison.  Atmos. Environ. 22: 789-793.
  7
  8     Winberry, W. T.; Forehand, L.; Murphy, N. T.; Ceroli, A.; Phinney, B.; Evans, A. (1993) Determination of
  9            reactive acidic and basic gases and paniculate matter in indoor air (annular denuder technique, method
 10            IP-9). In: Methods for determination of indoor air pollutants: EPA methods. Park  Ridge, NJ: Noyes Data
 11            Corporation.
 12
 13     Woskie, S.  R.; Shen, P.; Finkel, M.; Eisen, E. A.; Smith, T. J.; Smith, R.; Wagman,  D. H. (1993) Calibration
 14            of a continuous-reading aerosol monitor (MINIRAM) to measure borate dust exposures. Appl. Occup.
 15            Environ. Hyg. 8: 38-45.
 16
 17     Yamada, Y. (1983) A new method for the determination of collection efficiency of an aerosol sampler by
 18            electron microscopy. Atmos. Environ. 17: 369-372.
 19
20     Yamasaki, H.;  Kuwata,  K.; Miyamoto, H. (1982) Effects of ambient temperature on aspects of airborne  "
21            polycyclic aromatic hydrocarbons. Environ. Sci.  Technol. 16: 189-194.
22
23     Ye, H. (1993) Electrical technique. In: Willeke, K.; Baron,  P. A., eds.  Aerosol measurement principles,
24            techniques and applications. New York, NY: Van Nostrand Reinhold; pp. 410-426.
25
26     Ye, Y.; Tsai, C.; Pui, D.  Y. H.; Lewis, C.  W. (1991) Particle transmission characteristics of an annular
27            denuder ambient sampling system. Aerosol Sci. Technol.  14: 102-111.
28
29     Zhang, X!;  McMurry, P. H. (1992) Evaporative losses of fine paniculate nitrates during sampling. Atmos
30            Environ.  Part A 26: 3305-3312.
31
32     Zielinska, B.; Arey, J.;  Atkinson, R.; Winer, A. M. (1989) The nitroarenes of molecular  weight 247 in ambient
33            paniculate samples collected in southern California. Atmos. Environ. 23: 223-229.
34
35     Zoller, W. H.;  Gordon,  G. E. (1970) Instrumental neutron activation analysis of atmospheric  pollutants utilizing
36            Ge(Li) 7-ray detectors. Anal.  Chem. 42: 257-265.
        April 1995                                    4_12i        DRAFT-DO NOT QUOTE OR CITE

-------
  i                     5.   SOURCES  AND  EMISSIONS OF
  2                            SUSPENDED PARTICLES
  3
  4
  5      5.1  INTRODUCTION
  6           Excessive particulate concentrations in the atmosphere result from pollutant emissions,
  7      meteorological transport of those emissions between source and receptor, and chemical and
  8      physical changes during the transport period.  Importantly, both primary and secondary
  9      particles contribute to ambient PM mass concentrations.
 10           Primary particles are those which are directly emitted by sources,  and these particles
 11      often undergo few changes between source and receptor.  Atmospheric concentrations of
 12      primary particles are, on average, proportional to the quantities that are emitted.  Primary
 13      particles are emitted  in several size ranges, the most common being less than 1 ^m in
 14      aerodynamic diameter from combustion sources and larger than 1 /zm in aerodynamic
 15      diameter from dust sources.  Particles larger than 10 ^im in aerodynamic diameter usually
 16      deposit to the surface within a few hours after being emitted  and do not have a large effect
               »
 17      on urban or regional  scales.  These larger  particles may make significant contributions at
 18      receptors located within a few kilometers of the emissions  source,  however.  Coarse particles
 19      with aerodynamic diameters between 2 and 10 /xm may make significant contributions at
 20      distances exceeding 10 km from their emissions sources, while particles with aerodynamic
 21      diameters  < 2 /xm may affect receptors that are more than 100 km distant from emissions
 22      sources.
 23           Secondary  particles are those that form in the atmosphere from gases that are directly
24      emitted by  sources.  Sulfur dioxide, ammonia, and oxides of nitrogen are the precursors for
25      sulfuric acid, ammonium bisulfate, ammonium sulfate, and ammonium nitrate particles
26      (Seinfeld,  1986).  Several volatile organic compounds  (VOC) may also change into particles;
27      the majority of these  transformations result from intense photochemical reactions that also
28      create high ozone levels (Grosjean and Seinfeld, 1989). Secondary particles usually  form
29      over several hours or days, attain aerodynamic  diameters between 0.1 and  1 /xm, owing to
30      the complex chemical processes that form them, secondary particle concentrations are not
31      necessarily  proportional to the quantities that are emitted, and may affect receptors more  than
        April 1995                               5-1       DRAFT-DO  NOT QUOTE OR CITE

-------
 1      1,000 km distant from the sources of precursor gas emissions.  Several of these particles,
 2      notably those containing ammonium nitrate, are volatile and transfer mass between the gas
 3      and particle phase to maintain a chemical equilibrium.
 4           Ambient concentrations of secondary aerosol are not necessarily  proportional to primary
 5      emissions quantities since the rate at which they form may be limited by factors other than
 6      concentrations of the precursor gases.   For example,  secondary ammonium nitrate is not a
 7      stable compound; its equilibrium with gaseous ammonia and nitric acid is influenced strongly
 8      by temperature and  relative humidity (Watson et al.,  1994b).  Fugitive dust is predominantly
 9      a primary pollutant, but it does play a role in secondary particle formation.  Some
10      components of dust, such as ammonium nitrate fertilizer,  may volatilize into ammonia and
11      nitric acid gases, thereby contributing to secondary aerosol. Alkaline  particles,  such as
12      calcium carbonate, may react with nitric and hydrochloric acid gases while on the ground, in
13      the atmosphere, or on filter samples to form coarse-mode particle nitrates and chlorides.
14      Ammonium sulfate fertilizer components and  minerals such as gypsum (calcium sulfate) may
15      be mistaken for secondary sulfates when PM10 filters are  chemically analyzed.
16           Due to the complex and varying composition of suspended particles, it is necessary to
17      consider  sulfur dioxide,  oxides of nitrogen, ammonia, VOCs,  and primary particle emissions
18      as potential sources  of suspended  particles in  ambient air.  Major sources of these emissions
19      are classified  into categories of: (1) major point sources; (2) mobile sources; and  (3) area
20      sources (U.S.  Environmental Protection Agency, 1993).
21           Major point sources are ducted emissions that are subject to permit.  Emissions of a
22      single pollutant usually must exceed 25 tons/year to qualify for this category, though some
23      jurisdictions track individual thresholds at lower levels. This  category includes  stack
24      emissions from most industrial facilities in the United States such as steel mills, smelters,
25      cement plants, electric utilities, refineries, and incinerators. Non-ducted emissions from
26      these industries, such  as VOCs from leaking valves and primary particles from materials
27      handling  are termed "fugitive emissions".
28           Mobile sources include on-road and off-road motor vehicles, trains, aircraft, and ships,
29      with the majority of emissions resulting from on-road vehicles in most areas. Area sources
30      include many small  stationary emitters that, in their aggregate, can be significant  contributors
31      to suspended particles. These include residential wood and coal combustion, prescribed

        April  1995                                5-2       DRAFT-DO NOT QUOTE OR CITE

-------
  1      burning, space heating, cooking, paved and unpaved roads, construction and demolition,
  2      agricultural operations, and wind erosion.
  3           Point, mobile, and area source emissions are anthropogenic.  Natural sources of
  4      suspended particles such as sea salt, volcanic emissions, wild fires, and aeolian dust from
  5      undisturbed surfaces can also contribute to ambient PM concentrations in certain situations.
  6      One of the major challenges to air quality sciences is to distinguish among contributions from
  7      these different sources when suspended particle concentrations are high.
  8           This chapter is organized to present first a concise summary of key information on PM
  9      emissions  derived from the previous criteria review in the 1980's and then to provide a more
10      extensive discussion of newer information appearing in recent years.
11           The main objectives of the ensuing discussion this section are:
12           •     To identify the sources that are major contributors to suspended  particle
13                 concentrations in the United States.
14
15           •     To describe the particle sizes and chemical properties of source emissions.
16
17           •     To evaluate the limitations and uncertainties of emissions rate estimates and
18                 source contributions for suspended particles and their gaseous precursors.
19
20
21      5.2  SUMMARY OF 1982 CRITERIA DOCUMENT EMISSIONS
22           REVIEW
23           The U.S. Environmental Protection Agency  (EPA) (1982) examined 109 open-literature
24      references related to particulate emissions published between 1951 and 1981.  This
25      comprehensive review discussed the sources of emissions data and their accuracy, global
26      particulate emissions from natural sources, U.S. particulate and sulfur dioxide emissions
27      from manmade sources, and the size and chemical composition of emitted particles.  The
28      major findings of this review were  the following:
29
30           • Particulate emissions rates are very uncertain.  Two independent EPA estimates
31             yielded 6.4 (U.S.  Environmental Protection Agency, 1980a) and 3.9  (U.S.
32             Environmental Protection Agency,  1980b) million metric tons per year for the same
33             industrial process emissions of primary particles in 1977.  Particle emissions
34             estimates from mobile sources were believed to have even greater uncertainty.
35
       April 1995                                5.3       DRAFT-DO NOT QUOTE OR CITE

-------
 1           • Global participate emissions include 100 to 500 million metric tons/year (Robinson
 2             and Robbins, 1971; Vandegrift et al.,  1971) from aeolian dust, 900 million metric
 3             tons/year from sea spray (Robinson and Robbins, 1971), and 0.75 to 3.75  million
 4             tons/year from volcanoes (Robinson and Robbins, 1971; Granat et al., 1976). Much
 5             of these emissions are small particles that do not  attain great elevation above ground
 6             level and deposit close to their emissions points.  Significant fractions of volcanic
 7             emissions, however, are injected high  into the atmosphere and can have long
 8             atmospheric residence times.  Primary particle  emissions from U.S. wildfires were
 9             estimated to be 0.5 to 4.5 million metric tons/year, with most of these emissions
10             being small particles that do not deposit close to the emissions point (Robinson and
11             Robbins, 1971;  Yamate,  1973).
12
13           • Naturally-emitted volatile organic compounds that might be  secondary paniculate
14             precursors were estimated at 200 million metric tons/year for the world and
15             20 million metric tons/year for the United States  (Went, 1960).  Global emissions of
16             reduced sulfur compounds were estimated at 37 to 91 million metric tons per year
17             (Robinson and Robbins, 1971; Granat et al., 1976).
18
19           • Manmade particle emissions in the United States during 1978 included 10.5 million
20             metric tons/year from stationary point sources,  3.3  million metric tons/year from
21             non-ducted industrial processes, 110 to 370 tons/year from non-industrial fugitive
22             dust, and 1.3 million metric tons/year from  mobile sources.  Stationary point sources,
23             primarily electric utilities, accounted for 26.2 million tons/ year of sulfur dioxide
24             emissions, with the remaining 0.8 million tons/year emitted  by mobile sources.
25
26           • Primary particle emissions from stationary fuel combustion, industrial processes,
27             solid waste disposal, mobile sources, and burning in the United States decreased from
28             24.8 million metric tons per year  in 1940 to 12.5 million metric tons/year  in 1978.
29             All categories decreased expect emissions from transportation, which nearly tripled
30             from 0.5 to  1.3 million metric tons/year over this period. Sulfur dioxide emissions
31             increased from 19.5 to 27.0 million metric tons/year between  1940 and 1978, with
32             the major increase in stationary fuel combustion.
33
34           • Coal combustion was the major U.S. industrial emitter, with 3,090 out of  10,460
35             metric tons/year total for primary particles and 17,890 out of 26,180 metric tons  per
36             year total for sulfur dioxide during 1978. More than 80% of coal combustion
37             emissions derived from electricity generation.
38
39           • Emissions estimates differed substantially among  different parts  of the  United States,
40             especially between the eastern and western regions.  The midwest contained more
41             than 70% of particle and sulfur dioxide emissions from industrial sources.
42
43           • Fugitive dust source emissions were found primarily in particles sizes larger than
44             2.5 /mi.  Seventy to 100%  of the primary particle emissions from coal and oil
45             combustion and other ducted industrial emissions were smaller than 2.5 /j.m in
46             aerodynamic diameter.
47

        April  1995                                5.4       DRAFT-DO NOT QUOTE OR CITE

-------
 1          • Aluminum, silicon, calcium, potassium, and titanium were found to be abundant in
 2            many geologically-related emissions.  Several sources of primary particles showed
 3            enrichments with respect to crustal abundances for certain elements:  (1) copper,
 4            lead, and arsenic in smelter emissions; (2) selenium and arsenic in coal combustion
 5            emissions; (3) iron and manganese in steel mill emissions; (4) calcium and sulfate in
 6            cement emissions; and (5)  vanadium and nickel in residual oil combustion emissions.
 7
 8          The U.S. EPA (1982) emphasized emissions from industrial sources, and especially
 9     primary particles emitted by these sources.   Sulfur dioxide was the only precursor of
10     secondary aerosol that was considered.  Emission rates were quantified in terms of Total
11     Suspended Paniculate (TSP), i.e., particles with aerodynamic diameters < -50 /im,  since
12     PM10 had not  yet been defined as  the size fraction relevant to public health.  Measured
13     chemical compositions included many elements and sulfate, but did not include nitrate,
14     ammonium, and carbon.
15          At the time, this emphasis was appropriate owing to the then available information and
16     to the lack of  sufficient emissions  controls on many of these sources.   Since that time, many
17     ducted emissions have been controlled, yet air quality standards are still exceeded in many
18     areas.  Non-sulfate secondary aerosol is a major component of suspended particles in  many
19     areas. *The particle sizes relevant  to health effects are now believed to be much smaller than
20     50 /-on, and possibly substantially  smaller than the PM10 fraction which is currently subject to
21     NAAQS regulation.  Much new information has been published since  1982 that enhances and
22     expands upon  the conclusions about source contributions  to suspended particles.
23
24
25     5.3    SOURCE CONTRIBUTIONS TO SUSPENDED PARTICLES
26          Much of the current knowledge related to source contributions derives from evaluations
27     used in development of control strategies for non-attainment areas where PM10 concentrations
28     exceed 50 jug/m3 for an annual arithmetic average and/or 150 /ng/m3  for a 24-h average.
29     Seventy-five areas have been designated in non-attainment.  The 1990 Clean Air Act
30     Amendments and several years of ambient PM10 monitoring resulted in designation of
31     75 U.S. areas as "moderate" non-attainment areas for PM10 (Federal  Register, 1991;  1994),
32     5  areas as "serious" non-attainment (Federal Register, 1993), and the remaining areas as
33     unclassifiable.  Each state containing non-attainment areas must develop and submit State

       April 1995                                5-5       DRAFT-DO NOT QUOTE OR CITE

-------
 1     Implementation Plans (SIPs) that specify the means for reducing PM10 concentrations below
 2     the standards.  Extensive efforts were expended in many areas to apportion excessive PM10
 3     concentrations  to their sources.
 4           The U.S. EPA (1987) recommends the application of both source and receptor models
 5     to quantify the major contributors to excess PM10 concentrations, and linear rollback of
 6     emissions to estimate the effects of alternative emission controls (Pace and Watson,  1987;
 7     U.S. Environmental Protection Agency, 1987). Source models use emissions inventories and
 8     meteorological measurements to predict the PM10 concentrations measured at  receptors.
 9     Receptor models use the chemical  composition of source emissions and receptor
10     concentrations  to infer the  source contributions that constitute the measured PM10 mass.
11     Source models are most useful when sources  have been identified and emissions rates are
12     known.  In many non-attainment areas, however,  the majority of emissions emanate from
13     mobile and area sources that have  highly sporadic and often unknown emissions at different
14     locations and different times.  Several non-attainment areas in the mountainous western
15     United States experience highly variable meteorology induced by complex terrain, and most
16     dispersion models are not applicable.  In these situations, receptor-oriented source
17     apportionment models were found to the be the most appropriate methods to identify sources
18     and quantify their contributions to measured PM10.
19           Source contributions  to average PM10 for several areas where the Chemical  Mass
20     Balance  (CMB) receptor model (Watson et al., 1990;  1991) was applied are shown in
21     Table 5-1.  The values  in Table 5-1 are not entirely comparable in an absolute sense, since
22     published source apportionment studies usually report averages for different sample  selections
23     (often biased toward high PM10 levels) and different seasons.  The sampling sites represent a
24     variety of different source  characteristics within different regions of Arizona,  California,
25     Colorado, Idaho, Illinois, Nevada  and Ohio.   Several of these are background sites,
26     specifically Estrella  Park,  Gunnery Range, Pinnacle Peak, and Corona  de Tucson, AZ,
27     Anacapa Island, CA, San Nicolas  Island, CA, Vandenberg Air Force Base, CA, and Verdi,
28     NV.  Definitions of source categories also vary from study to study. In  spite of these
29     differences,  several  features can be observed from the values in this  table.
30           Fugitive dust (geological material) from roads, agriculture and  erosion is a major
31     contributor to PM10 at  nearly  all sampling sites, often contributing up to, but not generally

       April 1995                                5-6        DRAFT-DO NOT QUOTE OR CITE

-------
TABLE 5-1.  RECEPTOR MODEL SOURCE CONTRIBUTIONS TO PM10.
t— '•
I—1
VO















1
-J



O
!>
^
H
6
o

o
H
O
e;
O
H
tn
§
o
H
m




Sampling Site
Central Phoenix, AZ (Chow et al , 1991)
Corona de Tucson, AZ (Chow et al., 1992a)
Craycroft. AZ (Chow et al., 1992a)
Downtown Tucson, AZ (Chow et al., 1992a)
Hayden 1, AZ (Garfietd) (Ryan et al , 1988)
Hayden 2, AZ (Jail) (Ryan et al , 1988)
Orange Grove, AZ (Chow et al., 1992a)
Phoenix, AZ (Estrella Park) (Chow et al , 1991)
Phoenix, AZ (Gunnery Rg.) (Chow et al , 1991)
Phoenix, AX (Pinnacle Pk.) (Chow et al , 1991)
Rillito, AZ (Thanukoset al., 1992)
Scottsdale, AZ (Chow et al.. 1991)
West Phoenix, AZ (Chow et al., 1991)


Anacapa Island, CA (Chow et al., 1994b)
Anaheim, CA (Gray et al., 1988)
Anaheim, CA (Summer) (Watson et al , 1994a)
Anaheim, CA (Fall) (Watson et al , 1994a)
Azusa, CA (Summer) (Watson et al , 1994a)
Bakersfield, CA (Magliano, 1988)
Bakerfield, CA (Chow el al , 1992b)
Burbank, CA (Gray et al , 1988)
Burbank, CA (Summer) (Watson et al , 1994a)
Burbank, CA (Fall) (Watson et al., 1994a)
Chula Vista 1, CA (Bayside) (Cooper et al., 1988)
Chula Vista 2, CA (De! Ray) (Cooper et a!., 1988)
Chula Vista 3, CA (Cooper et al., 1988)
Claremont, CA (Summer) (Watson et al , 1994a)
Crows Landing, CA (Chow et al , 1992h)







Primary
Geological
33.0
17.0
13.0
26 0
5.0
21 0
20.0
37.0
20 0
7 0
42.7
25.0
30.0


2.2
21.2
11 4
13 2
34 9
27.4
42 9
21.3
14.0
11.0
6.7
82
9.7
19 4
32 2







Primary
Construction
0.0
0.0
0.0
5 1
2 Ob
4 Ob
00
0.0
00
0 0
13 8b
00
00


0.0
0.0
00
0.0
0 0
3 0
1.6
00
0.0
0.0
0.0
0 3
03
0 0
!) 0





Primary*
Motor
Vehicle
Exhaust
25.0
1.6
8.3
14 0
00
0.0
15 0
10.0
55
2 9
1.2'
19.0
25.0


4 9
4.1'
8.5
37.2
15 9
5.5
7.7
6 I1
17.0
39.1
0.8
1.5
1 4
14 4
2 2






Primary
Vegetative
Burning
2.3
00
00
00
0.0
0.0
0 0
0.9
0.0
1 0
00
7.4
10.0


0.0
00
0.0
00
00
9.6'
65
0.0
0.0
00
0.0
0.0
0.0
0.0
3 4






Secondary
Ammonium
Sulfate
0.2
1 9
0.7
1 0
40
4 0
0 4
1.6
1 0
0.9
0.
0.6
0.4


3 4
7.0
9.0
3.7
11 4
5 6
5.5
7 2
12.4
3 1
7 5
8.9
8.2
9 5
9 y




/ig/m3

Secondary
Ammonium
Nitrate
2.8
0.0
0.6
0.2
0.0
0.0
0 4
0.0
00
0 0
0.0
3.6
3.1


1 0
9.8
2.9
38 5
6 1
0.0
12.7
10.2
6 5
25 1
00
00
0.0
6 3
6 5







Misc.
Source 1
0.0
00
1.2*
1 3a
74 Oc
28 (f
0.0
0.0
0.0
0 0
11.6s
0.0
00


9.6h
0 41
OO1
00'
0 O1
0.5J
1 Om
0.1J
0.0'
0.0-'
0.41
0.61
0.6>
0 0'
0 5m







Misc.
Source 2
0.0
00
00
0.0
5.0d
00
00
0.0
0.0
0.0
00
0.0
0.0


0.0
1.4h
6.5"
3 lh
5 7h
0.0
1.5"
0911
5.7h
1.9h
2.7h
1.8h
1.7h
47h
1 5n







Misc.
Source 3
0.0
00
0.0
0.0
1.0e
1 Oc
0.0
0.0
0.0
0.0
0.0
0.0
0.0


00
8.2k
0.0
0.0
0.0
0.0
0.6k
9.8k
0.0
0.0
2.0k
0.0
- o.o
00
1.2k







Misc.
Source 4
0.0
0.0
00
00
00
0.0
0 0
0.0
0 0
0.0
0.0
0.0
0.0


0.0
0.0
00
0.0
0.0
0.0
00
00
0.0
0.0
0.0
0.0
0.0
0.0
0 0






Measured
PM10
Concentration
64.0
19.1
23 4
48.0
105 0
59 0
34 2
55 0
27.0
12 0
79 5
55.0
690


26.0
52 1
51 3
104 0
92 1
67.6
79.6
56 6
72.3
94.8
28 8
31 1
29.6
70 0
52 5





-------
TABLE 5-1 (cont'd).  RECEPTOR MODEL SOURCE CONTRIBUTIONS TO PM
                                                              10.
N±>'
>o
^o
Ul














i
00


o

>
H
6
o
z
o
H
O

o
H
tn
O
O
H
W




Sampling Site
Downtown Los Angeles, CA (Gray et al., 1988)
Downtown Los Angeles, CA (Summer) (Watson et al ,
1994a)
Downtown Los Angeles, CA (Fall) (Watson et al ,
1994a)
Fellows, CA (Chow et al., 1992b)
Fresno, CA (Magliano, 1988)
Fresno, CA (Chow et al., 1992b)
Hawthorne, CA (Summer) (Watson et al , 1994a)
Hawthorne, CA (Fall) (Watson et al., 1994a)
Indio, CA (Kim et al., 1992)
Kern Wildlife Refuge, CA (Chow et al., 1992b)
Lennox, CA (Gray et al., 1988)
Long Beach, CA (Gray et al., 1988)
Long Beach, CA (Summer) (Watson et al., 1994a)
Long Beach, CA (Fall) (Watson et al., 1994a)
Magnolia, CA (Chow et al., 1992c)
Palm Springs, CA (Kim et al., 1992)
Riverside, CA (Chow et al., 1992c)
Rubidoux, CA (Gray et al., 1988)
Rubidoux, CA (Summer) (Watson et al., 1994a)
Rubidoux, CA (Fall) (Watson et al., 1994a)
Rubidoux, CA (Chow et al., 1992c)
San Jose, CA (4th St.) (Chow et al., 1994a)
San Jose, CA (San Carlos St.) (Chow et al., 1994a)
San Nicolas Island, CA (Summer) (Watson et al , 1994a)
Santa Barbara, CA (Chow et al., 1994b)
Santa Barbara, CA (GTC) (Chow et al , 1994b)
Santa Maria, CA (Chow et al., 1994b)







Primary
Geological
23.8
12.7

9.4

29.0
17.1
31.8
7.5
8.9
33.0
15.1
16.0
20.7
11.1
11.3
31.7
16.4
32.6
43.1
34.9
19.2
48.0
13.1
11.8
1.6
9.5
3.2
7.4







Primary
Construction
0.0
0.0

0.0

1.4
07
0.0
0.0
0.0
3.0
2.0
0.1
0.0
0.0
0.0
0.0
1.4
0.0
4.01
4 5
16.1
00
0.0
00
0.0
0.0
0.0
0.0





Prirfiary
Motor
Vehicle
Exhaust
6.4'
16.2

41.1

2.1
4 0
6.8
5.6
35.1
4.4
2.2
4.6'
5.1'
6.3
42.8
11.2
2.3
7 0
5.6'
17.3
30.3
10.2
92
8.9
0.9
14 7
5.1
7.6






Primary
Vegetative
Burning
0.0
0.0

00

3.4
9.21
5.1
0.0
0 0
7 1
4.0
0.0
0.0
0.0
0.0
0.0
5.1
0.0
0.0
00
0.0
00
31.3
31.3
0.0
0.0
0.0
00






Secondary
Ammonium
Sulfate
7.6
13.0

3.9

5.1
1.8
3.6
15.0
5.1
3.6
3 3
7.6
8.0
10.9
3.8
4.9
3 7
4.8
6.4
9 5
2 1
5.3
2.3
2.1
3.7
3.2
2 8
3 1




//g/m3

Secondary
Ammonium
Nitrate
11.2
4.4

27.5

7 5
0.0
104
0.6
204
4.1
1.5
7.9
9.2
0.8
23.2
19.7
4.2
21.4
21.3
27.4
31.6
21 7
13 3
12.8
0 5
1 0
0 5
1 4







Misc
Source 1
0.0
00>

O.QJ

7.0m
0 1J
0.3m
O.O1
O.O1
0.2J
0.5m
02J
0.1J
0.1J
O.O1
0.3J
O.V
o.y
oy
O.QJ
O.QJ
0.41
0.9h
0.7h
00'
6.4h
63h
5.7h







Misc.
Source 2
1.3"
6.5h

1.8h

1.4"
0.0
1.0"
7.0h
3.7h
1.0h
1.5"
3.1h
2.0h
2.2h
2.7h
1.2"
0.5h
1.3h
1.0"
5.1"
1.1"
1.5h
00
0.0
4.3h
0.0 .
0.0
00







Misc.
Source 3
7.9k
0.0

0.0

1.4k
00
O.lk
0.0
0.0
00
0.7k
76k
6.4k
0.0
0.0
1.2°
0.0
1 1°
5.9k
0.0
0.0
5.7°
0.0
0.0
0.0
0.0
0.0
0.0







Misc.
Source 4
0.0
0.0

0.0

0.0
0.0
0.0
0.0
00
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
00
0.0
0.0
0.0
00
00
0.0
0.0
0.0
0.0






Measured
PM10
Concentration
60.2
67 6

98.6

546
48.1
71.5
45.9
85.1
58 0
47.8
46.9
51 9
46.1
96 1
66.0
35.1
64.0
87.4
114.8
112.0
87.0
684
64.9
17 4
34.0
20 5
27 0





-------
TABLE 5-1 (cont'd). RECEPTOR MODEL SOURCE CONTRIBUTIONS TO PM10.
Pj
h— *
VO







Primary


^ Motor Primary Secondary











vb



O
3
H
t
o
o
2;
o
H
O
cl
O
m




H
w

Sampling Site
Santa Ynez, CA (Chow et al., 1994b)
Stockton, CA (Chow et al., 1992b)
Upland, CA (Gray et al., 1988)
Vandenberg AFB, CA (Watt Road) (Chow et al., 1994b)
Telluride 1, CO (Central) (Dresser and Baird, 1988)
Telluride 2, CO (Society Turn) (Dresser amd Baird, 1988)
Pocatello, ID (Houck et al , 1992)
S. Chicago, IL (Hopke et al , 1988)
S.E. Chicago, IL (Vermette et al., 1992)
Reno, NV (Non-sweeping) (Chow et al , 1990)
Reno, NV (Sweeping) (Chow et al., 1990)
Reno, NV(Chow et al., 1988)
Sparks, NV (Chow et al., 1988)
Verdi, NV (Chow et al., 1988)
Follansbee, OH (Skidmore et al., 1992)
Mingo, OH (Skidmore et al , 1992)
Sewage Plant, OH (Skidmore et al., 1992)
Steubenville, OH (Skidmore et al., 1992)
WTOV Tower, OH (Skidmore et al., 1992)

aSmelter background aerosol.
bCement plant sources, including kiln stacks, gypsum pile, and
cCopperore.
dCopper tailings.
'Copper smelter building
fHeavy-duty diesel exhaust emission.
Background aerosol.
hManne aerosol, road salt, and sea salt plus sodium nitrate
'Motor vehicle exhuast from diesel and leaded gasoline.
Primary
Geological
4.6
34 4
25 4
4.5
32.0
12 1
8.3
27.2
14 7V
9.7
11 8
14 9
15 1
7.8
10.0
120
22.0
8.3
7.4


kiln area.






Primary
Construction
00
0.5
0.*
0.0
00
00
7.5,
2 4
00
0.0
0.0
0.0
0.0
00
0.0
0.0
0.0
00
00


Vehicle Vegetative Ammonium
Exhaust Burning
6.8 0.0
5.2 48
41' 00
3.2 0.0
00 98.7
00 73
0 1 0.0
2.8 0.0
09f 00
8.7 0.1
11.0 1.2
10.0 1.9
11.6 134
4.0 1 1
35.0 0.0
14.0 4.1
12.0 0.0
14.0 0 8
160 0.2

JResidual oil combustion
Sulfate
2 2
3.1
6 4
1.9
0 0
0 0
0 0
15.4s
7 7
06
0 8
1 3
2.7
0 9
160
15 0
13.0
14.0
15.0


/xg/m3

Secondary
Ammonium
Nitrate
0.6
7.0
14.5
1 0
0.0
0.0
00
00
0.0
0.2
0.2
0.6
0.9
0 1
-
--

--
-


^Secondary organic carbon.






'Biomass burning
mPnmary crude oil
"NaCl + NaNO3
°Lime
PRoad sanding material
^Asphalt industry.



Misc.
Source 1
4.0h
0.7m
0.6)
9 3h
61. 3p
7.3p
00
15.1'
08'
0.0
0.0
0.0
0.0
0.0
9.3'
3.4'
6.6'
3.8'
3.4'

sRegional



Misc.
Source 2
0.0
1.8"
0.6h
0.0
00
00
00
2.2"
0.3h
0.0
0.0
0.0
0.0
00
0.0
11 Ox
8.7X
5.0X
7.9X

sulfate.



Misc.
Source 3
0.0
0.0"
7.8k
0.0
0.0
0.0
84 lr
00
i.r
0.0
0.0
0.0
0.2k
0.0
0.0
0.0
0.0
0.0
0.0





Misc.
Source 4
0.0
0.0
0.0
0.0
0.0
0.0
0 0
0.0
7.78
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0




Measured
PM10
Concentration
19.0
62.4
58.0
20.6
208 0
27.0
100.0
80.1
41 0
20.4
24 9
300
41.0
15.0
66.0
60.0
62.0
46.0
490


'Steel mills.
"Refuse incinerator.
"Local road dust, coal
yard road
dust, steel
haul road dust.
"Incineration.
"Unexplained mass.














rPhosphorus/phosphate industry

-------
  1      more than 50% of the average PM10 mass concentration.  The average fugitive dust source
  2      contribution is highly variable among sampling sites within the same urban areas, as seen by
  3      differences between the Central Phoenix (33 ng/m3) and Scottsdale (25 /zg/m3) in Arizona.
  4     • It is also highly variable between seasons, as evidence by the summer and winter
  5      contributions at Rubidoux, CA. In general, these studies found that fugitive dust was
  6      chemically similar, even though it came from different emitters,  so that further
  7      apportionment into sub-categories was not possible. An exception was for road sanding in
  8      Telluride, CO.  Road sand often contains salts that allow it to be distinguished from other
  9      fugitive dust sources.  It is usually the only exposed fugitive dust source when other sources
10      are covered by snowpack.  Dust from some construction activities and  cement plants can also
11      be separated from other sources due to enrichments in calcium content  of these emissions, as
12      seen in studies at Rubidoux, CA and Rillito, AZ (near cement plants),  in Pocatello, ID (near
13      chemical and fertilizer production plants), and Tucson,  AZ (where a nearby community
14      center was undergoing renovation).
15           Primary motor vehicle exhaust contributions account for up to  approximately 40% of
16      average PM10 at many of the sampling sites.  Vehicle exhaust contributions are also variable
17      at different sites within the same non-attainment area.  Vegetative biomass burning, which
             »
18      includes agricultural fires, wildfires, prescribed burning, and  residential wood combustion,
19      was found to be significant at residential sampling sites such as:  Craycroft, Scottsdale, and
20      West Phoenix, AZ; San Jose,  Fresno,  Bakersfield, and Stockton, CA; Telluride,  CO;  Sparks,
21      NV; and Mingo, OH.  The predominance of these contributions  during winter months and
22      the local rather than regional coverage indicates that residential wood combustion was  the
23      major sub-category, even though chemical profiles are too similar to separate residential
24      combustion from other vegetative burning sources.  For example, Chow et al. (1988)  show
25      substantial differences between the residential Sparks, NV and urban-commercial Reno, NV
26      burning contributions even though these sites are separated  by less than 10 km.
27           Sites near documented industrial  activity show evidence of  that activity, but not
28      necessarily from primary particles emitted by point sources.  Hayden, AZ, for example,
29      contains a large smelter, but the major smelter contributions appear  to  arise from fugitive
30      emissions of or copper tailings rather than stack emissions.  Secondary sulfate contributions
31      at Hayden were low, even though sulfur dioxide emissions  from  the stack were substantial

        April 1995                               5-10      DRAFT-DO NOT QUOTE OR CITE

-------
 1     during the time of the study. Fellows, CA, is in the midst of oilfields facilities that burn
 2     crude oil for tertiary oil extraction. The Follansbee, Mingo, Sewage Plant, Steubenville, and
 3     Tower sites in Ohio are all close to each other in the Ohio River Valley and show evidence
 4     of the widespread steel mill emissions in that  area.
 5          Marine aerosol is found, as expected, at coastal sites such as Long Beach, San Nicholas
 6     Island, and Anacapa Island, CA, but these contributions are relatively low compared to
 7     contributions from manmade sources.
 8          Of great  importance are the contributions from secondary ammonium sulfate and
 9     ammonium nitrate. These are especially noticeable at sites in California's San Joaquin
10     Valley (Bakersfield, Crows  Landing, Fellow,  Fresno, Kern Wildlife, and Stockton), in the
11     Los Angeles area, and in the Ohio River Valley. Nitrate was not measured at the Ohio sites,
12     but there was a large portion of unexplained mass in the CMB source apportionments that
13     could be composed in part by ammonium nitrate.
14          Other aerosol characterization and receptor model source apportionment studies have
15     been performed for PM10 and PM2 5 that could be added to Table 5-1.  The general
16     conclusions drawn from this table would not change substantially.
17
18
19     5.4  NATIONAL EMISSION RATES AND ANNUAL TRENDS
20          Figure 5-1 (U.S. Envrionmental Protection Agency, 1993) shows the primary PM10
21     emissions estimated for the period  of 1985 through 1992 using the National Trends data base.
22     PM10 fugitive  dust emissions were not estimated prior to 1985.  Figure 5-1 shows fugitive
23     dust from paved and unpaved roads, agricultural operations, construction, and soil erosion to
24     constitute —90%  of nationwide primary emissions. All of the emissions have remained
25     relatively constant over the 8-year  period except for those from soil erosion.
26          A more detailed geographical breakdown of the  erosion emissions shows that the
27     majority of wind erosion occurs in the  "dustbowl" region of the United States that includes
28     the Oklahoma  and Texas panhandles (Barnard and Stewart, 1992; Fryrear, 1992). Wind-
29     induced erosion estimates are also  highly influenced by annual precipitation and wind-speed
30     distributions.   Erosion emissions estimates during 1986, 1989, and 1991 were approximately
31     twice those determined in other years owing to changes in these variables.

       April 1995                              5-11       DRAFT-DO NOT QUOTE OR CITE

-------
           70
           60
           50  i
          .-40
          CO
                        £.v«Wi$~:

           30
           20
           10
                          I Non-Fugitive Dust Sources
                          g] Wind Erosion
                          fH Unpaved Roads
                          | Paved Roads
                          [31 Mining and Quarrying    |
                          | Construction
                        |  [33 Agricultural Tilling
                 1985   1986   1987    1988    1989   1990   1991    1992
                                          Year
       Figure 5-1.  Primary PM10 emissions estimated for 1983 to 1992.


 1           Figure 5-2 (U.S. Environmental Protection Agency, 1993) expands the non-fugitive
 2     dust portion of Figure 5-1 into the indicated sub-categories.   PM10 emissions from these have
 3     been extrapolated from early TSP estimates back to 1983.  The major non-fugitive dust
 4     emitters are other industrial processes (several of which include materials handling which are
 5     sources of industrial fugitive PM10) and exhaust from highway vehicles.  Each of these
 6     emitters is only  —2% of the total emissions noted in Table 5-1.  Fuel combustion from
 7     utilities, industrial,  and other sources together contribute between 1 to 2% to total primary
 8     particle emissions.  Solvent use and petroleum storage and transport are included for
 9     comparability to gaseous emissions inventories and have  no primary particle emissions.
10           Industrial fuel combustion emissions were reduced by one-third and other fuel
11     combustion emissions were  reduced by one half between 1983 and 1992, mostly owing to
       April 1995
5-12
DRAFT-DO NOT QUOTE OR CITE

-------
             6  r
                                                                          Miscellaneous
                                                                          Off-Highway Vehicles
                                                                          Highway Vehicles
                                                                          Waste Disposal and Recycling
                                                                          Storage and Transport
                                                                          Solvent Utilization
                                                                          Other Industrial Processes
                                                                          Petroleum and Related Industries
                                                                          Metals Processing
                                                                          Chemical and Allied Product Manufacture
                                                                          Other Fuel Conbustion
                                                                          Industrial Fuel Combustion
                                                                          Electnc Utility Fuel Combustion
                 1983 1984 1985  1986  1987 1988 1989 1990  1991  1992
                                        Year
        Figure 5-2.  Sub-categories of non-fugitive dust emissions, 1983 to 1992.
 1      increased use of natural gas and the addition of particle removal devices as part of source
 2      permitting.  On-highway vehicle emissions increased by 50%, despite lower emissions and
 3      better gas mileage in newer vehicles.  This is primarily due to large increases in the number
 4      of vehicle miles traveled.  To the greatest extent possible, U.S. EPA (1993) adjusted
 5      previous years emissions to conform to current emissions estimation methods, so the year-to-
 6      year changes reflect real changes in emissions rather than changes in the estimation methods.
 7           Figures 5-3 through 5-5 show national emissions for sulfur dioxide, oxides of nitrogen,
 8      and VOCs for the 1983 through 1992  period.  As  found by U.S.  EPA (1982), electric
 9      utilities account for the largest fraction of sulfur dioxide, nearly 70% of total  emissions.
10      These emissions have not changed substantially over the 10 years reported, and when
11      translated into metric  tons, they are lower than the 19.4 million short tons/year estimated for
        April 1995
5-13
DRAFT-DO NOT QUOTE OR CITE

-------
           25
           20
           15
         CO
         o

         I
         o
           10
                  |  Miscellaneous
                  j§  Off-Highway Vehicles
                  j§  Highway Vehicles
                  f  Waste Disposal and Recycling
                  §1  Storage and Transport
                  |  Solvent Utilization
                  I  ]  Other Industrial Processes
                  [7]  Petroleum and Related Industries
                  H  Metals Processing
                  |  Chemical and Allied Product Manufacture
                  B  Other Fuel Conbustion
                  |  Industrial Fuel Combustion
                  [3  Electric Utility Fuel Combustion
               1983  1984  1985 1986 1987  1988 1989 1990  1991  1992
                                       Year

        Figure 5-3.  National emissions of sulfur dioxide, 1983 to 1992.
 1      utilities in 1978 by U.S. EPA (1982).  This difference may be due to methodological
 2      changes. Emissions from industrial fuel combustion increased by approximately 20% from
 3      1983 to 1985, then leveled off at about 3.1 million short tons/year.  A similar increase
 4      between 1983 and 1985 is found for chemical manufacturing, with a leveling off at 0.42
 5      million short tons/year after 1985.  Sulfur dioxide emissions from highway vehicles increased
 6      by 60% between 1983 and 1992, while off-highway vehicle emissions decreased to a low of
 7      0.23 million short tons/year in 1986, then slowly increased to 0.27 million short tons per
 8      year by 1990.  Major sulfur dioxide emissions reductions are observed for petroleum
 9      processing and other industrial processes, with decreases of 40% to 50% over the  ten year
10      period.  In total, however, sulfur dioxide emissions estimates in 1992 are identical to those
11      found in 1983 at 22.73 million short tons/year.
        April 1995
5-14
DRAFT-DO NOT QUOTE OR CITE

-------
              25
              20
              15 ]
            8
            I
            o
              10
                      Miscellaneous
                      Off-Highway Vehicles
                      Highway Vehicles
                      Waste Disposal and Recycling
                      Storage and Transport
                      Solvent Utilization
                      Other Industnal Processes
                      Petroleum and Related Industries
                      Metals Processing
                      Chemical and Allied Product Manufacture
                      Other Fuel Conbustion
                      Industrial Fuel Combustion
                      Electric Utility Fuel Combustion
                  1983 1984 1985  1986  1987  1988  1989  1990 1991  1992
                                         Year
        Figure 5-4.  National emissions for oxides of nitrogen, 1983 to 1992.
 1           Figure 5-4 shows a less than 5 %  increase in total nitrogen oxides emissions over the ten
 2      year period.  Utility and motor vehicle emissions are about equal at —7.5 million short
 3      tons/year, and together these account for two-thirds of total emissions.  Industrial and other
 4      fuel combustion and off-highway vehicles account for nearly all of the remaining third.
 5      There is little change in any of the source categories  from year to year, with slight reductions
 6      for highway vehicles and slight increases for utility, industrial and other fuel combustion, and
 7      off-highway vehicles.
 8           Volatile organic compound (VOC) emissions in Figure 5-5 are dominated by highway
 9      vehicles and solvent use that account for nearly 60%  of total emissions.  Off-highway
10      vehicles, petroleum-related industries, chemical manufacturing, and petroleum storage and
11      transport account for most of the remaining amounts.  VOC emissions from highway vehicles
        April 1995
5-15
DRAFT-DO NOT QUOTE OR CITE

-------
                                                                         Miscellaneous
                                                                         Off-Highway Vehicles
                                                                         Highway Vehicles
                                                                         Waste Disposal and Recycling
                                                                         Storage and Transport
                                                                         Solvent Utilization
                                                                         Other Industrial Processes
                                                                         Petroleum and Related Industries
                                                                         Metals Processing
                                                                         Chemical and Allied Product Manufacture
                                                                         Other Fuel Conbustion
                                                                         Industrial Fuel Combustion
                                                                         Electric Utility Fuel Combustion
                  1983 1984 1985 1986 1987 1988 1989 1990 1991 1992
                                         Year
        Figure 5-5.  National emissions for volatile organic compounds, 1983 to 1992.


 1      were reduced by 40%, in spite of increased vehicle mileage.  Most of this is due to the
 2      presumed effectiveness of emissions controls on newer vehicles.  VOC emissions from
 3      petroleum industries were also reduced substantially, as were  those from miscellaneous
 4      sources.  Emissions from other categories increased slightly or remained approximately the
 5      same.  Although U.S. EPA (1993)  includes a category for "natural sources," there are
 6      entries of zero  emissions for all years  for the category and it has, therefore, been omitted
 7      from Figure  5-5.
 8           The spatially  and temporally averaged emissions rates presented above were compiled
 9      according to  standardized procedures and provide a good starting point for the identification
10      of potential contributors to ambient paniculate concentrations  which exceed the PM NAAQS.
        April 1995
5-16
DRAFT-DO NOT QUOTE OR CITE

-------
  1     These estimates are insufficient, however, to develop effective emissions reductions strategies
  2     for specific standard exceedances that might affect public health.
  3           Several potentially important source categories  are not included in current inventories.
  4     For example, Hildemann et al. (1991) identify charbroiling and frying of meat in homes and
  5     restaurants as a potential PM source neglected in previous PM10 inventories.  Their studies
  6     show almost all of these emissions to be in the PM2 5 size fraction.  Emissions rates
  7     measured for regular and extra-lean hamburger meat which was charbroiled or fried on a
  8     restaurant-style grill (with commonly used grease traps) were:  40 g/kg for charbroiled
  9     regular meat; 7.1 g/kg for charbroiled extra-lean meat; 1.1 g/kg for fried regular meat; and
 10     1.4 g/kg for fried extra-lean meat.  Fugitive dust emitters such as golf course turf
 11     replacement, feedlots and dairies, equestrian events,  off-road vehicle competitions,  and
 12     parking lot sweeping are not quantified. These emissions may be small  on a national basis
 13     but could be important in specific non-attainment areas. Vegetative burning sources are also
 14     omitted,  as are  potential precursor VOC emissions from biogenic sources. Lamb et al.
 15     (1993) estimate national VOC  emissions from a variety of vegetative sources (based on land
 16     use maps) to be =47 million short tons/year.  This is twice the emissions from all  other
 17     sources combined in Figure 5-5.
 18          As  noted above, ammonia is a major participant in atmospheric reactions which form
 19     ammonium sulfate and ammonium nitrate,  but it is not  included in the National Trends
 20     inventory.  Only a small fraction of total ammonia emissions are related to activity levels
 21      which are quantified by current emissions inventory methods. Russell and Cass (1986)
 22     attributed 52% of all ammonia emissions in the Los Angeles area to livestock; 23% to dogs,
 23      cats, and humans (allocated by population); 15% to bare soil surfaces; 9% to sewage
 24      treatment; and 5% to fertilizer applications.  Less than 4% of ammonia emissions were
 25      attributed to stationary and mobile fuel combustion sources.  For 27 European countries,
 26      Buijsman et al.  (1987) estimated that 81% of anthropogenic ammonia is emitted by livestock
 27      waste, 17% by fertilizers, and  less than 2% by industrial sources.
28           Annual averages do not reflect the seasonality of certain emissions.  Residential wood
29      burning in fireplaces and stoves, for example, is  a seasonal practice which usually reaches  its
30      peak during December in response to very  cold weather.  Cold weather also affects motor
       April 1995                                5_17       DRAFT-DO NOT QUOTE OR CITE

-------
 1      vehicle exhaust emissions, both in terms of chemical composition and emission rates (e.g.,
 2      Watson et al., 1990c).  Planting, harvesting,  and fertilizing and harvesting are also  seasonal.
 3           Several of the sources  in Figures 5-1  through 5-5 are episodic rather than continuous
 4      emitters.  This is especially  true of prescribed and structural fires and fugitive dust
 5      emissions.  For example, Engineering Science (1988) based  windblown dust estimates for
 6      Phoenix, AZ, on high wind speeds that occur from 5 to 10 times per year.  Although
 7      windblown dust emissions are low on an annual average, they are likely to be quite large
 8      during those few episodes when wind speeds  are high.
 9           Source categories are not adequately desegregated.  This is especially true of the mobile
10      sources, in which paniculate emissions from  several vehicle types and emissions points are
11      combined in emissions models.  Construction dust estimates, for example, are based on acres
12      under construction rather than on the individual processes involved in construction.  This
13      complicates the evaluation of specific control strategies which might target specific vehicles
14      or processes.
15           As shown in Table 5-1, there is substantial variation in contributions to suspended
16      particles even for nearby sources.  The national averages in Figures  5-1 through 5-5 do not
17      take into account the regionality and zones of influence of different emissions sources.
18           Finally, the emissions  factors applied to activity data are of questionable accuracy.  For
19      example, Engineering  Science (1988) took  many samples from paved and unpaved roadways
20      in Phoenix, AZ,  and they found silt loadings ranging from 0.02 to 0.75 grains/ft2.  An
21      average of these measurements was used to represent the loadings for all paved roads.  It  is
22      very  likely that the loadings of suspendable dust on paved roads  varies over  time in ways
23      which can never be known.
24           Several of these deficiencies may be unimportant, and  the source apportionments
25      provided by receptor modeling allows their importance to be evaluated. Others of these
26      deficiencies can, and should be, corrected by better emissions modeling methods.  It is
27      revealing, for example, that fugitive dust constitutes 90% of the  annual average emissions
28      inventory, but it seldom averages more than 50% of the contribution to average PM10
29      concentrations as evidenced in Table 5-1.  The contributions from primary motor vehicle
30      exhaust, residential wood combustion, and industrial sources are definitely underestimated by
31      the relative  emissions from these sources in the national emissions inventory. Some of these

        April 1995                                5-18       DRAFT-DO NOT QUOTE OR CITE

-------
 1     deficiencies,  such as fugitive dust emissions, are probably intractable and the best that can be
 2     done is to estimate the uncertainties in these emissions and to value the data accordingly
 3     when decisions are made.  These issues are examined in greater detail for major source
 4     categories in the next section.
 5
 6
 7     5.5   EMISSIONS PROCESSES AND ESTIMATION METHODS
 8          The national emissions estimates shown in Figures 5-1  through 5-5 must be considered
 9     within the context of their estimation methods.  Many of these methods are empirically
10     derived, rather than  process-related, and they are often extrapolated beyond the original
11     empirical data sets used to  derive them. The U.S. EPA (1982) did a credible job of
12     describing estimation methods for industrial and  natural sources, but it did not describe
13     estimation methods for emissions from area and mobile sources.  Since Table 5-1 shows that
14     fugitive dust and motor vehicle are major contributors to suspended particles in several areas,
15     this ensuing review will focus on methods to estimate such emissions.
16
17     5.5.1  Fugitive Dust
18          Fugitive dust consists of geological material that is suspended into the atmosphere by
19     natural wind and by  anthropogenic activities from sources such as paved and unpaved roads,
20     construction  and demolition of buildings and roads, storage piles, wind erosion, and
21     agricultural tilling. Though qualitative descriptions of fugitive dust emissions are easy to
22     understand, translating these descriptions into quantitative estimates of emission rates,
23     locations, temporal variability, and contributions to suspended particles measured at receptors
24     has been a scientific and engineering challenge.
25          The movement of soil, and especially its suspension from the surface of the earth into
26     the atmosphere, has  been studied in many branches of science.  "Aeolian dust,"  named for
27     Aeolia, the Greek goddess  of the wind, is a major discipline in the fields of geology  and
28     archaeology  (Pye, 1987). Wind erosion is of great concern  to agriculturists and soil
29     scientists.  Meteorologists study interactions between the atmosphere and the earth's surface.
30     Air pollution scientists devise methods to estimate contributions from dust suspended  by the
31     wind and other mechanisms to particles which might cause adverse health effects.  The

       April 1995                                5-19      DPxAFT-DO NOT QUOTE OR CITE

-------
 1      published literature on the mechanisms for dust suspension are widely dispersed and not
 2      entirely comparable in terms of the properties studied and the units of measure.  Much of the
 3      work on dust suspension has been done by agriculturists to minimize soil erosion. Erosion
 4      estimates include all of the mechanisms which might affect the removal of topsoil from  a
 5      given area, one of which is suspension into the atmosphere.
 6           Air pollution scientists are only concerned with that portion of eroded soil which is
 7      removed by suspension into the atmosphere and transported reasonable distances (typically
 8      greater than 100 m, the  nominal dimension of significant fugitive dust emissions sources)
 9      without deposition to the surface.   This suspension depends on: (1) particle size  of the
10      erodible material; (2)  surface loadings; (3) moisture; (4) surface roughness; (5) wind speed
11      and direction; (6) suspension height; and  (7) vehicular ejection mechanisms.  The current
12      status of knowledge of each of these characteristics  is discussed below.
13           Very little is known about the respirable size fractions in dust, despite their adverse
14      health potential, long residence times, and high potential for vertical mixing.  The  most
15      comprehensive information on particle  sizes in geological material is contained in soil
16      surveys compiled by the Soil Conservation Service (SCS).  These surveys provide detailed
17      boundaries  for different  soil types on 7.5-minute maps corresponding to U.S. Geological
18      Survey (USGS) maps.  The codes are associated with data in a printed summary which
19      accompanies the maps for each survey  area.
20           Particle sizes in soil surveys are indicated by qualitative descriptions in  terms of the
21      amount of sand (50 to 2,000 /*m geometric diameter), silt (2 to 50 jum geometric diameter),
22      and clay (< 2 /*m geometric diameter).  These particle size fractions in the soil survey  are
23      estimated by the individuals conducting the survey based on the visual  similarity of the
24      observed soils to a sub-set of soil samples which are submitted to particle size analysis in a
25      laboratory.  AP-42 emissions factors for fugitive dust often contain silt loading  as one of
26      their input variables largely because of the availability of SCS soil survey data for  large
27      portions of the United States.
28           The particle sizing procedure  American  Society for Testing and Materials,  1990a;
29      1990b) which is most  commonly followed for soil surveys creates a soil/water suspension in
30      which soil aggregates  are broken into their component parts prior to sieving.  While this
31      disaggregation is useful for agricultural, construction, and other land uses, it  is not especially

        April 1995                                5-20      DRAFT-DO NOT QUOTE OR CITE

-------
  1      useful for estimating air pollution emissions because it does not accurately estimate the size
  2      of aggregates which are suspended by surface winds.  Gillette (1980) applied two methods to
  3      determine the small particle content of soil which might be entrained by winds  and cause
  4      pollution problems. The first method (i.e., gentle sieve) consists of drying the soil sample
  5      and sieving it gently with about twenty circular gyrations parallel to the plane of the sieve
  6      (Chepil,  1952).  The second method (i.e., hard sieve) consists of up to one-half hour of
  7      vigorous shaking (usually using a shaking machine). The gentle sieve method best represents
  8      the suspension properties of the soil in the state in which it was  sampled.  The hard sieve
  9      method represents the potential of that soil for resuspension when disaggregating activities
10      (e.g., vehicle traffic) occur.  Gillette's threshold suspension velocity measurements apply to
11      soil characteristics obtained by  the gentle sieve method.
12           The size distribution of dust particles affects the suspension process.  A flat bed of
13      particles  with diameters less than 20 /xm is very difficult to suspend by wind, as Bagnold
14      (1937) showed by blowing wind in excess of 100 cm/sec over a bed of fine Portland cement.
15      In this situation, there is no large cross section for wind to act on.  A bed of large particles
16      with diameters exceeding  1 mm interspersed with fine  particles also mitigates suspension.
17      Particles  larger than 0.5 mm cannot be lifted by wind, but they absorb wind energy as they
18      roll along the surface.  They also shelter smaller particles on their lee sides from the effects
19      of high wind speeds. Gillette and Stockton (1989)  sprinkled glass spheres with diameters
20      ranging from 2.4 to 11.2 mm onto a bed of glass spheres with sizes from 0.107 to  0.575 /mi
21      and found major reductions in suspension  of the smaller particles. Bagnold (1941) estimated
22      that 80 jum particles are the most susceptible to suspension by wind, even though their large
23      masses cause them to settle to the surface  very rapidly.
24           The amount of suspendable dust on a surface influences how much might be suspended.
25      Most surfaces are limited reservoirs, and the suspendable dust is depleted after a short time
26      period.  On exposed land, fine  particle which  are blown away often expose larger rocks
27      which then shield the suspendable particles from the wind. When surfaces are  continually
28      disturbed, however, by very intense winds or by vehicular movement, they may become
29      unlimited reservoirs which emit dust whenever winds exceed threshold suspension velocities.
30      There are few (<  500 for the entire U.S.) reported data on the surface loadings of silt
        April 1995                                5_2i       DRAFT-DO NOT QUOTE OR CITE

-------
 1     (< 75 /nm) and absolutely no data on surface loadings of respirable particle sizes for any
 2     surface included in fugitive dust emissions inventories.
 3           Water adheres to individual  soil particles, thus increasing their mass and mitigating
 4     suspension and transport.  It also  increases the cohesive forces among individual particles,
 5     and these forces  often persist after the water has evaporated as evidenced by aggregates  and
 6     surface crusts.  Chepil (1956), Belly (1964), Bisal and Hsieh (1966), and Svasek and
 7     Terwindt (1974) show that substantially greater wind forces are needed  when soil surface
 8     moisture is increased by less than 1 % from its dry state.
 9           Soil surveys include  plastic limits and moisture limits to the ability of soils to absorb
10     moisture. The "plastic limit" is the moisture content at which a soil changes  from a semi-
11     solid to a plastic. The plastic limit is determined by adding water to dry soil sample until it
12     can be rolled into a coherent cylinder. Soil surveys also report liquid limits  (the quantity of
13     water required to create a slurry with the consistency of water), the infiltration rate (the
14     movement of water through soil layers), and field moisture capacity.
15           Kinsey and Cowherd (1992) show how watering might reduce emissions at a
16     construction site. A large pollution control benefit might be derived from  initially  doubling
17     the area which is watered with lower benefits achieved as more water is applied to the site.
18     Control efficiency is ultimately limited because grading operations are continually exposing
19     dry earth and burying the moistened topsoil.  A portion of this moistened soil adheres to the
20     construction vehicles and can be carried out to paved and unpaved roads for subsequent
21     resuspension.
22           While the moisture capacities and retentions of different geological materials  are well
23     documented in the soil surveys, the actual moisture content at a given time or place is not
24     recorded.  Thornthwaite (1931) proposed the ratio of precipitation to evaporation as an
25      indicator of the availability of moisture for soils. Thornthwaite's major concern was the
26     agricultural potential of land in different areas.  The precipitation-evaporation effectiveness
27      index (P-E index) is 10  times the sum of the monthly precipitation to evaporation ratios.
28      Using precipitation, evaporation,  and temperature data taken prior to 1921 at twenty-one
29      U.S. monitoring sites, Thornthwaite (1931) devised an empirical precipitation-evaporation
30      index (P-E index) to classify all North America as wet (P-E index  > 128), humid
31      (64 <  P-E index < 128), sub-humid (32 < P-E index  < 64), semi-arid  (16 < P-E index

        April 1995                                5-22       DRAFT-DO NOT QUOTE OR CITE

-------
  1      < 32), or arid (P-E index  <  16).  Much of the western United States is in the arid and
  2      semi-arid categories, while much of the eastern United States is in the humid category.  The
  3      P-E index is cited for several AP-42 emission factors to estimate the moisture content of
  4      different soils.
  5           Moisture content of soils will vary throughout the year depending on the frequency and
  6      intensity of precipitation events, irrigation, and relative humidity and temperature of the
  7      surrounding air. Large amounts of rain falling during 1 mo of a year will not be as effective
  8      in stabilizing dust as the same amount of rain  interspersed at intervals throughout the year.
  9           Moisture also  causes dust to adhere to vehicle surfaces so that it can be carried out of
10      unpaved roads, parking lots, and staging areas.  Carry out also occurs when trucks exit
11      heavily watered construction sites.  This dust is deposited on paved roadway surfaces as it
12      dries, where it is available  for suspension far from its point of origin. Fugitive dust
13      emissions from paved roads are often higher after rainstorms in areas where unpaved
14      accesses are abundant, even though the rain may  have flushed existing dust from many  of the
15      paved streets.
16           Windblown dust is a major contribution from all exposed  surfaces. Each surface has a
17      threshold velocity that depends on the cohesiveness of the particles and the surface
18      roughness.  Surface roughness height is the distance  above average ground level at which the
19      average wind velocity approaches zero.  Surface roughness is also a  meteorological concept;
20      it is related to, but not identical to, measured heights of obstructions in an open  area. Larger
21      surface roughness shelters suspendable dust from suspension. Surface roughness depends
22      both on the height and spacing of roughness elements, but is not well characterized for  most
23      surfaces.  Friction velocity is the slope of the  logarithm of wind speed versus elevation  above
24      ground level.
25           Dust arises due to suspension of the disturbed surface by wind.  Chepil and Woodruff
26      (1963) and Gillette and Hanson (1989) show that the amount of soil which can be suspended
27      by wind depends on the particle size distribution, wind velocity  at the soil surface, the
28      roughness of the surface, the relative fractions of erodible (< 2 /mi diameter) and non-
29      credible (>  2 jum diameter) material, and the  cohesion of the soil particles with one another.
30      Values for each of these variables affect other variables.  For example, a higher moisture
31      content increases cohesion among particles and shifts the size distribution to larger particles.

        April  1995                                 5-23      DRAFT-DO NOT QUOTE OR CITE

-------
 1     Larger agglomerations of small particles increase surface roughness which decreases wind
 2     speeds at the surface.
 3          All of these variables collectively affect threshold friction velocity, which is
 4     experimentally determined by placing a wind tunnel over an example of the affected soil and
 5     measuring the velocity near the surface at which visible soil movement is first observed.
 6     Wind speeds must approach 0 near the earth's surface, and experiments show that wind
 7     speed increases exponentially with height. Both friction velocity and surface roughness are
 8     determined experimentally for different situations by taking measurements at different
 9     elevations.  When the actual friction velocity is less than the threshold friction velocity for
10     soil erosion, no particles are suspended into  the atmosphere.  Most ambient wind speed
11     measurements are made at elevations between 5 and 10 m above ground level, and these
12     must be translated to surface friction velocities to determine suspension.  For a given
13     threshold friction velocity, paniculate emissions factors for windblown dust use the fastest
14     mile wind speed as reported in the National  Weather Service (NWS) Local Climatological
15     Summaries.  Gillette (1980)  shows that threshold wind speeds vary from 0.19 to  1.82 m/s  for
16     disturbed soils. Even though emissions are initiated at these velocities, the wind  force
17     contains insufficient energy to suspend much of the erodible soil mass.  The amount of dust
18     suspended increases at approximately the cube of the wind speed above the threshold
19     velocity.
20           Particles suspended into the atmosphere are acted upon by  gravity in a downward
21     direction and by atmospheric resistance in an upward direction.  Every particle attains an
22     equilibrium between these forces at its terminal settling velocity.  The settling velocity
23     increases as the square of the particle diameter or when the particle density increases.  For
24     very small particles (<  10 ptm diameter), vertical air movements caused by turbulence can
25     counteract the gravitational settling  velocity  and such particles can remain suspended for long
26     times.  Particle deposition for particles larger than  ~20 /*m diameter is dominated by the
27      force of gravity, however.  Transport distance depends on the initial elevation of a particle
28      above ground level, the horizontal wind velocity component in the direction of interest at the
29     particle elevation, and the gravitational settling velocity.  For this  reason, large TSP
30      emissions rates do not give a good  impression of contributions to suspended particles that  are
31      very distant from the receptor site.  Pye  (1987) shows vertical profiles for different sized

        April 1995                                 5-24       DRAFT-DO NOT QUOTE OR CITE

-------
  1      particles which might be elevated through a 100 m depth during a wind storm.  The particles
  2      smaller than 10 /*m are nearly uniformly distributed through this depth, while the larger
  3      particles exhibit much higher concentrations closer to the surface.
  4           Dust on paved roads, unpaved roads, parking lots, and construction sites is suspended
  5      by natural winds and vehicular movement.  Vehicular traffic in these areas adds to
  6      suspension because tire contact with the road lifts particles into the air.  Vehicle wakes
  7      create turbulent eddies which act much like natural winds to raise particles.  The grinding of
  8      particles by tires against the road surface shifts  the size distribution toward smaller particles,
  9      especially those in respirable size fractions.
 10           Unpaved roads and other unpaved areas with vehicular activity are unlimited reservoirs
 11      of dust loading when vehicles are moving. These surfaces are always being disturbed, and
 12      wind erosion seldom has an opportunity to increase their surface roughness sufficiently to
 13      evade particle  suspension.  The U.S. EPA AP-42 emission factor (U.S. Environmental
 14      Protection Agency, 1988) for unpaved road dust emissions contains  variables which account
 15      for silt loading, mean vehicle speed, mean vehicle weight, mean number of wheels, and
 16      number of days with detectable precipitation, to determine annual PM10 dust emissions for
 17      each vehicle-kilometer-traveled.  These relationships are derived from imprecise correlations
 18      of variables, however, and a full physical understanding of the vehicle  suspension process is
 19      lacking for unpaved roads. Muleski and Stevens (1992) note that more  than 90% of the tests
 20      which acquired data for the AP-42 factor were conducted with vehicle speeds slower than
 21      56 km/h (35 mph), and more than 80%  was derived from industrial haul roads. AP-42
 22      emission factors may not be applicable  to publicly maintained unpaved  roads, desert
 23      shortcuts, and  agricultural roads which are common in most PM10 non-attainment areas.
 24           Dust on paved roads must be continually replenished, however, and reducing the
25      deposition of fresh dust onto these surfaces is a  viable method for reducing particulate
26      emissions. Dust loadings  on a paved road surface build up by being tracked out from
27      unpaved areas  such as construction sites, unpaved roads, parking lots, and shoulders; by
28      spills from trucks carrying dirt and other particulate materials; by transport of dirt collected
29      on vehicle undercarriages; by wear of vehicle components such  as tires, brakes, clutches, and
30      exhaust system components; by wear of the pavement surface; by deposition of suspended
31      particles from many emissions sources; and by water and wind erosion from adjacent areas.

        April 1995                                5_25       DRAFT-DO NOT QUOTE OR CITE

-------
 1          The relative contribution from each of these sources is unknown.  Axetell and Zell
 2     (1977) estimated typical deposition of 240 Ib/mile of curb/day for particles of all sizes from
 3     the following sources:  (1) 42% from mud and dirt carryout;  (2) 17% from litter;
 4     (3) 8% from biological debris;  (4) 8% from ice control compounds (in areas with cold
 5     winters); (5) 8% from erosion of shoulders and adjacent areas; (6) 7%  from motor vehicles;
 6     (7) 4% from atmospheric  dustfall; (8) 4% from pavement wear; and  (9) less than 1% from
 7     spills.  These proportions  are highly uncertain because they apply to  the TSP size fraction
 8     (rather than to the respirable size fractions) and because these investigators did not consider
 9     all of the sources cited above.  Axetell and Zell (1977) cite these fractions without describing
10     the methodology used to estimate them.  No other published quantitative apportionments of
11     paved road dust loadings to their sources were found.
12          Nicholson et al. (1989) and Mollinger et al. (1993) identify the turbulent wake of
13     vehicles  on roads as a major cause of dust suspension from roads; their research suggests
14     that vehicle shapes might  be altered to reduce emissions potential.  Mollinger et al. (1993)
15     mounted a cylinder, an elliptical cylinder, and a rectangular solid on a pendulum which
16     swung back and forth over dust-covered test areas.  After 20  passes by the cylinder and
17     elliptical cylinder, 65% and 45% of the dust remained in the  test area, respectively.  After
18     20 passes by the rectangular solid traveling at the same velocity, less than 20% of the dust
19     remained.
20          Other than the information inferred from the chemical composition of road dust and
21     from multivariate relationships between downwind concentrations and vehicle variables, there
22     is no detailed physical understanding of the effects of tire contact with particles and their
23     suspension into the atmosphere.  This knowledge is essential  to understanding how these
24     particles are suspended  and how far they are transported.
25          There are obvious discrepancies between the proportion of fugitive dust in primary
26     emissions and geological  contributions to PM10 calculated by receptor  models.  To some
27     extent, this is due to contributions from secondary aerosols, which are not included in the
28     primary  PM10  emission estimates.  Even when secondary aerosol is subtracted, however,
29     other sources  such as vegetative burning and wood combustion make larger  relative
30     contributions to ambient concentrations than  is indicated by the emissions inventories.
        April 1995                               5-26       DRAFT-DO NOT QUOTE OR CITE

-------
 1          Fugitive dust estimates are especially affected by the general limitations of emissions
 2     inventories described above.  Annual and spatial averages do not reflect the seasonality of
 3     certain emissions. Planting and harvesting are seasonal, as are high winds which might cause
 4     erosion.  Paved road dust emissions might be much higher after rain storms when dirt is
 5     tracked from unpaved areas onto paved roads.
 6          Many fugitive dust sources are episodic rather than continuous emitters.  For example,
 7     Engineering Science (1988) based windblown dust estimates for Phoenix, AZ, on high wind
 8     speeds that occur 5 to 10 times per year. Though windblown dust emissions are low on an
 9     annual average, they can be quite large during those few episodes when wind speeds are
10     high. Construction activities are also episodic in nature.  Reeser et al. (1992) reported
11     fugitive dust emissions during wintertime in Denver, CO to be 44% higher than those found
12     in the annual inventory using standard emissions inventory methods.  In Coachella Valley,
13     CA, the South Coast Air Quality Management District (SCAQMD) (1994) calculated 24-h
14     emissions based on a worst windy  day.  When wind gust speeds exceeded 96 km/h,  fugitive
15     dust emissions could account for 20% of the entire annual average emission rates.
16          As noted earlier, many  fugitive dust emitters are included  in the inventories. For
17     example, Chow et al. (1992) identified two cement plants and many roads with unpaved
18     shoulders near the Rubidoux, CA,  site.  The cement plants were not included in the
19     SCAQMD emissions inventory, and there is no distinction in any of the inventories between
20     curbed and swept roads and those with no shoulders that may be dirtier than others.
21          Finally,  the spatial  disaggregation for fugitive dust emissions is poorer than that for all
22     other source categories.  Whereas most mobile sources are confined to established roadways
23     and most area sources correspond to population density, suspendable dust is everywhere.
24     Most fugitive  dust emissions  are compiled on a county-wide basis and are not allocated to
25     specific fields, streets, unpaved roads, and construction sites possibly contributing to high
26     airborne PM concentrations.  Several of these limitations may be impossible to overcome,
27     but many result from old methods  being applied to the problem. Modern data bases,
28     computer systems, and information management software can be applied to improve  existing
29     inventories without major additional costs beyond initial investments in establishing an
30     inventory methodology.
31

       April 1995                                5-27      DRAFT-DO NOT QUOTE OR CITE

-------
 1     5.5.2  Mobile Source Emissions
 2          Mobile sources are major emitters of primary particles, oxides of nitrogen, and VOCs.
 3     They are also minor emitters of sulfur dioxide and ammonia.  On-road motor vehicles using
 4     gasoline- and diesel-fueled engines are by far the largest component of mobile source
 5     emissions, and the emissions estimation methods are most highly developed for these
 6     vehicles. The U.S. EPA (1994) has published the MOBILE model to estimate carbon
 7     monoxide, oxides of nitrogen, VOC, and primary particle emissions from on-road vehicles.
 8     The other model in common use is the California Air Resources Board's Emissions  FACtor
 9     (EMFAC) model (CARB,  1993). These models use as their inputs variables such as vehicle
10     speed, vehicle age distribution, vehicle classification, ambient temperature, and laboratory
11     emissions test data. The laboratory test data generally use dynamometers with Federal Test
12     Procedure (FTP) driving cycles.
13          MOBILE and EMFAC model estimates of carbon monoxide, oxides of nitrogen, and
14     VOC emissions are variable as a function of the input variables.  Primary particle emissions
15     from tire and break wear and from exhaust are constant, however, regardless of assumed
16     driving conditions.  Emissions estimates from these models have recently been called into
17     question by  a series of on-road emissions measurement experiments, which consist of
18     measurements made in tunnels, along roadsides, and by pulling over vehicles for on-road
19     inspection and dynamometer testing.
20          The Southern California Air Quality Study (SCAQS) Van Nuys Tunnel Study (Ingalls et
21     al.,  1989), conducted in October and December, 1987 in the Sherman Way tunnel under the
22     Van Nuys Airport,  first called modeled emissions into question by noting that the emissions
23     models underestimated measured VOC emissions by nearly a factor of four. Nitrogen oxide
24     emissions in the tunnel, however, were consistent with those calculated by the models.
25     Pierson et al. (1990) concluded that the results were, in the main, correct. Even if the
26     absolute values of the emission factors determined were somewhat biased due to inaccurate
27     air flow measurements (one of the major difficulties pointed out by Pierson et al.) the
28     emission factor ratios are still valid.
29          Other tunnel experiments were carried out in the Ft. McHenry Tunnel in Baltimore,
30     Maryland, in June 1992  and in the Tuscarora Mountain Tunnel on Interstate 76 in south-
31     central Pennsylvania, in September 1992 (Pierson et al., 1995).  Both tunnels have

       April 1995                               5-28      DRAFT-DO NOT QUOTE OR CITE

-------
 1      considerable fetches of freeway before them; thus, all vehicles should have been in the hot-
 2      stabilized mode when they entered.  In both tunnels the traffic maintained steady speeds with
 3      low run-to-run and vehicle-to-vehicle variability.  This is in distinct contrast to the Van Nuys
 4      tunnel where there was considerable variation in vehicle speeds and many different speeds.
 5           Both MOBILE4.1 and MOBILES modeled running losses fell within the range of
 6      source contributions determined by the CMB model from VOC source profiles for tailpipe
 7      exhaust and gasoline headspace (Gertler et al., 1995)  For the driving mode in these tunnels,
 8      failure to model hydrocarbon emissions is not exclusively due to the method which estimates
 9      running loss evaporative emissions.  Fuel volatilities were lower at both Ft. McHenry and
10      Tuscarora than  during the Van Nuys tunnel experiment.  While this may lower evaporative
11      emissions, it is  not enough to account for the  discrepancy between the observed and modeled
12      emission  factors at Van Nuys.
13           The emissions models tend to overestimate at Tuscarora and only slightly underestimate
14      emissions at Ft. McHenry, which are reasonable results given the sources of the emission
15      factors in the models and traffic conditions at  each tunnel.  As stated previously, the models
16      derive their emission factors from the FTP, which has many accelerations and decelerations
17      and very  little steady-speed driving.  At Tuscarora, there was little acceleration or
18      deceleration.  The tunnel is virtually flat and contains no turns.  It is more than 10 km from
19      the nearest interchange and many, if not most, commuter vehicles travel for much longer
20      distances.  These vehicles were relatively new (median model year was 1989 during the 1992
21      experiment) and presumably well maintained.  The average speed of light-duty vehicles in the
22      tunnel (determined by a hand-held radar gun)  was  59.4 mph with a vehicle-to-vehicle
23      variability of 5.6 mph, which is negligible variability compared to that seen at the Van Nuys
24      tunnel.  Traffic  in the Tuscarora tunnel is expected to emit at low levels because of:
25      (1) a flat  roadway, (2) steady driving speeds,  (3) relatively  new and well maintained
26      vehicles,  and (4) vehicles in hot-stabilized driving mode. These conditions are consistent
27      with basic assumptions of the emissions models.
28           The Ft. McHenry tunnel is a different situation,  with up- and down-grades reaching
29      +3.76%  and more speed variability.  The vehicle  average speeds were 51 mph upon entering
30      the tunnel and 43 mph at the exit.  The median model year was 1989 for automobiles, and
31      given the  tunnel's location  on a major freeway, it is assumed that all vehicles were in hot-

        April 1995                                5_29       DRAFT-DO NOT QUOTE OR CITE

-------
 1      stabilized driving mode.  Though Ft. McHenry is a more complex driving situation than
 2      Tuscarora, it is still considerably less complex than the Van Nuys situation.  At Ft.
 3      McHenry, MOBILE4.1 underestimates and MOBILES overestimates emissions, but only
 4      slightly in either case. None of the discrepancies are as  severe as the underestimates at Van
 5      Nuys.  The models agree well with the Tuscarora and Ft. McHenry tunnel observations,
 6      even when deviations from assumptions such as road grades are significant, but when
 7      challenged by a complex urban driving situation such as  Van Nuys, it is uncertain at this
 8      time how well the models perform.
 9           Stedman et al. (1994) and Ashbaugh et al. (1992) conducted a study in the Los Angeles
10      area in 1991 using remote sensing devices for the detection of VOC and carbon monoxide
11      emissions from individual vehicles. During the study, more than 60,000 vehicles' VOC and
12      CO emissions were measured by remote sensing.  Of that group, more than 300 high-
13      emitting vehicles were identified by the remote sensors and pulled over for inspection.
14      Sixty-seven percent of the inspected vehicles had emission control  systems that were
15      defective or had been tampered with; more than 90% of them failed the Smog Check
16     inspection.
17           During the same time, approximately 80 of those inspected vehicles were given an
18     IM240 test (a 240 second, loaded mode test on a dynamometer).  These vehicles previously
19     had been stopped because of high remote sensing CO and/or VOC readings.  All but one of
20     the tested  vehicles  exceeded the California certification emission standards. The ten highest
21      CO and VOC  emitters had mass emission rates that were 24 to 70 times higher than the
22     standards.  NOX emissions also were measured in the IM240 tests, and three of the vehicles
23     had NOX mass emission rates greater than 10 g/mi.   Even though the vehicles had been
24     stopped because of their CO or VOC  remote sensing readings, high NOX emitters were
25     encountered.  This unexpected finding suggests  the presence of high NOX emitters in the fleet
26     (Knapp, 1992) and has important implications for the accuracy of NOX emissions in the
27     mobile source emissions inventory. Though particle emissions were not measured in these
28     experiments, it is likely that inefficient combustion and pollution controls on these vehicles
29     enhance the production of carbonaceous aerosols.
30          Ashbaugh and Lawson (1991) analyzed data collected in 1985, 1987 and 1989  random
31     roadside surveys and showed that, for the low-idle test,  20% of the vehicles were responsible

       April 1995                               5-30       DRAFT-DO NOT QUOTE  OR CITE

-------
 1     for 80% of the CO emissions and 20% (not necessarily the same vehicles) were responsible
 2     for 80% of the VOC hydrocarbon (HC) emissions.
 3          While several tests have examined on-road emissions of both nitrogen oxide and VOC
 4     precursors, only Hansen and Rosen (1990) report  individual on-road vehicle measurements of
 5     particulate emissions.  They measured the ratio of light-absorbing carbon to CO2 in the
 6     exhausts of 60 gasoline-fueled vehicles in Berkeley,  CA.  The ratio of carbon to excess CO2
 7     above background provides an estimate of emissions per unit of fuel combusted.  By making
 8     reasonable assumptions about speed and gasoline mileage, these ratios can be translated into
 9     grams/vehicle mile traveled.  Their experiment found a factor of 250 between the highest and
10     lowest ratio of light-absorbing carbon to CO2 for the 60 vehicles tested.  This is  at odds with
11     use of a single value for primary particle emissions, regardless of vehicle type  or operating
12     condition, typically employed in current emissions models.
13          These studies show that while vehicle emissions models may  function well under
14     idealized conditions,  they underestimate the effects of high emitting vehicles that may be
15     major sources of VOCs.  Though data are lacking to verify the primary particle emissions in
16     these models, it  is very unlikely that the values currently in use account for the variability in
17     emissions from different types of vehicles.
18
19
20     5.6     SIZES DISTRIBUTIONS OF PRIMARY PARTICLE EMISSIONS
21          Recent measurements of the size distributions of primary particles confirm U.S.
22     Environmental Protection Agency (1982a) conclusions that most fugitive dust emissions are
23     in particles larger than 2.5 pm and  that the majority of emissions from combustion sources
24     are in sizes smaller than 2.5 /mi. Figures 5-6 and 5-7 from Houck et al. (1989, 1990)  were
25     derived from a major characterization of different source emissions in California conducted
26     during 1986.  Hot exhaust samples  were diluted to ambient temperatures prior  to sampling
27     onto filter media through impactor inlets with 50% cut-points  of 1, 2.5, 10, and  —30 /mi.
28     These figures show that combustion products are nearly always less than 2.5 /im in size.
29          Figure 5-8 shows examples of size distributions in dust from paved and unpaved roads,
30     agricultural soil, sand and gravel, and alkaline lake bed sediments  which were  measured in a
31     laboratory resuspension chamber as part of the California study (Chow et al.,  1994).  This

       April 1995                               5-31       DRAFT-DO NOT QUOTE OR CITE

-------
         100
'

80


60
Q_
1
0)
a.
40

20
0
~


-





—







|H
in
1
f
fK^:--:':~':.
y//M'/3aftf-'faf>'#.




52.3% (<10^)




10.7% (<2.5n)


^^^g





,^\>.
x V"V,
\^4
r -U' " •
•;';.i"* V
1|
92.8% (<10n)
82.7% (<1 n)
81.6%(<2.5n)











••••••

-^"^x?"
,;|

\>^:

;^%-
IS
1
||
IsJ
1
95.8% (<10n)
93.1%(<2.5n)
92.4% (<1 n)


Code
^I^\ >1Q\i
2.5u - 10u
•^^ ™ • ** "
^ 1^-2.5(1
E3 
-------
         100
          80
          60
      Q_
      P
      I
      o>
      Q_
          40
          20
                        96.2%

                        92.3% (<2.5\i
                        91.8%
                   Diesel
                   Truck
                   Emissions
                                         99.2%
                                         97.4%
                                         87.4%
Crude Oil
Combustion,
Mean of 2
Composites
                                                              Code
                                                           34.9% (<1 Oji)
                                                           5.8%
                                                           4.6%
     Fresno
     Construction
Figure 5-7. Size distribution of California particle emissions, 1986.
April 1995
  5-33
DRAFT-DO NOT QUOTE OR CITE

-------
            100
                    Paved       Unpaved     Agriculture    Soil/Gravel
                   Road Dust    Road Dust        Soil
                          EI3<1.0nm [ZI|<2.5nm  ••< 10 urn
                                                                    Alkaline
                                                                    Lake Bed
       Figure 5-8.  Particle size distribution in laboratory resuspension chamber.
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
figure shows substantial variation in primary particle size among some of these sources.  The
PMj o abundance (6.9%) in the alkaline lake bed dust is twice its abundance in paved and
unpaved road dust.  Approximately 10% of TSP is in the PM2 5  fraction and approximately
50% of TSP is in the PM10 fraction.  The sand/gravel dust sample shows that 65% of the
mass consists of particles larger than the PM10 fraction. The PM2 5 fraction of TSP in
alkaline lake beds and sand/gravel is approximately 30% to 40% higher than the other soil
types.  Particle emissions in the respirable fraction can be expected to vary substantially
among these different fugitive dust sources.
       April 1995
                                        5-34
DRAFT-DO NOT QUOTE OR CITE

-------
 1     5.7 CHEMICAL COMPOSITIONS OF PRIMARY PARTICLE
 2          EMISSIONS
 3          The widespread use of receptor models since U.S. EPA (1982a) has resulted in many
 4     published chemical compositions for different paniculate sources.  Table 5-2 shows the
 5     relative amounts of different chemical species in the profiles from different source types.
 6     These are deduced from many different profiles compiled by Radian (1989) and by Chow
 7     et al. (1994) for EPA's SPECIATE source composition library.
 8          Figures 5-9 through 5-12 show examples of the chemical abundances in specific PM2 5
 9     profiles for paved road dust, motor vehicle exhaust, residential  wood combustion, and a coal-
10     fired power plants.  These were measured in Denver, CO during 1987 (Watson and Chow,
11     1994).  Substantial differences in chemical composition exist for these emitters,  while these
12     differ from site to site.
13          The road dust profile in Figure 5-9 contains large abundances  of aluminum, silicon,
14     potassium, calcium, and iron.  Though total potassium is abundant in road dust,
15     water-soluble potassium constitutes less than one-tenth of the total.  Strontium and lead are
16     also present at detectable levels.  Paved road dust is much like  an ambient PM10 sample,
17     with a complex combination of particulate matter from a wide variety of sources, especially
18     other geological source types.  This complexity is evident in the comparison of a paved road
19     dust profile reported by Chow et al. (1991) for Phoenix, AZ, with profiles from other
20     geological sources in the area. Chow et al. (1991) noted that the abundance of organic
21     carbon in the profile was 11+9%,  larger and more variable than its abundance in profiles
22     from agricultural land,  construction sites, and vacant lots.  The presence of tire  wear,
23     detritus,  and engine oils can account for this  enrichment.  This organic carbon  content
24     places an effective upper limit on the contribution from tire wear and other carbon sources to
25     suspendable paved road dust.  Approximately 25 %  of tire material consists of styrene-
26     butadiene rubber (SBR) (Pierson and Brachaczek, 1974). Ondov (1974) measured elemental
27     components of tire material and found minor concentrations for most species, with S (-2%),
28     Cl (~ 1.5%), and Zn (~ 1 %) being the most abundant components.
29          The abundances of Pb and Br in Phoenix paved road dust  were more than  double the
30     concentrations in the other geological profiles, indicating the presence of tailpipe exhaust
       April 1995                              5.35       DRAFT-DO NOT QUOTE OR CITE

-------
1
1— 1
VO



i
0-v
O
6
O
O
H
O
Cl
O
H
W
O
50
n
TABLE 5-2. TYPICAL CHEMICAL ABUNDANCES IN SOURCE EMISSIONS.
Chemical Abundances

Source Type
Geological Material
Motor Vehicle
Vegetative Burning
Residual Oil Combustion
Incinerator
Coal Fired Power Plant
Marine


Dominant
Particle Size
Coarse
Fine
Fine
Fine
Fine
Fine
Coarse



< 0.1%
Cr, Zn, Rb, Sr, Zr
Cr, Ni, Y+
Ca, Mn, Fe, Zn,
Br, Rb, Pb
K+, OC, Cl, Ti,
Cr, Co, Ga, Se
V, Mn, Cu, Ag, Sn
Cl, Cr, Mn, Ga,
As, Se, Br, Rb, Zr
Ti, V, Ni, Sr, Zr,
Pd, Ag, Sn, Sb, Pb



0.1 to 1%
Cr, N03-, S04=,
NH4+, P, S, Cl, Ti,
Mn, Ba, La
NH4+, Si, Cl, Al,
Si, P, Ca, Mn, Fe,
Zn, Br, Pb
NOj, S04=, NH4+,
Na+
Na+, NH4+, Zn,
Fe, Si
K+, Al, Ti, Zn,
Hg
NH4+, P, K, Ti, V,
Ni, Zn, Sr, Ba, Pb
K, Ca, Fe, Cu, Zn,
Ba, La, Al, Si



1 to 10% > 10%
OC, EC, Al, K, Si
Ca, K, Fe
S, Cl, NO§, OC, EC
S04=, NH4+
K+, K, Cl, Cl- OC, EC
Ni, OC, EC, V S, SO4=
NO§, Na+, EC, SO4=, NH4+,
Si, S, Ca, Fe, OC, Cl
Br, La, Pb
SO4=, OC, EC, Si
Al, S, Ca, Fe
NOj, S04=, Na + , Na, Cl",
OC, EC Cl



-------
                                                        <3> v <3*

                                 Chemical Compound

Figure 5-9.  Chemical abundances for PM2 5 profiles of road dust.
                 ^w-  ^    
-------
            ^^^^^^^/^^^f
                         VN
                                Chemical Compound
Figure 5-11. Chemical abundances for PM2 5 profiles of wood burning.
          10
                            Coal-fired Power Plant
                                Chemical Compound
Figure 5-12. Chemical abundances for PM2 5 profiles of coal-fired power plant.
April 1995
5-38
                                            DRAFT-DO NOT QUOTE OR CITE

-------
  1     from vehicles burning leaded fuels.  These will not be good indicators of exhaust
  2     contributions today because tetraethyl lead is no longer used as a fuel additive.  Enrichments
  3     in species from clutch and brake wear were absent in the Phoenix paved road dust profiles.
  4     These are often composed of asbestos and/or semi-metal carbon composites. Ondov (1974)
  5     measured abundances of ~ 14% Mg, -2% Ca, ~4% Fe, and -1% Ba in asbestos brake
  6     shoes, while Anderson et al. (1973)  reported Si abundances of ~ 10%.  Cooper et al.  (1987)
  7     examined the elemental composition of semi-metal brake shoes and found abundances  of
  8     -45% Fe, -2% Cu, -0.5%  Sn,  -3% Ba, and -0.5% Mo.  None of these  species was
  9     found in the Phoenix paved road dust profiles at levels significantly in excess of their
 10     abundances in other geological  sub-types.
 11          The motor vehicle exhaust profile contains high concentrations of organic  and elemental
 12     carbon; but their ratios are much different from those found in wood combustion with the
 13     abundance of elemental carbon  being nearly equal to the organic carbon abundance. Bromine
 14     and lead are also much larger components in vehicle exhaust than in other source profiles.
 15     However,  Br and Pb have  been phased out of most U.S. gasolines, and this example does not
 16     represent current motor vehicle emissions profiles  found in the United States today.
 17          Pierson and Brachaczek (1976; 1983) pioneered the chemical characterization of
 18     paniculate emissions from  motor vehicles.  These  results, acquired from 1970 through 1981,
 19     were used extensively in the early days of CMB modeling of total  suspended particulate
 20     matter (e.g., Watson, 1979; Kowalczyk et al.,  1978).  Little additional work on the chemical
 21      characterization of particulate motor  vehicle emissions was conducted until promulgation of
 22      the PM10 NAAQS in 1987. It was soon recognized that fleet composition,  emissions
 23      controls, fuels, engine designs,  and vehicle maintenance had changed considerably  since the
 24      tests of Pierson and Brachaczek (1976; 1983), and that new tests, and types of testing, were
 25      needed to obtain chemical source profiles.
 26          Watson et al. (1988b) obtained  six roadside samples under a freeway overpass, in a city
 27      bus yard,  and near busy intersections in Reno,  NV, in 1986 and 1987 as part of the State of
28      Nevada Air Pollution Study (SNAPS) (Chow et al., 1988).  Cooper et al. (1987) and NBA
29      (1990a; 1990b; 1990c), in studies conducted for SCAQMD (Gray et  al., 1988; Zeldin  et al.,
30      1990), measured exhaust from eleven unleaded-gasoline-fueled vehicles, 3 leaded-gasoline-
31      fueled vehicles, and 2 heavy-duty diesel-fueled  trucks  operating on laboratory dynamometers

        April 1995                               5.39      DRAFT-DO NOT QUOTE OR CITE

-------
 1     in 1986.  The FTP and a steady-state test at 35 mph were followed for the gasoline-fueled
 2     vehicle tests.  Diesel-fueled trucks were tested under modified FTP and steady-state
 3     conditions.  Cooper et al.  (1987) also took three roadside samples in a tunnel under the Los
 4     Angeles Airport.
 5          Watson et al. (1990b; 1990c) report the results from dynamometer tests of eight
 6     unleaded-gasoline-fueled vehicles, three leaded-gasoline-fueled vehicles, and three light- to
 7     medium-duty diesel-fueled vehicles conducted in 1988 during a Denver visibility  study.  The
 8     FTP was applied to vehicles under conditions (temperatures <  40 °F) similar to  those found
 9     in wintertime Denver. In 1987, Houck et al. (1989) took three samples of heavy-duty diesel-
10     fueled truck exhaust at a roof monitor over the Wheeler weigh  station near Bakersfield, CA.
11     These measurements were used with SCAQMD source profiles for PM10 source
12     apportionment  in California's San Joaquin Valley (Chow et al., 1992; 1993d).  The TOR
13     method described by Chow et al. (1993b) was applied in all of these tests except that of
14     Cooper et al. (1987). The Cooper et al. (1987) carbon analysis monitored light transmission
15     instead of reflectance to implement the pyrolysis correction (Gary, 1990).  Watson et al.
16     (1994) report profiles derived in Phoenix, AZ during 1989.
17          There are significant similarities and differences among the chemical compositions of
18     these different  motor vehicle profiles measured in different areas, at different times,  and by
19     different methods.  The Denver diesel-fueled profiles have a much higher abundance of
20     elemental carbon (74+21 %) than the SCAQMD (52+5%), Wheeler Station (43 +8%), or the
21     PHDIES (33±8%) profiles.  The organic carbon in these profiles is 23±8% for  Denver,
22     36±3% for SCAQMD, 49 + 13% for Wheeler Station,  and 40±7% for PHDIES.
23          For leaded-gasoline-fueled vehicles, the OC and EC abundances were  found to be
24     67±23% and 16±7% in the Denver tests,  52±4% and 1.3 + 1%  in the SCAQMD steady-
25     state tests, and 31+20% and 15±2% in the SCAQMD FTP tests, respectively.   For
26     unleaded-gasoline-fueled vehicles, the OC and EC abundances were found to be 76+29%
27     and 18±11% for Denver tests, 93 ±52% and 5 ±7% for SCAQMD steady-state tests, and
28     49 + 10% and 39+8% for SCAQMD FTP tests, respectively.  These compare to  OC and EC
29     abundances in  PHAUTO of 30 + 12% and 14±8%, respectively.
30          In roadside and tunnel tests which included mixtures of diesel-, leaded-, and unleaded-
31     fueled vehicles, Watson et al. (1988a;1990c) found abundances of 50 + 24% OC and

       April 1995                               5-40       DRAFT-DO NOT QUOTE OR CITE

-------
 1      28 ±  19% EC, and Cooper et al. (1987) found abundances of 38 ± 6% OC and 38 + 5%
 2      EC.  These are similar to the 39 + 19% OC and 36 + 11% EC abundances found in this study.
 3           All of these profiles represent emissions from a small number of vehicles relative to the
 4      entire vehicle population. In each of these studies, as in this one, there were individual
 5      samples which deviated in their compositions from the majority of profiles.  Further study is
 6      needed to determine why these deviations exist and how they are related to the mixture of
 7      vehicles tested. For wood burning in fireplaces and stoves,  organic carbon is by far the most
 8      abundant constituent, followed by elemental carbon.  Chlorine and potassium also approach
 9      an abundance of 1% in these emissions, as do sulfate and nitrate.  The water soluble
10      potassium equals total potassium in wood burning  emissions.
11           The coal-fired power plant profile in  Figure 5-12 has several chemical abundances that
12      are similar to those of the road dust profile.  The sulfate level is much higher in these
13      emissions than in the other profiles, and the organic and elemental carbon fractions  are much
14      lower. Selenium is clearly detectable in the power plant profile, whereas it is below
15      quantifiable limits in the other profiles.
16           The chemicals identified in Figures 5-9 through 5-12 represent the major components
17      that contribute to atmospheric light extinction (i.e., crustal, sulfate, nitrate, ammonium,
18      organic carbon, elemental carbon) as  well as chemical patterns that help to distinguish one
19      source of primary emissions from others (e.g., selenium, potassium, aluminum,  silicon,
20      organic carbon, elemental carbon). In addition to  these commonly measured components, it
21      is possible that isotopic ratios in source emissions may vary  in an informative way with the
22      nature of the combustion process and with  the geologic age and character of the source input
23      material. Carbon-14, for example, has been widely used to  separate contemporary carbon
24      due to vegetative burning from  carbon emitted by fossil fuel combustion (Currie et al.,
25      1984). Certain isotopic ratios might also distinguish coal-fired  power plant emissions from
26      other sources.  Fuels from different mines  may have isotopic differences that allow otherwise
27      similar source  emissions from different power plants to be distinguished.
28
29
30
        April 1995                                5_41       DRAFT-DO NOT QUOTE OR CITE

-------
 1     5.8  EMISSIONS MODELS AND EMISSIONS INVENTORIES
 2           As noted above, major discrepancies exist between relative amounts of emissions and
 3     contributions to suspended particles found in many areas. Major re-design is  needed to
 4     create more accurate emissions models to improve quantification of source-receptor
 5     relationships.
 6           Emissions models are intended to estimate the emissions rates as a function of space
 7     and time of selected pollutants from point, area,  and mobile sources.  In contrast to an
 8     emissions inventory, which is a static catalogue of emissions estimates for a given
 9     geographical area and averaging time, an emissions model is capable of accessing activity
10     data bases from a multitude of information- gathering agencies and determining actual
11     emissions for relatively small regions and averaging times.
12           Emissions models have, at their base, several activity surrogates  that are considered to
13     have  some relationship to emissions. Activities relevant to the major source contributions
14     identified in Table 5-1 include: (1) number of acres burned, locations, and durations of
15     wildfires and prescribed burns; (2) cords of wood and tons of coal sold for  residential
16     heating; (3) vehicle miles traveled for mobile source emissions; (4) heads of livestock,
17     pounds  of fertilizer, and number sewage treated for ammonia;  and (5) amount of product
18     produced for industrial sources (e.g. power generated for power plant emissions; tons of
19     coke  produced for a coke oven).  Wherever possible, different activity data for the same
20     sources and emitted species are used to  evaluate uncertainties.  For example, Watson et al..
21     (1990) showed for Denver, CO, that proportional differences between gasoline and diesel
22     powered mobile source activities might be separately estimated from vehicle counts at
23     selected roadways,  areawide fuel sales,  and vehicle registrations.  Differences between
24     separate activity estimates can be used to quantify the uncertainty  of emissions rates.
25           Upon  these activity data are imposed emissions/activity relationships (commonly termed
26     "emissions factors") that may require meteorological inputs such as temperature,  relative
27     humidity, or wind speed. These relationships are easily replaceable and include the ability to
28     propagate the precisions specified for the input data. The emissions/activity relationships are
29     specific for different source sub-types.   As an example, certain roads may have  greater or
30     lesser proportions of diesel truck traffic, older and newer vehicles, and vehicles which were
31     recently started and ones which are fully warmed up.  In the idealized emissions model, each

       April 1995                                5-42      DRAFT-DO NOT QUOTE OR CITE

-------
 1      road segment is assigned a classification for a specific distribution of vehicles, and a separate
 2      emissions/activity relationship is determined for each classification.
 3           Emissions models should have several attributes, i.e., they should:  (1) be based on
 4      documented scientific and engineering principles;  (2) be composed of special purpose
 5      modules which can be updated with new information and new science when available;
 6      (3) have activity levels and emissions/activity relationships specific to a non-attainment area;
 7      (4) contain error propagation algorithms to provide precision estimates on outputs; (5) use
 8      independent activity data bases and emissions/activity relationships of equivalent quality to
 9      estimate accuracy; (6) adjust the emissions/activity relationship in response to environmental
10      variables,  especially meteorology; (7) allow the addition, subtraction, or modification of
11      emissions  for special events; (8) retain traceability of all information to allow quality
12      auditing; (9) provide output displays, statistics, and data bases which can be used  for
13      modeling,  data analysis, control  strategy development, and quality assurance;  and
14      (10) calculate the effects  of changes, such as population growth, implementation of PM10,
15      control measures, and changes in land use.
16           Unfortunately, there are no emissions models used or even currently available that
17      attain all these attributes.  Attempts have been made to improve the  current state of the art
18      with creation of the Flexible Response Emissions  Data System (FREDS) (Lebowitz et al.,
19      1987) for the National Acid Precipitation Assessment Program (NAPAP) and  the Emissions
20      Preprocessor System (EPS) for the Urban  Airshed Model (SAI, 1990).  The closest approach
21      is the Geographical Emissions Modeling and Assessment Program (GEMAP; Dickson and
22      Oliver, 1993) developed for the San Joaquin Valley Air  Quality Study and Atmospheric
23      Utility Signatures, Predictions and Experiments  (SJVAQS/AUSPEX) in California (Solomon,
24      1994).
25           Assumptions inherent in using a static emissions inventory to represent short-term
26      events are: (1) emissions rates are constant,  typically averaged over a year and sometimes
27      over a season; (2) emissions  factors relating activities to emissions apply to all emitters at all
28      locations at all times; (3) the values of activity variables are highly correlated with actual
29      emissions; and (4) all major emitters have  been  identified and included.
30           It is not difficult to  find major deviations from these assumptions, even in a  small area
31      over a short time period.   As might be expected averaging over large spatial scales and time

        April  1995                               5.43       DRAFT-DO NOT QUOTE OR CITE

-------
 1     periods evens out much of the variability inherent in applying these assumptions to individual
 2     episodes and locations.  However, short-term episodes of a few hours or a few days duration
 3     and source influences over spatial scales of less than 100 km are exactly what is needed to
 4     evaluate source contributions to excessive 24-h values of paniculate concentrations in non-
 5     attainment areas.
 6          Emissions inventories are often used to develop emissions reductions strategies by the
 7     process of linear rollback Earth (1970). This approach assumes that pollutant concentrations
 8     and the effects of those concentrations (e.g., health effects) are directly proportional to
 9     emissions within a selected geographical region.   For this approach, an area is defined that
10     receives emissions from the  sources to be controlled and ambient concentrations of the
11     pollutants of interested are measured within and  outside this area.  If there are contributions
12     from the suspected sources,  the within-area concentrations will be significantly higher than
13     the outside-area concentrations, and the outside-area concentrations  are subtracted to obtain
14     the incremental amount contributed by in-area sources.  Emissions of the pollutant or the
15     precursor within the study area are reduced by an amount proportional to the desired
16     reduction in ambient concentrations.
17          The linear rollback method is simple and its application in urban areas has resulted in
18     reduced ambient concentrations for primary particles when emissions estimates are accurate,
19     spatial scales are  large, and  averaging times are  long.  It is currently being tested for U.S.
20     utility  sulfur dioxide emissions that are to be reduced to 8.95 million tons per year (by more
21     than 50% from current emissions) by the year 2002.  The rollback method,  however,  has not
22     been proven to be accurate for short duration events over small spatial scales for secondary
23     aerosol when emissions estimates are uncertain.
24
25
26     5.9  SUMMARY AND CONCLUSIONS
27          The ambient atmosphere contains both primary and secondary particles; the  former are
28     emitted directly by sources, and the latter are formed from gases  (SO2, NOX, HN4, VOCs)
29     that are directly emitted by  sources.  Fugitive dust  is a primary pollutant, and also has a role
30     in secondary particle formation.  Major sources  of particle emissions are classified as major
        April 1995                               5-44       DRAFT-DO NOT QUOTE OR CITE

-------
  1     point sources, mobile sources, and area sources; these are anthropogenic.  Natural sources
  2     also contribute to ambient concentrations.
  3           The 1982 Criteria Document emphasized emissions from industrial sources, especially
  4     primary particles.  SO2 was the only precursor of secondary particles considered.  Since
  5     1982, many of these sources have been controlled, yet particle standards are exceeded in
  6     many areas.
  7           Source and receptor models are used to quantify major contributions to excess PM10
  8     concentrations.  Source models use emissions inventories and meteorological data to product
  9     particle dispersion and PM10 concentrations measured at receptors.  Receptor models use
 10     chemical composition of emissions and receptor concentrations to estimate the contribution of
 11     sources.  The  latter are more appropriate to identify sources in non-contaminant areas.
 12           Fugitive  dust is a major contribution to PM10 at nearly all sampling site, although the
 13     average fugitive dust source contribution is highly variable among sampling sites within  the
 14     same areas, and is highly variable between seasons.
 15          Primary  motor vehicle exhaust makes up as  much as 40% of average PM10 at many
 16     sampling sites. Vegetative burning outdoor and residential wood burning are significant
 17     sources in residential areas.  Fugitive dust from paved and unpaved roads, agricultural
 18     operations, construction, and soil  erosion constitute -90% of nationwide primary emissions.
 19     All of the emissions have remained relatively constant over the 8-year period except for  those
 20     from soil erosion.
 21           The majority of wind erosion occurs in the dustbowl region; estimates are influenced by
 22     annual precipitation and wind-speed distribution.  The major non-fugitive dust emitters are
 23      other industrial processes and exhaust from highway vehicles.  Fuel combustion from
 24      utilities, industrial, and other sources together contribute between  1 to 2%  to total primary
 25      particle emissions.  Industrial fuel combustion emissions were reduced by one-third and other
 26      fuel combustion emissions were reduced by one half between 1983 and 1992.  On high-way
 27      vehicle emissions increased by 50%, primarily due to large increases in the number of
 28      vehicle miles traveled.  Electric utilities account for the largest fraction of sulfur dioxide,
29      nearly 70% of total emissions. These emissions have not changed substantially over the  10
30      years reported.  Annual averages do not reflect the seasonality of certain emissions,
31      residential wood burning in fireplaces and stoves,  for example.  Cold weather also affects

        April 1995                                5.45       DRAFT-DO NOT  QUOTE OR CITE

-------
 1     motor vehicle exhaust emissions, both in terms of chemical composition and emission rates.
 2     Planting, harvesting, and fertilizing and harvesting are also seasonal.  Fugitive dust consists
 3     of geological material that is suspended into the atmosphere by natural wind and by
 4     anthropogenic activities from sources such as paved and unpaved roads,  construction and
 5     demolition of buildings and  roads,  storage piles, wind erosion, and agricultural tilling.  There
 6     are obvious discrepancies between the proportion of fugitive dust in primary emissions and
 7     geological  contributions to PM10 calculated by recptor models, due to contributions from
 8     secondary  aerosols, which are not  included in the primary PM10 emission estimates.  Even
 9     when secondary aerosol is subtracted, however, other sources such as vegetative burning and
10     wood combustion make larger relative contributions to ambient concentrations than is
11     indicated by the emissions inventories.   Fugitive dust estimates are especially affected by the
12     general limitations of emissions inventories.  Annual and spatial averages do not reflect the
13     seasonality of certain emissions. Planting and harvesting are seasonal, as are high winds
14     which might cause erosion.  Paved road dust emissions might be much higher after rain
15     storms when dirt is tracked  from unpaved areas onto paved roads.  The  spatial disaggregation
16     of fugitive dust emissions is poorer than that for all other source categories.  Whereas most
17     mobile sources are confined to established roadways, and most area sources correspond to
18     population density,  suspendable dust is everywhere.  Modern data bases, computer systems,
19     and information management software could be applied to improve existing inventories
20     without major additional costs after the  initial investment in establishing  an inventory
21     methodology.  Mobile sources are major emitters of primary particles, oxides of nitrogen,
22     and volatile organic compounds.  They  are also minor emitters of sulfur dioxide and
23     ammonia.  On-road motor vehicles using gasoline- and diesel-fueled engines are by far the
24     largest component of mobile source emissions, and the emissions estimation methods are
25     most highly developed for these vehicles.  Studies show that while vehicle emissions models
26     may function well under idealized  conditions, they underestimate the effects of high emitting
27     vehicles that may be major sources of VOCs.  Motor vehicle exhaust contains high
28     concentrations of organic and elemental carbon, but their ratios are much different from
29     those found in wood combustion with the abundance of elemental carbon being nearly equal
30     to the organic carbon abundance.  There are major discrepancies between the relative
31     amounts of emissions and contributions to suspended particles found in many areas.  Some

       April 1995                               5-46       DRAFT-DO NOT  QUOTE OR CITE

-------
1     major re-design is needed to create more accurate emissions models that can improve the
2     quantification of source-receptor relationships.  Emissions models are intended to estimate
3     the emissions rates as a function of space and time of selected pollutants from point, area,
4     and mobile sources.  In contrast to an emissions inventory, which is a static catalogue of
5     emissions estimates for a given geographical area and averaging time, an emissions model is
6     capable of accessing activity data bases from a multitude of information- gathering agencies
7     and determining actual emissions for relatively small regions and averaging times.
8
9
      April 1995                               5.47       DRAFT-DO NOT QUOTE OR CITE

-------
 1      REFERENCES
 2
 3     American Society for Testing and Materials. (1990a) Standard test method for amount of material in soils finer
 4            than the no. 200. In: Annual book of ASTM standards. Philadelphia, PA: American Society for Testing
 5            and Materials; pp. 193-194; document no. D1140.
 6
 7     American Society for Testing and Materials. (1990b) Standard test method for particle-size analysis of soils. In:
 8            Annual book of ASTM standards.  Philadelphia, PA: American Society for Testing Materials; pp. 94-100;
 9            document no. D422.
10
11     Anderson, A. E.; Gealer, R. L.; McCune, R. C.; Sprys, J. W. (1973) Asbestos emissions from brake
12            dynamometer tests. Warrendale, MI: Society of Automotive Engineers; document SAE 7305-49.
13
14     Ashbaugh, L.; Lawson, D. R. (1991) A comparison of emissions from mobile sources using random roadside
15            surveys conducted in 1985, 1987,  and 1989. Presented at:  the 84th annual meeting of the Air & Waste
16            Management Association; June; Vancouver, BC, Canada. Pittsburgh, PA: Air & Waste Management
17            Association.
18
19     Ashbaugh, L. L.; Lawson,  D. R.; Bishop, G. A.; Guenther, P. L.; Stedman, D. H.;  Stephens, R.  D.; Groblicki,
20            P. J.; Parikh, J. S.; Johnson, B. J.; Huang, S. C. (1992) On-road remote sensing of carbon monoxide
21            and hydrocarbon emissions during several vehicle operating conditions. In: Chow,  J. C.; Ono,  D.  M.,
22            eds. PM10 standards and nontraditional paniculate source controls, v. II. Pittsburgh, PA: Air & Waste
23            Management Association; pp. 885-898.
24
25     Axetell, K.; Zell, J.  (1977) Control of reentrained dust from paved streets. Kansas City, MO: U.S.
26            Environmental  Protection Agency, Region VII; EPA report no. EPA-907/9-77-007. Available from:
27            NTIS, Springfield,  VA; PB-280 325.
28
29     Bagnold, R. A. (1937) The transport of sand by wind. Geogr. J. 89: 409-438.
30
31     Bagnold, R. A. (1941) The physics of blown sand and desert dunes. London, United Kingdom: Methuen & Co.
32
33     Barnard, W. R.; Stewart, M. A. (1992) Modification of the 1985 NAPAP wind erosion methodology to assess
34            annual PM10 emissions for EPA's  emissions trends report. In: Chow, J.  C.; Ono, D. M., eds.  PM10
35            standards and nontraditional paniculate source controls. Pittsburgh, PA:  Air and Waste Management
36            Association; pp. 121-130.
37
38     Barnard, W. R.; Carlson,  P. M.; Stewart, M. A.  (1992) Fugitive PM10  emissions estimates developed for EPA's
39            emission trends report. In: Chow,  J. C.;  Ono, D.  M., eds. PM10 standards and  nontraditional paniculate
40            source controls. Pittsburgh, PA: Air & Waste Management Association; pp. 110-120.
41
42     Earth, D. S. (1970) Federal motor vehicle emission goals for CO, HC and NOX based on desired air quality
43            levels. J. Air Pollut. Control Assoc. 20: 519-523.
44
45     Belly, P. Y. (1964) Sand movement by wind. U.S. Army Coastal  Engineering Research Center; pp. 1-38;
46            technical memo 1.
47
48     Bisal, F.; Hsieh, J. (1966)  Influence of moisture on erodibility of soil by wind.  J. Soil Sci. 102: 143-146.
49
50     California Air Resources Board. (1993) EMFAC7F workshop jointly sponsored  by the Technical Support
51            Division and the Mobile Source Division; June; Sacramento, CA.  Sacramento, CA: California Air
52            Resources Board.
53
         April 1995                                    5-48        DRAFT-DO NOT QUOTE OR CITE

-------
  1     Gary, R.  (1990) Sunset Laboratory thermal-optical analysis for organic/elemental carbon analysis—analysis
  2            description. Forest Grove, OR: Sunset Laboratory.
  3
  4     Chepil, W.  S. (1952) Improved rotary sieve for measuring state and stability of dry soil structure. Soil Sci. Soc.
  5            Am. Proc. 16: 1113-1117.
  6
  7     Chepil, W.  S. (1956) Influence of moisture on erodibility of soil by wind. Soil Sci. Soc. Am. Proc. 20:  288-292.
  8
  9     Chepil, W.  S.; Woodruff, N. P. (1963) The physics of wind erosion and its control. Adv. Agron. 15: 211-299.
 10
 11     Chow, J.  C.; Watson, J.  G.  (1994) Contemporary source profiles for geological material and motor vehicle
 12            emissions [final report]. Research Triangle Park, NC: U.S. Environmental Protection Agency, Office of
 13            Air Quality Planning and  Standards.
 14
 15     Chow, J.  C.; Watson, J.  G.; Frazier, C. A.; Egami, R. T.; Goodrich, A.; Ralph,  C. (1988) Spatial and
 16            temporal source contributions to PM10 and PM2 5 in Reno, NV. In: Mathai, C. V.; Stonefield, D. F.,
 17            eds. PMi0: implementation of standards. Pittsburgh, PA: Air Pollution Control Association; pp. 438-457.
 18
 19     Chow, J.  C.; Watson, J.  G.; Richards, L. W.; Haase,  D. L.; McDade, C.; Dietrich, D. L.; Moon, D.; Sloane,
 20            C. (1991) The 1989-90 Phoenix PM10 study, v. II: source apportionment [final report]. Phoenix,
 21            AZ: Arizona Department  of Environmental Quality; report no. 8931.6F1.
 22
 23     Chow, J.  C.; Watson, J.  G.; Lowenthal, D. H.; Frazier, C.; Hinsvark, B.;  Prichett, L.; Neuroth, G. (1992a)
 24            Wintertime PM10 and PM2 5 mass and chemical compositions in Tucson, Arizona. In: Chow, J. C.; Ono,
 25            D. M., eds. Transactions:  PM10 standards and  nontraditional paniculate source controls. Pittsburgh,
 26            PA: Air and Waste Management; pp. 231-243.
 27
 28     Chow, J.  C.; Watson, J.  G.; Lowenthal, D. H.; Solomon, P. A.; Magliano, K. L.; Ziman,  S. D.; Richards, L.
 29            W. (1992b) PM10 source apportionment in California's San Joaquin Valley. Atmos. Environ. Part A
 30            26: 3335-3354.
 31
 32     Chow, J.  C.; Liu, C. S.;  Cassmassi, J.; Watson, J. G.; Lu, Z.; Pritchett, L. C. (1992c) A neighborhood-scale
 33            study of PM10 source contributions in Rubidoux, California. Atmos. Environ. Part A 26: 693-706.
 34
 35     Chow, J.  C.; Watson, J.  G.; Ono, D. M.; Mathai, C.  V.  (1993) PM10 standards and nontraditional paniculate
 36            source controls: a summary of the A&WMA/EPA international specialty conference. Air Waste
 37            43: 74-84.
 38
 39     Chow, J.  C.; Watson, J.  G.; Houck, J. E.;  Pritchett, L. C.; Rogers, C. P.; Frazier, C. A.;  Egami, R. T.; Ball,
 40            B. M. (1994) A laboratory resuspension chamber to measure fugitive dust size distributions and chemical
 41            compositions. Atmos. Environ. 28: 3463-3481.
 42
 43     Cooper, J. A.; Redline, D. C.; Sherman, J. R.; Valdovinos, L.  M.; Pollard, W. L.; Scavone, L. C.;
 44            Badgett-West, C.  (1987) PM10 source composition library for the South Coast Air Basin: volume I,
45            source profile development documentation final report. El Monte, CA:  South Coast Air Quality
46            Management District; July 15.
47
48     Cooper, J. A.; Sherman,  J. R,; Miller, E.; Redline,  D.; Valdonovinos, L.; Pollard, W. (1988) CMB source
49            apportionment of  PM10 of an oil-fired power plant in Chula Vista, California. In: Mathai, C. V.;
50            Stonefield, D. H., eds. PM10: implementation of standards. Pittsburgh, PA: Air & Waste Management
51             Association; pp. 495-507.
52
53      Federal Register. (1991) Designations and classifications for initial PMIO nonattainment areas: 40 CFR Part 81.
54            F. R. 56: 11101.


         April 1995                                    5.49        DRAFT-DO NOT QUOTE OR CITE

-------
  1      Federal Register. (1993) Reclassification of moderate PM10 nonattainment areas to serious areas: 40 CFR Part
  2             81. F. R. 58: 3334.
  3
  4      Federal Register. (1994) Designations of areas for air quality planning purposes: 40 CFR Part 81. F. R.
  5             58: 67334.
  6
  7      Fryrear, D. (1992) Measured wind erosion of agricultural lands. In: Chow, J. C.; Ono,  D. M., eds. PM10
  8             standards and nontraditional paniculate source controls. Pittsburgh, PA: Air & Waste Management
  9             Association; pp. 433-439.
10
11      Gertler, A. W.; Fujita, E. M.; Pirson, W. R.; Wittorff, D. N. (1995) Apportionment of NMHC tailpipe vs. non-
12             tailpipe emissions in the Fort McHenry and Tuscarora tunnels. Atmos. Environ.: accepted.
13
14      Gillette, D. (1980) Major contributions of natural primary continental aerosols: source mechanisms. In: Kniep,
15             T. J.;  Lioy, P.  J., eds. Aerosols: anthropogenic and natural, sources  and transport; January 1979; New
16             York,  NY. Ann N. Y.  Acad.  Sci.  338: 348-358.
17
18      Gillette, D. A.; Hanson, K. J.  (1989) Spatial and temporal variability of dust production caused by wind erosion
19             in the  United States. J. Geophys. Res.  94: 2197-2206.
20
21      Gillette, D. A.; Stockton, P.  H. (1989) The effect of nonerodible particles on wind erosion of erodibile surfaces.
22             J. Geophys. Res. [Atmos.] 94: 12885-12893.
23
24      Gillette, D. A.; Adams, J.; Endo,  A.; Smith, D.; Kihl, R. (1980) Threshold  velocities for input of soil particles
25             into the air by desert soils. J.  Geophys. Res. 85: 5621-5630.
26
27      Granat, L.; Rodhe, H.; Hallberg, R. O. (1976) The global sulphur cycle. In: Svensson,  B. H.; Soderlund, R.,
28             eds. Nitrogen, phosphorus and sulphur: global cycles. Ecol. Bull. 22: 89-134. (SCOPE report 7).
29
30      Grosjean, D.;  Seinfeld, J. H. (1989) Parameterization of the formation potential of secondary organic aerosols.
31             Atmos. Environ. 23:  1733-1747.
32
33      Hansen, A. D. A.; Rosen, H. (1990) Individual measurements of the emission factor of aerosol black carbon in
34             automobile plumes. J. Air  Waste Manage. Assoc. 40: 1654-1657.
35
36      Hildemann, L. M.; Markowski, G. R.; Jones,  M. C.; Cass, G. R.  (1991) Submicrometer aerosol mass
37             distributions of emissions from boilers, fireplaces, automobiles, diesel trucks, and meat-cooking
38             operations. Aerosol Sci. Technol.  14: 138.
39
40      Houck, ]. E.;  Chow, J. C.; Watson, J.  G.; Simons, C. A.; Pritchett, L. C.;  Goulet, J. M.; Frazier, C. A.
41             (1989) Determination of particle size distribution and chemical composition of particulate matter from
42             selected sources in California. Sacramento,  CA; California  Air Resources Board.
43
44      Houck, J. E.;  Goulet, J. M.; Chow, J. C.; Watson, J. G.; Pritchett, L. C. (1990) Chemical  characterization of
45             emission sources contributing to light extinction. In:  Mathai, C. V., ed. Visibility and fine particles.
46             Pittsburgh, PA: Air &  Waste  Management Association; pp. 437-446.
47
48      Ingalls, M. N.; Smith, L. R.; Kirksey, R. E. (1989) Measurement  of on-road vehicle emission factors in the
49             California South Coast Air Basin:  volume I, regulated emissions.  Atlanta, GA: Coordinating Research
50             Council, Inc.; report no. CRC-APRAC-AP-4-SCAQS-l. Available from: NTIS,  Springfield, VA; PB89-
51             220925.
52
53      Kinsey, J. S.;  Cowherd, C. (1992) Fugitive emissions. In: Buonicore, A. J.;  Davis, W.  T., eds. Air pollution
54             engineering manual. Pittsburgh, PA: Air & Waste Management Association; pp. 133-146.


         April 1995                                    5-50       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Knapp, K. T. (1992) Dynamometer testing of on-road vehicles from the Los Angeles in-use emissions study.
  2            In: Chow, J. C.; Ono, D. M., eds. PM10 standards and nontraditional particulate source controls, v. II.
  3            Pittsburgh, PA: Air & Waste Management Association; pp. 871-884.
  4
  5     Kowalczyk, G. S.; Choquette, C.  E.; Gordon, G. E. (1978) Chemical element balances and identification of air
  6            pollution sources in Washington, D.C. Atmos. Environ.  12: 1143.
  7
  8     Lamb, B.; Gay, D.; Westberg, H.; Pierce, T. (1993) A biogenic hydrocarbon emission inventory for the U.S.A.
  9            using a simple  forest canopy model. Atmos. Environ.  Part A 27: 1673-1690.
 10
 11     Landry,  B.; Liu, C. S.; Henry, R. C.; Cooper, J. A.; Sherman, J. R. (1988) Receptor modeling for PM10
 12            source apportionment in the south coast air basin of California. In: Mathai, C. V.; Stonefield, D. H.,
 13            eds. PM10: implementation of standards. Pittsburgh, PA: Air Pollution Control Association; pp. 399-418.
 14
 15     Lebowitz, L. G.; Ackerman, A. S. (1987) Flexible regional emissions data system (FREDS) documentation for
 16            the 1980 NAPAP (National Acid Precipitation Assessment Program) emissions inventory. Research
 17            Triangle Park,  NC: U.S. Environmental Protection Agency, Air and Energy Engineering Research
 18            Laboratory; EPA report no. EPA/600/7-87/025A. Available from: NTIS, Springfield, VA; PB88-129499.
 19
 20     Magliano, K. L.  (1988) Level I PM10 assessment of a California air basin. In: Mathai,  C. V.; Stonefield, D. H.,
 21            eds. Transactions, PMi0: implementation standards. Pittsburgh, PA: Air and Waste Management
 22            Association,; pp. 508-517.
 23
 24     Mollinger, A. M.; Nieuwstadt, F. T.  M.; Scarlett, B.  (1993) Model experiments of the resuspension caused by
 25            road  traffic. Aerosol Sci. Technol. 19: 330-338.
 26
 27     Muleski, G.; Stevens, K. (1992) PM10 emissions from public unpaved roads in rural Arizona. In: Chow, J. C.;
 28            Ono, D. M., eds. Transactions, PM10 standards and nontraditional particulate source controls. Pittsburgh,
 29            PA: Air and Waste Management Association; pp. 324-334.
 30
 31     NBA. (1990) Enhancement of the  south coast air basin source profile library for chemical mass balance  receptor
 32            model application. Beaverton, OR: NBA.
 33
 34     Nicholson, K. W.; Branson, J. R.; Giess, P.; Cannell, R. J. (1989) The effects of vehicle activity on particle
 35            resuspension. J. Aerosol Sci.  20: 1425-1428.
 36
 37     Ondov, J. M. (1974) A study of trace elements on particles from motor vehicles. College Park, MD: University
 38            of Maryland.
 39
 40     Pace, T. G. (1987) Protocol for applying and validating the CMB model. Research Triangle Park, NC: U.S.
 41            Environmental  Protection Agency; report no. EPA-450/4-87-010.
 42
 43     Parmar, S. S.; Grosjean,  D. (1990) Laboratory tests of KI and alkaline annular denuders. Atmos. Environ. Part
44           A 24: 2695-2698.
45
46     Pierson, W. R.;  Brachaczek, W. W.  (1974) Airborne particulate debris from rubber tires. Rubber Chem.
47           Technol. 47:  1275.
48
49     Pierson, W. R.;  Brachaczek, W. W.  (1976) Particulate matter associated with vehicles on the road. Presented at:
50           the automotive  engineering congress and exposition; February; Detroit,  MI. Warrendale, PA: Society of
51            Automotive Engineers; SAE technical paper no. 760039.
52
53      Pierson, W. R.;  Brachaczek,  W. W.  (1983) Particulate matter associated with vehicles on the road. II. Aerosol
54           Sci. Technol. 2: 1-40.


         April 1995                                     5.51        DRAFT-DO NOT QUOTE OR CITE

-------
 1     Pierson, W. R.; Gertler, A. W.; Bradow, R. L. (1990) Comparison of the SCAQS tunnel study with other on-
 2            road vehicle emission data. J. Air Waste Manage. Assoc. 40:  1495-1504.
 3
 4     Pye, K. (1987) Aeolian dust and dust deposits. London, United Kingdom: Academic Press.
 5
 6     Reeser, W. K.; Zimmer, R. A.; Cummins, P.; Briggs, K. R. (1992) An episodic PM10 emissions inventory for
 7            Denver, Colorado. In: Chow, J. C.; Ono, D. M., eds. Transactions, PM10 standards and nontraditional
 8            paniculate source controls. Pittsburgh, PA:  Air and Waste Management Association; pp. 131-145.
 9
10     Robinson, E.; Robbins, R. C. (1971) Emissions,  concentrations, and fate of paniculate atmospheric pollutants.
11            Final report. Washington,  DC: American Petroleum Institute; API publication no. 4076.
12
13     Ryan, W. M.; Badgett-West, C. R.; Cooper, J. A.; Ono, D. M. (1988) Reconciliation of the receptor and
14            dispersion modeling impacts of PM10 in Hayden, AZ. In: Mathai, C. V.; Stonefield, D. H., eds.
15            Transactions, PM10: implementation of standards. Pittsburgh,  PA: Air and Waste Management
16            Association; pp. 419-429.
17
18     Seinfeld, J. H. (1986) Air pollution: physical and chemical  fundamentals. 2nd ed. New York, NY: McGraw Hill.
19
20     Solomon, P. A. (1994) Planning and managing air quality modeling and measurements studies: a perspective
21            through SJVAQS/AUSPEX. Chelsea, MI: Lewis Pubishers; San Ramon, CA: Pacific Gas and Electric
22            Company; pp. 369-382.
23
24     South Coast Air Quality Management District. (1994) State implementation plan for PM10 in the Coechella
25            Valley: 1994 "BACM" revision. Diamond Bar, CA: South  Coast Air Quality Management District.
26
27     Stedman, D. J.; Bishop, G. A.; Beaton, S.  P.; Peterson, J. E.; Guenther, P. L.; McVey, I. F.; Zhang, Y.
28            (1994) On-road remote sensing of CO and HC emissions in California. Sacramento, CA: California Air
29            Resources Board; final report under contract A032-093.
30
31     Svasek, T. N.; Terwindt, J. H. (1974) Measurement of sand transport by wind on a natural beach.
32            Sedimentology 21:311 -322.
33
34     Systems Applications, Inc. (1990)  User's guide for the urban airshed model volume IV: user's guide for the
35            emissions preprocessor system. San Rafael, CA: Systems Applications, Inc.
36
37     Thanukos, L. C.; Miller, T.;  Mathai, C. V.; Reinholt, D.; Bennett, J. (1992) Intercomparison of PM10 samplers
38            and source  apportionment  of ambient PM10 concentrations in Rillito, Arizona. In:  Chow, J. C.;  Ono, D.
39            M., eds. Transactions, PM10 standards and nontraditional particulate source controls. Pittsburgh, PA: Air
40            and Waste Management Association; pp.  244-261.
41
42     Thornwaite, C. W. (1931) the climates of North America: according to a new classification. Geophys. Rev.
43            21:633-655.
44
45     U.S. Environmental Protection Agency. (1980a) National air pollutant emission estimates, 1970-1978. Research
46            Triangle Park, NC: Office of Air Quality Planning and Standards; EPA report no. EPA-450/4-80-002.
47            Available from: NTIS, Springfield, VA; PB82-232687.
48
49     U.S. Environmental Protection Agency. (1980b)  1977 National emissions report: national emissions data system
50            of the Aerometric and Emissions Reporting System. Research Triangle Park, NC: Office of Air Quality
51            Planning and Standards; EPA report no. EPA-450/4-80-005. Available from: NTIS, Springfield, VA;
52            PB80-222532.
53
         April 1995                                    5-52        DRAFT-DO NOT QUOTE OR CITE

-------
  1     U.S. Environmental Protection Agency. (1982) Air quality criteria for paniculate matter and sulfur oxides.
  2            Research Triangle Park, NC: Office of Health and Environmental Assessment, Environmental Criteria
  3            and Assessment Office; EPA report no.  EPA-600/8-82-029aF-cF. 3v. Available from: NTIS, Springfield,
  4            VA; PB84-156777.
  5
  6     U.S. Environmental Protection Agency. (1987a) PM10 SIP development guideline. Research Triangle Park, NC:
  7            U.S. Environmental Protection Agency, Office of Air Quality Planning and Standards; EPA report no.
  8            EPA-450/2-86-001. Available from:  NTIS, Springfield, VA; PB87-206488.
  9
 10     U.S. Environmental Protection Agency. (1987b) Protocol  for reconciling differences among receptor and
 11            dispersion models. Research Triangle Park, NC: U.S. Environmental Protection Agency, Office of Air
 12            Quality Planning and Standards; EPA report no. EPA-450/4-87-008. Available from: NTIS, Springfield,
 13            VA; PB87-206504.
 14
 15     U.S. Environmental Protection Agency. (1988) Compilation of air pollutant emission factors. Volume  I:
 16            stationary point and area sources. Research Triangle Park, NC: Office of Air Quality Planning and
 17            Standards, Office of Air and Radiation.
 18
 19     Watson, J. G., Jr. (1979) Chemical element  balance receptor methodology for assessing the sources of fine and
 20            total suspended paniculate matter in  Portland, Oregon [Ph.D. thesis]. Beaverton, OR: Oregon  Graduate
 21            Center.
 22
 23     Watson, J. G.; Chow, J. C.; Richards, L. W.; Neff, W.  D.; Andersen, S. R.; Dietrich, D. L.; Houck, J. E.;
 24            Olnez, I. (1988a) The 1987-88 metro Denver brown cloud study: v. I, II, III. Reno, NV: Desert
 25            Research Institute; final report no. 8810.1F(l-3).
 26
 27     Watson, J. G.; Chow, J. C.; Egami, R. T.;  Frazier, C. A.;  Goodrich, A.; Ralph, C. (1988b) PM10 source
 28            apportionment in Reno and sparks, Nevada for state  implemenation plan development. Volume I:
 29            modeling methods  and results. Reno, NV: Desert Research Institute; document 8086.2F1.
 30
 31     Watson, J. G.; Chow, J. C.; Mathai, C. V.  (1989) Receptor models in air resources management: a summary of
 32            the APCA international specialty conference. JAPCA 39: 419-426.
 33
 34     Watson, J. G.; Chow, J. C.; Pritchett, J. A.; Houch, R. A.; Ragazzi, R.  A.; Burns, S. (1990) Chemical  source
 35            profiles for paniculate motor vehicle exhaust under cold and high  altitude operating conditions. Sci. Total
 36            Environ. 93:  183-190.
 37
 38     Watson, J. G.; Chow, J. C.; Pritchett, L. C.; Houck, J. A.; Ragazzi, R. A.; Burns, S. (1990) Composite source
 39            profiles for paniculate motor vehicle exhaust source  apportionment in Denver, CO. In: Mathai, C. V.,
40            ed. Transaction, visibility and  fine particles. Pittsburgh, PA: Air and Waste Management Association;
41            pp. 422-436.
42
43     Watson, J. G.; Robinson, N. F.; Chow, J. C.; Henry, R.  C.; Kim, B. M.; Pace, T. G.; Meyer, E. L.; Nguyen,
44            Q. (1990) The USEPA/DRI chemical mass balance receptor model, CMB 7.0. Environ. Software
45            5: 38-49.
46
47     Watson, J. G.; Chow, J. C.; Pace, T. G. (1991) Chemical mass balance.  In: Hopke, P. K., ed. Data handling in
48            science and technology: v. 7, receptor modeling for air quality management. New York, NY: Elsevier
49            Press;  pp. 83-116.
50
51     Watson, J. G.; Chow, J. C.; Lowenthal, L. C.; Pritchett,  C. A.; Frazier,  C. A.; Neuroth, G. R.; Robbins, R.
52            (1994) Differences in the carbon compostion of source profiles for diesel- and gasoline-powered vehicles.
53            Atmos. Environ. 28: 2493-2505.
54


         April  1995                                    5.53        DRAFT-DO NOT QUOTE OR CITE

-------
 1     Watson, J. G.; Chow, J. C.; Lu, Z.; Fujita, E. M.; Lowenthal, D. H.; Lawson, D. R.; Ashbaugh, L. L. (1994)
 2            Chemical mass balance source apportionment of PM10 during the Southern California Air Quality Study.
 3            Aerosol Sci. Technol. 21: 1-36.
 4
 5     Watson, J. G.; Chow, J. C.; Lurmann, F. W.; Musarra, S. P. (1994) Ammonium nitrate, nitric acid, and
 6            ammonia equilibrium in wintertime Phoenix, Arizona. Air Waste 44: 405-412.
 7
 8     Watson, J. G.; Chow, J. G.; Roth, P. M. (1994) Clear sky visibility as a challenge for society. Annu. Rev.
 9            Energy Environ. 19: 241-266.
10
11     Went, F. W. (1960) Organic matter in the atmosphere, and its possible relation to petroleum formation. Proc.
12            Natl. Acad. Sci. U.S. A. 46: 212-221.
13
14     Yamate, G. (1973) Development of emission factors for estimating atmospheric emissions from forest fires.
15            Research Triangle Park, NC: U.S. Environmental Protection Agency, Office of Air Quality Planning and
16            Standards; EPA report no. EPA-450/3-73-009. Available from: NTIS, Springfield, VA; PB-230 889.
17
18     Zeldin, M.; Kim, B.-M.; Lewis, R.; Marlia, J. C.; Chan, S.; Hogo, H.; Uesugi, K. (1991)  PM10 source
19            apportionment for the South Coast Air Basin: final air quality management plan, 1991 revision. Diamond
20            Bar, CA: South Coast Air Quality Management District; final technical report V-F.
         April 1995                                   5-54       DRAFT-DO NOT QUOTE OR CITE

-------
  i               6.   ENVIRONMENTAL CONCENTRATIONS
  2
  3
  4      6.1   BACKGROUND, PURPOSE AND SCOPE
  5           This chapter summarizes the concentrations of particulate matter over the United States,
  6      including the spatial, temporal, size and chemical aspects.  This chapter is aimed to support
  7      the quantification of particulate matter effects and to aid the PM standard setting process.
  8      The information needs for assessing the major aerosol effects of concern is summarized in
  9      Table 6.1.  Depending on the effect, different aspects (dimensions) of aerosol concentrations
10      are important.  The effects on human health are considered most serious and this chapter is
11      to provide relevant aerosol concentration information to help in quantifying these effects.
12      Concern also exists for aerosol effects on visibility as well as damage to manmade  materials.
13           Health effects are concerned with people, and the  geographic areas of importance are
14      those with high population densities.  The spatial resolution and detail needed for health
15      effect assessment is rather high, because of strong gradients in population densities. Health
16      effects are believed to occur both as a consequence of short-term acute episodic exposure, as
17      well as through cumulative chronic long-term exposure.  The relevant particle sizes are in the
18      inhalable size range (< 10 /mi), but it is known that submicron particles penetrate  deeper
19      into the lungs.  It is not well known which ambient aerosol chemical species are most potent
20      in causing health effects.  However, sulfates, particularly in acidic form are believed to be
21      important along with toxic trace metals and carcinogenic organic substances.  Also, health
22      damage is usually the consequence  of the combined influence of multiple, coexisting
23      pollutants, weather, and other environmental conditions.
24           Preventing the degradation of visibility, particularly in pristine national parks  has been
25      one of the provisions of the Clean Air Act.  The main cause of visibility degradation is
26      atmospheric haze that is contributed mainly  by fine particles, except during dust events.
27      Sulfates, organics, nitrates, absorbing carbon and to some extent dust particles are the
28      contributors to  visibility degradation. Both  short term fine particle episodes, seasonal pattern
29      and long term fine particle trends are relevant to visibility degradation.  The potential
30      climatic effects of aerosols are influenced by roughly the same factors that determine the
31      visibility degradation except on a global scale.
        April 1995                                6-1       DRAFT-DO NOT QUOTE OR CITE

-------
            TABLE 6.1  AEROSOL INFORMATION NEEDS FOR ASSESSING EFFECTS

Concern
Space Resolution
Time Resolution
Particle Size
Chemistry
Health
People
Local
Short, Long
Inhalable (Fine?)
Acidity, Toxicity
Visibility
National Parks
Regional
Short, Long
Fine
SO4, NO3, Organics
Materials
Damage
Costs
Local
Long
Fine
Acidity, Soot
 1            Aerosol effects on man-made materials include soiling and corrosion. These materials
 2      are located mainly in populated areas  and high spatial resolution for concentrations is needed.
 3      Soiling is due to carbonaceous smoke and soot and settling dust while corrosion increases are
 4      due to acidifying sulfurous aerosols. The materials damage occurs over the period of years,
 5      but daily and seasonal cycles are  also important. The quantification of materials damage is
 6      particularly sensitive to  the interaction with weather elements, particularly moisture.
 7            Other regional and global aerosol effects include acid  deposition and effects on
 8      climate. Acid deposition and its relationship to aerosols is treated extensively elsewhere
 9      (NAPAP, 1991). Direct aerosol perturbation of the radiative climate and the indirect aerosol
10      influence through changing the cloud  properties and pattern is beyond the scope of this
11      chapter.
12            The commonality among these  effects (Table 6.1) is that  the overall damage is driven
13      by the concentration of  relevant aerosol parameters, the spatial pattern and density of
14      receptors and by the receptors' sensitivity.  The receptor densities, e.g. population densities,
15      national parks, etc. are  not discussed  here.  The sensitivity (damage functions) are treated in
16      the respective chapters on aerosol effects.
17
18      6.1.1   Dimensionality  and  Structuring of the Aerosol Data  Space
19            Aerosol concentration patterns  contain endless detail and  complexity in space, time,
20      size,  and chemical composition. Chemically analyzed aerosol samples over the conterminous
21      United States reveal the coexistence of sulfates, secondary organics, nitrates, smoke,  soot,
22      soil dust, sea salt, and trace metals in most aerosol samples.  Each of these chemicals is also
23      distributed in different size particles. This chemically rich aerosol mix arises from the
        April 1995
6-2
DRAFT-DO NOT QUOTE OR CITE

-------
  1      multiplicity of contributing aerosol sources, each having a unique chemical mixture for the
  2      primary aerosol at emission time.  The primary aerosol chemistry is further enriched by the
  3      addition of species during atmospheric chemical processes. Finally, the immensely effective
  4      mixing ability of the lower troposphere stirs these primary and secondary particles into a
  5      mixed batch with various degrees  of homogeneity, depending on location and time.  The
  6      result is a spatial temporal, size, and chemically heterogeneous aerosol pattern that is
  7      probably unparalleled in the domain of atmospheric sciences.
  8            In order to characterize the aerosol pattern  that is useful for effects assessment, it is
  9      necessary to organize and structure the aerosol pattern analysis.  This  "user-driven" aspect of
 10      aerosol concentration structuring demands that it be consistent with the information needs
 11      stated in Table 6.1.
 12            Another consideration in structuring the aerosol pattern analysis is  that it has to be
 13      consistent with physical and chemical processes that determine the concentrations. The
 14      principles of atmospheric sciences state that the concentration of particulate matter, (C) at
 15      any given location and time is determined by the combined interaction of emissions, (E),
 16      dilution,  (D),  and chemical transformation and removal,  (T), processes expressed as:
 17
 18                                           C  = f (D,T,E)
 19
20            Dilution, transformation/removal and emissions, D,T,E are generic operators and can,
21      in principle, be determined from suitable measurements and models.  However, for
22      consideration of aerosol pattern analysis it is sufficient to  recognize and separate these three
23      major causal factors influencing the aerosol concentration pattern.
24            This section outlines the main organizing principles for the analysis of PM pattern. It
25      is convenient to categorize the highly variable aerosol signal along the following major
26      dimensions: space, tune, size and chemical composition.  The space and time dependence of
27      concentrations area common to all pollutants.  However, both the distribution with respect to
28      particle size as well as the chemical distribution within a given size range constitute unique
29      dimensions of particulate matter that is not present for other pollutants. The  concentration of
30      single-compound gaseous pollutants can be fully characterized by their spatial and temporal
31      pattern.  This classification by dimensions is consistent with the size-chemical composition

        April 1995                                6-3        DRAFT-DO NOT QUOTE OR CITE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
distribution function introduced by Friedlander (1976).  It could be said that particulate
matter is a composite of hundreds of different pollutants having a distribution in space and
time.

6.1.2   Spatial Pattern and Scales
      The spatial dimension covers  the .geographic scale and pattern of aerosols.  Based on
consideration of emissions, meteorology, and political boundaries, the spatial dimension can
be classified as global,  national, regional-synoptic, meso, urban, and local. Some of the
characteristics of these  spatial scales are illustrated in Table 6-2.

                   TABLE 6-2.  SPATIAL  REGIONS AND SCALES
Global
Continent
10,000 -
50,000 km
National
Country
5,000 -
10,000 km
Regional
Multi-state
1,000-
5,000 km
Meso
State
100-
1,000km
Urban
County
10 - 100 km
Local
City-center
1 - 10km
 1     6.1.3   Temporal Pattern and Scales
 2           The time dimension of aerosols extends over at least six different scales (Figure 6-1).
 3     A significant, unique feature of the temporal domain  is the existence of periodicities.  The
 4     secular time scale extends over several decades or centuries.  Given climatic and chemical
 5     stability of the atmosphere the main causes of secular concentration trends are changes in
 6     anthropogenic emissions.  The yearly scale is imposed by seasonal variation of solar
 7     radiation.  Emissions, atmospheric dilution, as well as chemical/removal processes are
 8     influenced by the seasonal cycle.  The weekly periodicity is unique among the time scales in
 9     that it is imposed exclusively by human-induced emission changes.  The synoptic scale
10     covers the duration of synoptic meteorological events (3-5 days). Its role is primarily
11     reflected in dilution and chemical/removal processes.   The daily cycle is again imposed by
12     solar radiation and  it strongly influences the emissions, dilution, and chemical/removal
13     processes. Microscale defines variation of the order  of an hour caused by short-term
14     atmospheric phenomena.  In the analysis that follows we will emphasize  secular trends and
        April 1995
                                          6-4
DRAFT-DO NOT QUOTE OR CITE

-------
Emissions
X
Dilution
X
Chemistry /Removal
=
Concentration
             Secular
              Yearly
             Weekly
             Synoptic
                Daily
           Micro-scale
                           10-100 yrs
                            Minutes
1
2
3
4
5
6
Figure 6-1.  Time scales for particle emissions.

yearly cycles, with some consideration of daily aerosol pattern.  The microscale patterns will
be largely ignored.

6.1.4    Space-Time Relationships
      The spatial time scales of aerosol pattern are linked by the atmospheric residence time
of particles.  Short residence tunes restrict the aerosol to a short transport distance from a
      April 1995
                                         6-5       DRAFT-DO NOT QUOTE OR CITE

-------
 1      source, causing strong spatial and temporal gradients.  Longer residence times yield more
 2      uniform regional pattern caused by long range transport.  The relationship between spatial
 3      and temporal scales for coarse and fine particles is illustrated in Figure 6-2.
 4            The aerosol residence time itself is determined by the competing rates of chemical
 5      transformations and removal rates.   Secondary aerosol formation tends to be associated with
 6      multi-day long range transport because of the time delay necessary for the formation.  For
 7      sulfates, for example, the residence time is 3-5 days.  For fine particles, 0.1 jurn and above,
 8      the main removal mechanism involves cloud processing, while coarse particles above 10 ptm
 9      are deposited by sedimentation.  Ultrafine particles, below 0.1 /*m also coagulate to form
10      particles in the 0.1 to 1.0 /mi size range.  As a consequence of low removal rates,  aerosols
11      in 0.1-1.0 pirn size range reside in the atmosphere for longer periods than either smaller or
12      larger particles (Figure 6-3).  If aerosols are lifted into the mid- or upper-troposphere their
13      residence time will increase to several weeks.  Large scale aerosol injections into the
14      stratosphere through volcanoes or deep convection extend their atmospheric  residence to  1-2
15      years.
16            In the context of the specific analysis that follows, the  space-time-concentration
17      relationship  in urban and mountainous areas is of particular importance (Figure 6-4). Urban
18      areas have strong spatial emission gradients and also corresponding concentration gradients,
19      particularly in the winter under poor horizontal and vertical transport conditions.  In the
20      summer most urban areas have  similar concentrations to their non-urban background.
21            In mountainous regions, the strong concentration gradients  are caused by both
22      topography that limits transport as well as the prevalence of emissions in valley floors.
23      Strong wintertime inversions tend to amplify the valley-mountain  top concentration
24      difference.   Fog formation also accelerates the formation of aerosols  in valleys.
25
26      6.1.5   Particle Size Distribution
27            The aerosol size distribution is of importance in quantifying both the  formation
28      (generation) as well as the effects of aerosols.   Condensation  of gaseous substances during
29      combustion in the atmosphere generally produces fine particles below 1 jum in diameter.
30      Forced resuspension of soil dust and dispersion of sea spray produces coarse particles above
31      1
        April 1995                                6-6        DRAFT-DO NOT QUOTE OR CITE

-------
                         108-
                     c/)
                     az
                     uu
                     i-
                     LJ
                     UJ
                     u
                     I
                     CO
                     cr
                     o
                     Q.
                     ce
                                        i	1
                                        Global
                                       C02,CH4
                               Synoptic-
                               Regional
                             SOx.NOx.Os
                             ine Particles
                               <2um)
                     Mesoscale
                    NO.NOz, 03
                      Coarse
                     Particles
                   20pm)
                         101
                                   101     102     103    104    105    106
                                       RESIDENCE TIME,  SECONDS
                                              107
       Figure 6-2. Relationship of spatial and temporal scales for coarse and fine particles.
 1           The size distribution of particles also influences both the atmospheric behavior and the
 2     effects of aerosols.  Atmospheric coagulation, cloud scavenging,  and removal by impaction
 3     and settling are strongly size dependent.  The effects on human health depend on
 4     size-dependent lung penetration.  Light scattering hi visibility and climatic effects  is also
 5     strongly dependent on particle size.
 6           Measurements over the past decades (Whitby et al,  1972, Whitby 1978) show that
 7     atmospheric aerosols may be classified as fine or coarse particles. The size distribution of
 8     atmospheric particles is discussed hi Section 3.7. The sources, formation mechanisms, and
 9     the chemical composition of these two aerosol modes are different. In general, the two
10     aerosol size modes have independent spatial and temporal pattern as described throughout this
11     chapter.  Coarse, dust particles tend to be more variable hi space and tune and can be
12     suspended through natural or human-induced activities.  Fine particles are largely  of
       April 1995
              6-7
DRAFT-DO NOT QUOTE OR CITE

-------
                                   10'3  10'2  10'1   10°   101   102   103
                                     RADIUS .
       Figure 6-3.  Residence time in the lower troposphere for atmospheric particles between
                   0.1 and 1.0 /im.
       Source:  Jaenicke, 1980
 1     secondary origin and their spatial-temporal pattern is more regional. Notable exceptions are
 2     urban-industrial hotspots and mountain valleys where primary submicron size smoke particles
 3     can prevail.
 4
 5     6.1.6   Aerosol Chemical Composition
 6           The chemical composition of atmospheric aerosol is believed to  influence the effect on
 7     human health.  While the causal mechanisms are not fully understood,  the acidity,
 8     carcinogenicity, and other forms  of toxicity are chemical properties considered relevant to
 9     human health.
10           The aerosol chemical composition has also become an important property for
11     identifying source types based on chemical "fingerprints" in the ambient aerosol.  Since
       April 1995                                6-8       DRAFT-DO NOT QUOTE OR CITE

-------
        Summer
        rural
        Winter
                       urban
        rural
                        urban
                                           rural
                                           rural
                                                       Summer
                                                          mountain
                                                       Winter
valley
                                                         mountain
valley
                                                                                      mountain
                                                                                     mountain
        Figure 6-4. Space-time relationship in urban and mountainous areas.
 1      aerosols reside in the atmosphere for days and weeks, there is a substantial amount of mixing
 2      that takes place among the contributions of many sources. At any given "receptor"  location
 3      and time, the aerosol is a mixture of tens or hundreds of source contributions each having a
 4      chemical signature for possible source type identification.
 5            Fine particles  are generally composed of sulfates, organics, nitrates,  elemental carbon
 6      (soot), as well as trace metals  (Section 6.6). Each major chemical species have sub-species
 7      such as acidic and  neutral sulfates, light and heavy organics, ammonium and sodium nitrates,
 8      etc.
 9            The chemical composition of coarse particles is dominated by the elements of the earth
10      crust, Si, Al, Fe, suspended from soil. Near roadways, coarse particles may be
11      contaminated by lead and other trace metals.  At ocean shores, coarse particles may consist
12      of sea salt arising from breaking of waves.  Both resuspended dust and sea  salt are primary
13      particles, carrying  the chemical signature of then- sources.
        April 1995                                 6-9       DRAFT-DO NOT QUOTE OR CITE

-------
 1     6.1.7   Chapter Organization and Approach
 2            The general approach in preparing this chapter was to organize, evaluate, and
 3     summarize the existing large scale aerosol data sets over the United States.  Emphasis was
 4     placed on complete national coverage  as well as the fusion and reconciliation of multiple data
 5     sets.  The aerosol concentrations are considered from the point of view of spatial, temporal,
 6     size and chemical pattern.
 7            The main organizing dimension used to structure this chapter is space.  The
 8     concentrations are presented on global, continental, national, regional, and
 9     sub-regional/urban scales.  Within each spatial domain, the spatial-temporal structure,  aerosol
10     size and chemical composition  is presented. The presentation of aerosol pattern begins with a
11     global and continental perspective  (Section 6.2).  Next, the national aerosol patterns are
12     examined (Section 6.3) as derived from non-urban and urban PM10 and PM2 5 monitoring
13     networks.  In Section 6.4 the aerosol  characteristics over seven  subregions of the
14     conterminous United States are examined in more detail. The ten year trends,  seasonal
15     patterns, as well  as the PM2 s/PM10 relationship and fine particle chemical composition is
16     examined for each region.  Section 6.5 focuses further on the sub-regional and urban-scale
17     aerosol pattern over representative areas of the United States.
18            The aerosol concentration pattern over the United States has been reported by many
19     aerosol researchers over the past decade. In particular the research groups  associated  with
20     the IMPROVE aerosol monitoring networks have been prolific producers of high quality
21     data, reports, and analysis of non-urban data.  This section draws heavily on their
22     contribution but the maps, charts,  and computations have been re-done for sake of
23     consistency with other (urban)  data from the AIRS network. Each of the sections are
24     augmented by suitable but not complete references to the pertinent-literature.
25
26
27
28
29
30
31

        April 1995                               6-10       DRAFT-DO NOT  QUOTE OR CITE

-------
 1      6.2  CONTINENTAL AND GLOBAL AEROSOL PATTERNS
 2            There are two data sets which can be used to provide information on fine particles
 3      concentration  patterns on a continental and global scale.  Routine visibility distance
 4      observations,  recorded hourly at many U. S. airports by the U. S. Weather Service,
 5      provide an indication  of fine particles pollution over the United States.  The visibility
 6      distance data  has been converted to aerosol extinction coefficient and used to access
 7      patterns and trends of aerosol pollution over the U.S. (Husar et al., 1994). Routine
 8      satellite monitoring of backscattered solar radiation  over the oceans by the Advanced
 9      Very High Resolution Radiometer  sensors  on polar orbiting  meteorological satellites
10      provides a data set which can be used to give an indication of aerosol pollution over the
11      world's oceans (Husar and Stowe,  1994).
12            Aerosol detection over the oceans is facilitated by the fact the ocean reflectance
13      at 0.6 fjm is only 2%. Hence, even small  backscattering from aerosols produces a
14      measurable aerosol signal. The backscattering is converted to a vertically integrated
15      equivalent aerosol optical thickness assuming a shape for the aerosol size distribution
16      or phase function. Clouds are eliminated by a cloud mask, so the data are biased toward
17      clear-sky conditions.  The oceanic aerosol  maps represent a two-year average (July
18      1989-June 1991) prior to the eruption of Mt. Pinatubo,  while  the stratosphere was
19      unusually clear from aerosol. Consequently,  the images represent the  spatial pattern of
20      tropospheric aerosol.
21            These  two data sets may  then  be merged  to  provide  a  continental-scale
22      perspective. Some results for North America are shown in Figures 6-5a, 6-5b, and 6-5c.
23      The oceanic aerosol for the entire globe is shown seasonally in Figure 6-6. The average
24      aerosol map of Eastern North America for June, July and August (Figure 6-5a)  shows
25      areas of high  optical depth over the Mid-Atlantic States and over the Atlantic Ocean.
26      The oceanic aerosol concentration  is higher near the coast and declines with distance
27      from the coast. This indicates that the aerosol is of continental origin and represents the
28      plume of Eastern North America heading north-east across the Atlantic ocean.  This
29      plume can also be seen in the spring and  summer season oceanic  aerosol patterns
30      shown in Figure 6-6.
       April 1995                             6_n      DRAFT-DO NOT QUOTE OR CITE

-------
                      JUNE, JULY, AUGUST
                                           HUSAR AND STOWE, 1994
                                                            X
 Figure 6-5a.  Continental scale pattern of aerosols derived from visibility
          observations over land and satellite monitoring over the oceans:
          Eastern North America.

 Source:  Husar and Stowe, 1994
April 1995
6-12      DRAFT-DO NOT QUOTE OR CITE

-------
                       Mar.,  Apr.,  May.
 Figure 6-5b. Continental scale pattern of aerosols derived from visibility
          observations over land and satellite monitoring over the oceans:
          Western North America.

 Source:  Husar and Stowe, 1994
April 1995
6-13
DRAFT-DO NOT QUOTE OR CITE

-------
 2.
 c/i
ON
H
6
o
o
c!
§
                                            Mar.,  Apr.,   May.
n
HH
H
W
Figure 6-5c.  Continental scale pattern of aerosols derived from visibility observations over land and satellite monitoring
         over the oceans: Southern North America.
     Source: Husar and Stowe, 1994

-------
 1           The continental aerosol extinction coefficient data for the southwest coast of North
 2     America indicate elevated aerosol extinction over southern California. The area includes
 3     the hazy South Coast and the San Joaquin Valley air basins. It is interesting to note that
 4     somewhat elevated aerosol optical  thickness is also recorder over the Pacific near
 5     Southern California.  However, the low aerosol signal and the semi-quantitative satellite
 6     data preclude a clear cause-effect association.
 7           The average aerosol map for Southern North America (Figure 6-5c) April, May
 8     and June shows that the oceans adjacent to southern Mexico have high aerosol optical
 9     thickness, both on the Gulf side and the Pacific side. The aerosol concentration is higher
10     near the coasts and declines toward the  sea. This  indicates that the aerosol is of
11     continental origin,  over southern Mexico.  The haze  off the Mexican coasts is most
12     pronounced in the spring season.  Visibility observations at meteorological stations also
13     indicate a spring maximum in horizontal  extinction. The  region is known for extensive
14     springtime slash burning.  Photographs taken by astronauts show numerous areas of
15     biomass burning in the spring season. Visible geostationary satellite images taken in the
16     spring  also  show  the haze plumes emanating from  southern Mexico. However  the
17     composition and sources of the Mexican haze are not established.
18           The seasonal aerosol pattern over the oceans  reveals that the highest aerosol
19     signal is near the tropics, where wind-blown dust and  biomass  combustion from Africa
20     and southern Asia produce 5,000 km long aerosol plumes (Figure 6.6).  Further aerosol
21     belts of marine  origin are observed just north of the Equator and at 30 to 60° latitudes
22     in both hemispheres. The backscattering in the summer hemispheres exceed the winter
23     values by a factor of 5 to 10. There is a pronounced seasonality in each aerosol region
24     (Figure 6-7); the higher aerosol  levels appear in the summer hemispheres although
25     many continental and marine aerosol regions show a spring maximum. Thus, the global
26     tropospheric aerosol is a dynamic collection of independent aerosol regions, each having
27     unique sources and temporal pattern.
28           The seasonal oceanic aerosol maps show two distinctly different spatial patterns:
29     aerosol plumes originating  from  continents,  and  oceanic aerosol patches that  are
30     detached from the continents. The continental aerosol plumes are characterized by high
      April 1995                             6_16      DRAFT-DO NOT QUOTE OR CITE

-------
                                           70 Q^ N. Zeland
                                                    N Hemisphere
                          f
Figure 6-7.   Seasonal pattern of oceanic aerosols derived from satellite observations.




 Source: Husar and Stowe, 1994
 April 1995
6-17
DRAFT-DO NOT QUOTE OR CITE

-------
 1     values near the coastal areas and a decline with distance from the coast. The most
 2     prominent aerosol plume is seen over the equatorial Atlantic, originating from West Africa
 3     and crossing the tropical Atlantic.  It is the well known Sahara dust plume. Additional
 4     continental  plumes emanate form Southwest Africa, Indonesia, China-Japan, Central
 5     America and  eastern North America.  Aerosols of marine origin dominate large zonal
 6     belts (30 to 60° N and S) in the summer hemispheres as well as near the Equator.   In
 7     summary, the global tropospheric aerosol is a collection of largely in dependent aerosol
 8     regions,  each  having a bio-geochemically active source and unique spatial temporal
 9     pattern.
10           Based  on  the above  global  and  continental-scale observations,  it  can  be
11     concluded that the continental plume from eastern North America is not as intense as
12     those from other industrial and non-industrial regions of the world. However, quantitative
13     aerosol comparisons of global regions are not available.
14
15     6.3  U.S. NATIONAL AEROSOL PATTERN AND  TRENDS
16          Our current understanding of the U.S. national aerosol pattern arises from non-urban,
17     regional background monitoring networks the Interagency Monitoring of Protected Visual
18     Environments (IMPROVE) and the Northeast States for Coordinated Air Use Management
19     (NESCAUM) and from the mainly urban network,  the Aerometric Information Retrieval
20     System (AIRS).  The non-urban and urban networks yield markedly  different national
21     patterns, particularly over the western US. For this reason the results from the two sets of
22     observations are presented separately and the differences between two networks are
23     evaluated.
24          An early compilation of the chemical and  size resolved aerosol studies list 31 aerosol
25     data sets gathered since the 1970's.  However, these databases are widely dispersed,  and are
26     not generally available for study or evaluation (Chow and Watson, 1988).
27
28     6.3.1 Non-Urban National  Aerosol Pattern
29          Non-urban aerosol concentrations are measured at remote sites, away from
30     urban-industrial activities.  Size-segregated aerosol  mass and chemical composition data are
31     available for 50 sites, through the IMPROVE (Eldred et al, 1988) and NESCAUM (Poirot

       April 1995                             6-18a     DRAFT-DO NOT QUOTE OR CITE

-------
 1     et al., 1990; Flocchini et al., 1990) networks.  These are located mostly in national parks
 2     and wilderness areas.  The PM10 and PM2 5 mass concentrations are sampled and analyzed
 3     on separate filters. The sampling frequency was generally twice a week (Wednesdays and
 4     Saturdays) for 24-hours. The PM2 5 samples are analyzed for chemical composition which
 5     make the data sets suitable for chemical mass balance computations (e.g. Sisler et al., 1993).
 6     The IMPROVE/NESCAUM aerosol data are available from 1988 through 1993.
 7           The results of the national spatial and temporal pattern analysis are presented in
 8     quarterly contour maps and monthly seasonal time charts.  The contours drawn for the
 9     eastern United States are derived from only 15 to 20 stations.  As a consequence, these
10     contour lines are to be taken as guides to the eye and not as actual pattern.  The quarters of
11     the year are calendrical.
12
13     6.3.1.1 Non-urban PM2 5 Mass Concentrations
14           Maps of seasonal average  non-urban PM2 5 concentrations are shown in Figure 6-8.
15     The maps show that the country can be divided roughly into east and west halves.  The
16     eastern United States is covered by large, contiguous PM2 5 concentrations that range from
17     10 jug/m3 in Quarter 1, and 17 /xg/m3 in Quarter 3. During the transition seasons (Quarters
18     2 and 4) the eastern U.S. non-urban PM2 5 concentrations are at about 12 /xg/m3. Within the
19     eastern US,  there are subregions such as New England that have lower concentrations
20     ranging between 8 to 12 ^g/m3.
21           The lowest non-urban PM2 5 concentrations are measured over the central mountainous
22     western states.  The low winter concentrations are at about 3 /xg/m3, while the  summer
23     valuesare around 6 ftg/m3.  Somewhat elevated PM2 5 concentrations are observed over the
24     southwestern border adjacent to Mexico as well as in California and the Pacific Northwest.
25           The non-urban fine  particle mass clearly show multiple aerosol regions over the
26     conterminous US, each exhibiting a unique spatial and  seasonal  characteristics.
27
28     6.3.1.2  Non-urban PMCoarse  Concentrations
29           The non-urban coarse aerosol mass concentration in the size range 2.5 to 10 fj.m is
30     given in the seasonal maps Figure 6-9.  It is plotted on the same scale as the non-urban
31     PM2.s  and PM10 maps to show that the non-urban coarse mass concentration is less than the

       April 1995                               6-18b     DRAFT-DO NOT QUOTE OR CITE

-------
 1     fine mass concentration over most of the country.  The lowest non-urban coarse particle
 2     concentration is recorded during the first and fourth calendar quarters when virtually the
 3     entire conterminous United States showed values  < 10 ^tg/m3. It is remarkable that during
 4     the quarters 1  and 4, the  industrialized Midwest, adjacent to the  Ohio River shows low
 5     PMCoarse  concentration  (< 10 jug/m3) comparable to the pristine mountainous Rocky
 6     Mountains  states. The highest non-urban coarse mass concentrations is shown during
 7     quarters 2 and 3. In quarter 2, the southwestern United States adjacent to the Mexican
 8     border shows the highest non-urban coarse mass concentrations.  In quarter 3, the monitoring
 9     sites in Florida and Great Smoky Mountains exhibit high concentrations  (> 12 /ng/m3).
10
11     6.3.1.3 Non-urban PM10  Mass Concentrations
12            Maps of seasonal average non-urban PM10 concentrations  are shown in Figure 6-10.
13     PM10 is the sum of the PM2 5 and PMCoarse.  The spatial pattern, including the delineation
14     of aerosol regions is similar to the PM2 5.  However, the PM10 concentrations exceed the
15     PM2 5 by up to factor of two depending on region and season.
16            The  eastern U.S.  PM10 concentrations range between 12 /*g/m3 in Quarter 1, and 25
17     Mg/m3 in Quarter 3. During the transition seasons (Quarters 2 and 4) the eastern U.S.
18     non-urban PM10 concentrations are at about 15 jug/m3, except in New England.   The lowest
19     non-urban PM10concentrations are measured over the central mountainous states, 5 /xg/m3 in
20     Quarter 1,  10 ng/m3 in Quarter 3, and 7 /ig/m3 during the transition seasons. Higher PM10
21     concentrations, between 10 to 20 ^tg/m3 were  measured over the southwestern United States
22     as well as over the  Pacific  states from California to the Northwest.
23
24     6.3.1.4 PM2 5/PM10 Ratio at Non-urban Sites
25            The  PM10 aerosol mass is composed of fine mass  (PM2 5) and coarse mass, below
26     10/xm (Figure 6-10).  Both the sources and the effects of fine particles differ markedly  from
27     those of coarse particles.  For this reason it is beneficial to examine the relative contribution
28     of PM2 5 and PM10 concentrations.  Figure 6-11 shows the seasonal fine mass as a fraction
29     of PM10.
30            Nationally, the fine fraction at non-urban sites ranges between 0.4 and 0.8.  The
31     highest fine fraction is recorded east of the Mississippi River, where 75% of the PM10  mass

       April 1995                               6_2ia      DRAFT-DO NOT QUOTE OR CITE

-------
 1      is in particles <2.5 pm in size.  This is also the region that shows the highest PM10
 2      concentrations, thus fine particles dominate the non-urban aerosol concentrations east of the
 3      Mississippi River.  The fine fraction also exceeds the coarse fraction at the non-urban
 4      northwestern sites.  The fine fraction is the lowest in the southwestern United States
 5      (<  50%) particularly in the spring season (Quarter 2).  Evidently, the southwestern PM10 is
 6      dominated by coarse particles in the  spring season.
 7            Spatial and seasonal variation  of the fine fraction is a further indication for the
 8      existence of different aerosol regions over the conterminous US.  This is further illuminated
 9      in Section 6.4 where the aerosol characteristics over  different regions of the United States are
10      discussed.
11
12      6.3.1.5 Non-urban Fine Particle Chemistry
13            The chemical composition of non-urban fine particles over the conterminous  United
14      States is now reasonably well understood.   The IMPROVE/NESCAUM network provides
15      over 5 years of aerosol mass and chemical composition data. The detailed and almost
16      complete fine particle chemistry  data from these networks allows the chemical apportionment
17      of the fine particle mass into aerosol types such as sulfates, nitrates,  organics, soot, and fine
18      soil (Schichtel and Husar,  1992;  Sisler et al., 1993, Sisler and Malm, 1994). The
19      quantification of these aerosol types is relevant to both the determination of the aerosol
20      effects as well as for source apportionment of fine particle species.  It should be emphasized
21      that the chemical composition as well as the absolute concentrations of the chemical species
22      is likely to be different in urban  areas and mountain  valleys, than at the remote monitoring
23      sites.  Also, the  quantification of organics, nitrates, and other metastable species is  subject to
24      major uncertainties.
25            The following discussion is a summary of the  national fine particle chemistry derived
26      from non-urban monitoring networks.  The national spatial pattern for fine particle  sulfate,
27      nitrate, organics and soot will be presented.  These aerosol types along with wind blown dust
28      account for virtually all aerosol fine  mass  in the conterminous US.  Presentation of the
29      detailed chemical pattern in urban-industrial areas as well as in poorly ventilated air sheds
30      over the mountainous western states  would be desirable but it is unavailable at this  time.
        April  1995                                6-21b      DRAFT-DO NOT QUOTE OR CITE

-------
 1            In the non-urban East the fine aerosol is dominated by sulfur aerosol types (sulfate and
 2      ammonium ions and associated water)  and organics which together constitute approximately
 3      80% of the fine paniculate mass.  Over the Northeast, organics dominate the fine particle
 4      mass, particularly during the winter season.  In the Southwest, fine soil is also a major
 5      component accounting for 25 to 30% of fine mass while sulfates are a less dominant
 6      component (as shown in Figure 6-12) (Schichtel and Husar, 1992).
 7            The national pattern of annual fine particle sulfate, nitrate, organics, and soot
 8      concentrations from the IMPROVE network is shown in Figure 6-13  (Sisler et al., 1993).
 9      The station density,  particularly over the eastern United States is limited.  The contour lines
10      in the annual  average maps are to be used as guides to the  eye, rather then actual values.
11      The eastern U.S. sulfate (Figure 6-13a) exceeds the concentrations over the mountainous
12      western states by factor of five or more.  Elevated  sulfate in excess of 1 /xg/m3 is also
13      reported over the Pacific coast states.  Sulfates also contribute over 50% of the eastern fine
14      particulate mass, while in the West sulfates  contribute 30% or below.
15            Fine particle nitrates (Figure 6-13b) are most prevalent over California, exceeding 4
16      Mg/m3 at most sites-  Their share of the fine mass at several California sites exceeds 20%.  It
17      is well known, however, that accurate measurement of nitrate concentration has been plagued
18      by numerous  sampling problems, and uncertainties  in chemical analysis.
19            Organic carbon concentrations are high over California, northwestern sites, as well as
20      at the eastern U.S. sites.  Relative to fine particle mass organics contribute over 50% over
21      the Northwest, and about 30% throughout the eastern US.  Sampling and chemical analysis
22      problems of organics are comparable to that of the  nitrates. For this  reason, the
23      concentration estimates of these meta-stable species are  continuously being revised.
24            Light absorbing elemental carbon/soot concentrations are high over the Northwest,
25      southern California, as well as at the Washington DC site.  In the Northwest, soot exceeds
26      10% of the fine mass concentration, but over most of the country it is 5% or less.
27            The chemical composition of PM10 and PM2 5 aerosols in the IMPROVE network
28      (Eldred et  a., 1994) revealed that the average coarse mass does not differ significantly
29      between the East and Wet, however, the fine mass  is higher in the East. Also about 80% of
30      soil elements  and 20 % of sulfur were found in the coarse  fraction.  Most trace elements
31      were found in the fine fraction, both in the  East and in  the West.  The spatial and seasonal

        April 1995                                6.24a       DRAFT-DO NOT QUOTE OR CITE

-------
                                                                                  <»
                                                                                 •a
                                                                                  o

                                                                                 •£
                                                                                 
-------
                                    o



                                    T3
                                    OS

                                    CD

                                    1/5


                                    O



                                    1
                                    •«->


                                    O)
                                    o



                                    I

                                    O)
                                    •a
                                    a
                                    O)
                                    ai


                                    £8
                                    0)
                                        /
                                         o
6-25      DRAFT-DO NOT QUOTE OR CITE

-------
                                                                         03
                                                                         ex
                                                                         o

                                                                         aT

                                                                         I*

                                                                         .2
                                                                         CA
                                                                         03
                                                                         I
                                                                         8
                                                                         tu
                                                                         o
                                                                         en
                                                                         en

                                                                         &
                                                                         >>
                                                                         ~
                                                                              en
                                                                              O\
                                                                              O\
April 1995
6-26      DRAFT-DO NOT QUOTE OR CITE

-------
  1      pattern in particle concentration and optical extinction in the United States from the
  2      IMPROVE network were also summarized by Malm et al., 1994.
  3            Studying the regional patterns of non-urban trace metals in the IMPROVE network
  4      (Eldred et al.,  1994) found a good correlation between selenium and sulfur at all sites in the
  5      East.  The correlation in the West is lower.  Comparison of the S/Se rations for summer and
  6      winter  shows that there is approximately twice the sulfur relative to selenium in summer
  7      compared to winter. Zinc is highest at the sites in the central East.  It does not correlate
  8      well with sulfur.   Lead and bromine are relatively uniform, with slightly higher mean
  9      concentrations  in the East. There is poor correlation between lead and bromine. Copper and
 10      arsenic are highest n Arizona copper smelter region.  Copper is also higher in the central
 11      East.
 12            Trends (1982 to 1992) of non-urban fine particle sulfur,  zinc,  lead, and soil  elements
 13      were reported by Eldred et al. (1994) using the IMPROVE network data.  They observe that
 14      in the southwest, sulfur trends in spring, summer, and fall decreased, while most of the
 15      winter trends increased.  The trends in the Northwest increase slightly.  The two eastern sites
 16      (Shenandoah and Great Smokey Mountains), have increased almost 4% per year in summer,
 17      increased 1 to 3 % in spring and fall, and  decreased 2% in winter.  The annual increase was
 18      between 2  to 3%. Generally, there were no significant trends in zinc and the soil elements.
 19      Lead at all sites decreased sharply through 1986, corresponding to the shift to unleaded
20      gasoline.  The ten year trends reported by  Eldred et al.  (1994)  have not been compared and
21      reconciled  with other compatible data.
22
23      6.3.1.6  Seasonality of the Non-urban Chemistry
24            This section discusses the seasonality of size segregated chemical composition at
25      non-urban monitoring sites (IMPROVE/NESCAUM) over the entire U.S. (Figure 6-14).
26            The nationally aggregated average PM10, PM2 5  and PMCoarse is shown in Figure
27      6-14b.  The non-urban PM10 concentration ranges from 8 /xg/m3 in the winter, December
28      through February to about 15 jug/m3 in June to August.  On the national scale the PM10
29      seasonality is clearly sinusoidal  with a peak.   Fine particles over the non-urban conterminous
30      United States account for about 50 to 60%  of the PM10  mass concentration throughout the
       April 1995                               6-27       DRAFT-DO NOT QUOTE OR CITE

-------
              PM2.5 - Nonurban PM Monitoring Sites
             03

             0.8

             0.7

             0.6

             0.5

             0.4

             0.3

             02

             0.1

             0.0
              1989 Feb Mar Apr May Jim Jul Aug Sep Od Hov Dec

          -A- Sulfate           -a Organics

          •+• Soil            ; -o- Soot	r

          -0- SulffOrfl+SoiKSoot
         PM10, PM2.5 and PMC Monthly Avg.
             U.S. - IMPROVE/NESCAUM Networks
          60,000,	.	.	.	.	.	.	,	.	,	.	,
                                                         55,000

                                                         50,000

                                                         45,000

                                                         40,000

                                                         35,000

                                                         30,000

                                                         26,000

                                                         20,000

                                                         15,000

                                                         10,000

                                                         5,000
                                                     -B- PM10
              1989 Fob Mar Apr May Jun Jul Aug Sep Oct Nov Dec


                    I -4- PM2.5
                                                                                  -A- PM Coarse
                                                                         US
          4,000


          3,500


          3,000


          2,500
       •5  2,000
       c
          1,500


          1,000


            500
             \j	
             1989 Feb Mar Apr May Jun Jul Aug  Sep OM  Nov Dec

        -A- Sulfur   -B- Selenium  -4- Vanadium I-o- S/Se 	!

   Scale   0^000       0-4        0-10     0^000
Figure 6-14.        Seasonal pattern of non-urban aerosol concentrations for the entire
                     U.S.  a) Average concentration,  b) PM10, PM2 5, and PMCoarse.  c)
                     Chemical fraction of sulfate, soil, organics, and soot,  d) Tracer
                     concentrations.
April 1995
6-28
DRAFT-DO NOT QUOTE OR CITE

-------
  1     year.  The coarse mass accounts for 40 to 50% throughout the year.  Hence, the fine-coarse
  2     aerosols ratio does not change dramatically for the average non-urban aerosol.
  3           The relative chemical composition of the aggregated non-urban aerosol pattern is
  4     shown in Figure 6-14c, including sulfates, organics, soil, and soot as a fraction of the fine
  5     particle mass concentration.  The Figure also shows the sum of these four aerosol species to
  6     indicate the fraction of the fine aerosol mass that is not accounted for.  Most notable among
  7     the missing species is the contribution of nitrates.
  8           Throughout the year,  sulfate aerosol, including the ammonium cation accounts for
  9     30 to  40% of the fine mass.  There is a remarkably mild seasonality in the nationally
 10     aggregated sulfate fraction. Organics also contribute 30 to 40%  of the nationally averaged
 11     fine particle mass. Thus, sulfates and organics are the two dominant species contributing to
 12     about 70% of the fine aerosol mass.
 13           The contribution of soil dust to the fine mass ranges between 4% in the winter months
 14     to 12% during April  through July.  Soot, i.e. elemental carbon is about 2% during the
 15     summer and 5% during the winter.
 16           The sum of the four measured fine mass components,  sulfates, organics, soil and soot
 17     add up to about 80% of the measured fine mass throughout the year.  The remaining,
 18     unaccounted fine mass may be contributed by nitrates, trace metals (e.g. Pb, Br),  sea salt
 19     (NaCl), etc.
 20           The seasonal pattern of concentration of primary emission tracers, selenium, Se and
 21      vanadium, V is shown Figure 6-14d.  Se  is a known tracer for coal combustion (Miller and
 22     Friedlander,  1992?),  while V is a trace constituent of fuel oil.  The Figure also shows the
 23      monthly average concentration of fine particle sulfur as well as the S/Se ratio.  If all the fine
 24      particle sulfur was contributed  by coal combustion then S/Se  ratio would be a measure of
 25      secondary  sulfate formation.
 26            The national average Se concentration is rather uniform over the  seasons, ranging
 27      between 400 to 600 ng/m3.   Since Se is a primary pollutant, the seasonal invariance means
 28      that the combined effect of emissions and dilution is seasonally invariant over the year.
29            The concentration of V is between 500 to 700 ng/m3, with the higher concentrations
30      occurring in the winter season.  Evidently, emission of V bearing fuel oil is more
31      pronounced during the cold season.  The monthly average sulfur aerosol exhibits the highest

        April 1995                                6_29       DRAFT-DO NOT QUOTE OR CITE

-------
 1     concentrations 1.5 /^g/m3, during June, July, and August, and the lowest values 0.9 /-ig/m3,
 2     during November, December, and January.
 3           The S/Se ratio is about 700 during November to January and climbs to about 1,500
 4     during April through September. The higher S/Se ratio during the warm season is an
 5     indication of secondary sulfate production.   Alternatively, the summer sulfates could be
 6     contributed by sources other than coal combustion.
 7
 8     Eastern United States
 9           The seasonal pattern of the eastern U.S. aerosol chemistry is shown in Figure 6-15.
10           The concentration of PM10, PM2 5, PMCoarse (Figure 6-15b) indicates a similar
11     seasonality, highest concentrations in the summer, and lowest in the winter.  The PM10 levels
12     range between 12 to 24 ^g/m3, the PM2 5 are between 8 to 12 /ig/m3, while PMCoarse are 4
13     to 7 /zg/m3. The size segregated aerosol data for the non-urban East  show that the fine mass
14     concentration (8  to 12 /xg/m3) is higher than the  national average (4 to 8 ^g/m3),  while the
15     coarse mass concentration is comparable to the national average.  Consequently, eastern U.S.
16     non-urban  fine particles contribute 60 to 70% of the fine mass throughout the year.
17           The apportionment of the fine particle mass into its chemical components
18     (Figure 6-15c) favors sulfates which amount to 40 to 50%  of the fine mass throughout the
19     year, compared to about 30% of organics.  The  contribution of soil dust is about  5%
20     throughout the year, while soot is more important in the winter (6%)  than in the summer
21     (3%).  The above three aerosol chemical components account for  85 to 90%  of the measured
22     fine particle mass, leaving only marginal contribution to nitrates, trace metals, and sea salt.
23           The coal tracer selenium (Figure 6-15d) exhibits a modest winter peaked seasonality
24     between 600 to 800 ng/m3.  Vanadium on the other hand, is factor of two higher in the
25     winter (1,500 ng/m3) compared to the summer (750 ng/m3). Evidently, the primary
26     contribution from fuel oil is winter peaked.  The  S/Se ratio is about 1,000 in the winter, and
27     it is over 2,000 in the summer months. This suggests the seasonality of secondary sulfate
28     formation during the summer months, but other  factors can not be excluded.
       April 1995                               6-30      DRAFT-DO NOT QUOTE OR CITE

-------
            PM2.S - Nonurban PM Monitoring Sites
        PM10, PM2.5 and PMC Monthly Avg.
        East of Rockies - IMPROVEyNESCAUM NetworKs
         60,000,	,	,	^	,	,		
                                                   55,000

                                                   50,000

                                                   45,000

                                                   40,000

                                                   36,000

                                                   30,000

                                                   25,000

                                                   20,000

                                                   15,000

                                                   10,000

                                                   5,000
                                 &--A.
                                                                                 tt--£r--'
                                               i-e- PM10
           1989 Fob Mar Apr May Jun Jul Aug Sep Oct Nov Dec

           	   ! -4- PM2.5       -A- PMCoQrse
                    Eastern US
                                                                Eastern US
03

0.8
0.7
0.6
0.5
0.4
0.3
02
0.1
19!
-&- Sulfi
4 Soil
— Sulf
,
- -;---,••''' ~~~' •*<.... fr. .-—-'
-


- »s*^-+ s^*^
^^ \/ ^s^_.
^~~~ -Q. /* '~^' — ^
-
.:.5^-*"-*^L-i:*,-*-^.-..t.-S--:
9 Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
ite -B- Organics
-o- Soot
•Org+SoiH-Soot
4,000
3,600

3,000
2,600
ro
c
1,500
1,000
500
0
19t
-A- Surfu
Scale 0-400(

-

-
-
c
''ft
A 4. _
t '""*"• S^ ~~~*^ *-^ * •'
. ^^^D-^B-^*-*
9 Fob MM Apt May Jun Jul Auj Sep Oe> Hoy Dec
-a- Selenium -+- Vanadium I -o- S/Se ~1
3 0-4 0-10 0-4000
Figure 6-15.       Seasonal pattern of non-urban aerosol concentrations for the Eastern
                   U.S.  a) Average concentration,  b) PM10, PM2 5, and PMCoarse.  c)
                   Chemical fraction  of sulfate, soil, organics, and soot,  d) Tracer
                   concentrations.
April 1995
6-31
DRAFT-DO NOT QUOTE OR CITE

-------
 1      Western United States
 2            The aggregated western U.S. aerosol seasonality is presented in Figure 6-16. The
 3      non-urban aerosol concentrations for PM10, PM2 5, and PMCoarse are well below the
 4      concentrations over the  eastern United States (Figure 6-16b).  Evidently, the non-urban
 5      western United States differs from the eastern United States, having lower fine mass
 6      concentrations, which range between 3 to 5 ^ig/m3.  The  coarse mass concentration (4 to
 7      8 jug/m3) and seasonality is similar over the East and the  West.  It is worth emphasizing,
 8      however, that these measurements are at remote national  parks and wilderness areas in both
 9      East and West. The examination of monitoring data in urban areas and confined airsheds
10      (Sections 6.4 and 6.5) reveals a highly textured pattern in space and time.
11            The fine particle  chemical mass balance (Figure 6-16c) for the aggregated western
12      United States shows the dominance  of organics, which account for 30 to 45% of the fine
13      mass.  The higher organic fraction occurs in the November through January season.  Sulfates
14      hover  at 20 to 25% throughout the year.   Soil dust plays  a prominent role  in the western fine
15      mass balance, contributing 20% in April through May, but declining to 5% by January.  Soot
16      ranges between 5% in the winter and  2 % during the summer.  About 25% of the fine mass
17      over the western United States is not accounted for by sulfates,  organics, soil , and soot. It
18      is known (Cahill???) that nitrates are major contributors to the fine particle mass in the South
19      Coast  Basin, as well as  other western regions.
20            The concentration of the trace substances (Figure 6-16d) selenium and vanadium
21      shows both low concentrations and weak seasonality.  The sulfur concentrations are also less
22      than half of the eastern  U.S.  values.  The S/Se ratio is about 500 in the winter months and
23      1,000  during the summer.  The  low summer S/Se values  would indicate less sulfate yield per
24      selenium  in the summer.  However, these observations need to be tempered by the fact that
25      selenium  emitting coal-fired power plants are not the only sources of western U.S. sulfur.
26      The S/Se ratio is included here for sake of completeness.
27            The above discussion of national pattern of chemical and size dependence  hide a rich
28      spatial and temporal texture of the U.S. aerosol pattern discussed in the following sections.
29      However, it provides  the national scale gross features and serves as a broader context for the
30      more detailed examinations.
        April 1995                                6-32      DRAFT-DO NOT QUOTE OR CITE

-------
                 PM2.5 • Nonurban PM Monitoring Sites
            PM10, PM2.5 and PMC Monthly Avg.
            West or Rockies - IMPROVE/NESCAUM Networks
                                                          55,000

                                                          50,000

                                                          45,000

                                                          40,000

                                                          35,000
                                                       c
                                                       §  30,000
                                                       c
                                                          26,000

                                                          20,000 -

                                                          15,000

                                                          10,000

                                                           5,000
                                                             1989 Feb Mar Apr May Jun Jul  Aug Sep OO Nov Dec


                                                                   j -4- PM2.5        -A- PM Course
                         Western US
                          Western US

03

0.8

0.7
0.6
05


0.4

0.3
02
0.1
n n
1
_

"- _.-,--~
..-•'"'""" --., ,- --• ~



	 t
i .
-\ 0"
"s, ^ c- -a
'^^^^^^--^-V
r^~^rtr^~^ ^%^-^
" ,-•'' + "" '""•*•-
>•- lj o. - -o- -*---
4.UUU

3,500

3,000

2,500
„
I 2,000
c

1,500

1,000
500
D.O 	
1989 fell Mar Anr May Jun Jul Aug Sen Oct Nov Dec n


-

-

-






o - '"'- '•* ,
l^:^^^^^^^

             -+  Soil            | -o- Soot

             -<%  Sufft-Org+Soil+Soot
                1989 Feb MM Aft Uay Jun Jul  Aug Sep On  Nov Dec

            -A- Sulfur   -g- Selenium -4- Vanadium }-e- S/Se	;

       Scale    0-4000      0-4        0-10      0-4000
Figure 6-16.       Seasonal pattern of non-urban aerosol concentrations for the Western
                    U.S.  a) Average concentration,  b) PM10, PM2-5, and PMCoarse.  c)
                    Chemical fraction  of sulfate, soil, organics, and soot, d) Tracer
                    concentrations.
April 1995
6-33
DRAFT-DO NOT QUOTE OR CITE

-------
 1     6.3.2  Urban National Aerosol Pattern - AIRS
 2           The urban monitoring network is operated by state and local agencies as mandated by
 3     the Clean Air Act.  The data from this network constitute the sensory input for the detection
 4     of exceedences over the paniculate matter standards.  Federal regulations also require that
 5     these monitoring data be submitted to the EPA Aerometric Information and Retrieval System
 6     (AIRS).  In what follows,  AIRS PM10 refers to the PM10 mass concentration extracted from
 7     the AIRS database.
 8           The AIRS PM10 stations are mostly in urban areas but some suburban and non-urban
 9     sites are also reported.  The aerosol mass concentrations for PM10 and PM2 5  are sampled and
10     weighed by a variety of devices.   The sampling frequency is generally every  sixth day for
11     24-hours.  The analysis presented in this section is based on PM10 and PM2 5 data retrieved
12     from AIRS in October  1994.
13           The PM10 station density has been increasing over time.  Particulate matter sampling
14     with the size cut-off of 10/xm,  PM10 begun in the early 1980s.  By 1985 about 200 samplers
15     were operational and the number of sampling  stations has grown to 1,350 by 1994 as shown
16     in Figure 6-17.  The emergence of new stations  appeared in rough proportion to the final
17     station density shown in Figure 6-17. In other words, in 1985, the national coverage had a
18     similar pattern to 1994, except less dense.
19           The results of AIRS PM10 aerosol pattern analysis are presented in quarterly contour
20     maps, as well as seasonal time charts. For valid monthly and quarterly aggregation, it was
21     required to have at least two samples a month, and six samples per quarter.  For the seasonal
22     maps all the available data between 1985 to  1994 were used.
23           The seasonal contour maps also show the  location of the PM10 monitoring sites.  The
24     size of the rectangle at each site is proportional to the quarterly average PM10 concentration
25     using all available data between 1985 to  1994. Hence, sampling biases due to station density
26     that changed over time can not be excluded.
27           The quarterly concentration pattern of PM10 is  shown in Figure 6-18.  The high
28     sampler density  allows  the resolution of spatial texture on the scale of 100 km, particularly
29     over major metropolitan areas.  However, remote regions in the  central and western states
30     have poor  spatial density.  In the absence of rural monitoring data computerized contour
       April 1995                               6-34      DRAFT-DO NOT QUOTE OR CITE

-------
o\
O
O
2

3
O
c
O

3
H
W
                                                            i
                                                                         Valid PM10 Stations

                                                                             US, All Stations
    Figure 6-17.
                                                1984    1986



Trend of valid PM10 monitoring stations in the AIRS database.
1988
1990
1992

-------
OS
H

b
o
O
G
O
H
W


i

n

                          ,,rw^u
                                                                          PM10 Mass
Figure 6-18.      AIRS PM10 quarterly concentration maps using all available data.

-------
 1      plotting of PM10 is biased toward extrapolating (spreading) high concentrations over large
 2      areas. This bias is particularly evident in the maps for Quarters 1 and 4 in the western
 3      states, where the area of high concentration hot spots is exaggerated.
 4            The AIRS PM10 concentrations over the eastern United States are lowest during
 5      Quarter 1, ranging between 20 to 30 /xg/m3.   The higher concentrations exceeding 30 /ng/m3
 6      are confined to metropolitan areas.
 7
 8      6.3.2.1 National  Pattern and Trend of AIRS PM10
 9            The trend (1985 to 1993) of national average PM10 concentrations is shown in
10      Figure 6-19b.  During the decade there was a remarkable reduction in PM10 concentrations
11      from 48/ig/m3 to  25 ptg/m3.  Between 1986  and 1993, the reduction was 38%.  The
12      Figure 6-19b also  shows the standard deviation among the yearly average PM10
13      concentrations for each year.   On the national scale the standard deviation of yearly average
14      concentrations is about 40% of the mean.
15            The concentration of PM2 5 and PM10 are compared in the scatter chart in
16      Figure 6-19c.  Each point represents a pair of PM2 5-PM10 monthly average concentrations.
17      The diagonal line is the 1:1 line and shows the fine particle concentration ranges between
18      20 and 85% of PM10.  The heavy solid line is derived from linear best fit regression. The
19      detailed correlation statistics is reproduced in the upper-left corner of the scatter charts.
20      The ratio  of overall average PM2 5 and overall average PM10 is also indicated.  For the data
21      when both PM2 5 and PM10 data were available,  nationally aggregated PM2 5 particles
22      accounted for 57% of the PM10 mass.
23            The seasonal pattern of the national PM10 concentration  is also depicted in
24      Figure 6-19d, utilizing all available  data in AIRS. The national average PM10 seasonality
25      ranges between 27 jiig/m3 in March  and April, and 33 /xg/m3 in July and August, yielding a
26      modest 16% seasonal  modulation.  There is also evidence of slight bimodality with the
27      December through January peak.
28            The seasonal chart also shows the annual variation of PM2 5, and PM10-PM2 5
29      (i.e. coarse particles).  The national fine particle concentration  shows clear evidence  of
30      bimodality with peaks in July and December. It is shown below that the fine particle winter
31      peak arises from western sites,  while the  summer peak is due to eastern U.S. contributions.

        April 1995                               6-37      DRAFT-DO NOT QUOTE OR CITE

-------
               Average AIRS PM10 -1985-1993
     AIRS PM2.5 vs. PM10 - Monthly Avgerages
                 Conterminous U.S.
          Avg X
          Avg Y
          Avg Y/Avg X
          Corr Coeff
          Slops
          Y offset
          Data Points
         0       30      60      90     120

              PM10JVVG UWCU METER (25 C)
        AIRS PM10 Concentration Trends
                        U.S.
     u
     U>
                                                      60

                                                      65

                                                      50

                                                      45

                                                      40

                                                      35

                                                      30

                                                      26

                                                      20

                                                      15

                                                      10

                                                       6
                                                      1985  1986 1987  1988  1989 1990  1991 1992


                                                               I -Q- PM10AVC-SIG  -+ PM10AVC + SIG
                                                     PM10 AVG
     AIRS PM10, PM2.5 AND PM2.5-10  MONTHLY CONC,
                  Conterminous U.S.
        60
                                                  3
                                                  5
        55 -

        50 -

        45


        «°

        35

        30

        25

        20

        16

        10

         5
         198S    Mar Apr May J«n Jul Aug S«p Oct Nov D«c


     -a- PM10       -+- PMZ6       | -A- PM2.6-10   ~|
Figure 6-19.       AERS PM10 and PM2 5 concentration pattern for the conterminous
                    US.
April 1995
6-38
DRAFT-DO NOT QUOTE OR CITE

-------
 1     The national average coarse particle concentration has a 50 % yearly modulation with a
 2     single peak in July.
 3           Summarizing the national PM10 concentrations one can observe that the country has
 4     several major  aerosol regions.  Each region has a discernible geographic extent as well as
 5     seasonal  pattern.  Over the plains of the eastern United States the spatial texture of PM10 is
 6     driven by the  pattern of the emission fields, while the seasonality of concentrations is likely
 7     to be determined by the chemical transformation and removal processes,  as well as by the
 8     regional  dilution.  In the mountainous western and Pacific states US, pockets of wintertime
 9     PM10 concentrations exist that well exceed the eastern U.S. values. It is believed that haze
10     and smoke in  confined mountain valleys and air basins are strongly influenced by topography
11     which in turn  influences the emission pattern, dilution, as well as the chemical transformation
12     and removal rate processes.
13           Given the regionality of the aerosol concentration pattern  much of the discussion that
14     follows will be focused on the characteristics  of these aerosol regions.  The Rocky Mountains
15     produce  a natural division between the eastern and western aerosol regimes which will be
16     discussed next.
17
18     6.3.2.2  Eastern U.S. PM10 Pattern and Trend
19           The eastern U.S. PM10 concentration (Figure 6-20b) shows a 29% downward  trend of
20     yearly average PM10 concentrations from 35 /zg/m3 in 1985 to 25 /wg/m3 in 1993.  The
21     decline is rather steady over time.
22           The highest eastern U.S. AIRS PM10 concentrations are recorded in Quarter 3 (Figure
23     6-20d).  The peak concentrations are over the Ohio River Valley stretching from Pittsburgh
24     to West Virginia, southern Indiana and St. Louis. In this region, the PM10 concentration
25     over the  industrialized Midwest exceeds 40 pig/m3.  Additional hot-spots with > 40 /ug/m3
26     are recorded in Birmingham, AL, Atlanta, GA,  Nashville, TN, Philadelphia, PA and
27     Chicago. IL.  The summer time PM10 concentrations in New England and upstate Michigan
28     are  < 20 jiig/m3.
29           The transition seasons  Quarters 2 and 4 (Figure 6-20d) show about 30 jwg/m3 over
30     much of the eastern US, with concentration hot-spots over the industrial Midwest as  well as
31     in the Southeast, Atlanta, GA and Birmingham,  AL.  It is quite remarkable, however,

       April 1995                               6-39      DRAFT-DO NOT QUOTE OR CITE

-------
             Average AIRS PM10 -1985-1993
                                                    AIRS PM10 Concentration Trends
                                                               East of Rockies
   AIRS PM2.5 vs. PWI10 - Monthly Avgerages
                East of Rockies

    160rCOkREt.ATiON' SfATi:
        Corr Coeff
    120|-siapB
        Y offset
    1101-Data Points
       1985  1986 1987  1988 1989 1990  1991 1992

   -A- PM10AVG     -a PM10AVO-SIO   +PM10AVO + SIG


   AIRS PM10, PM2.5 AND PM2.5-10 MONTHLY CONC.

                 East of Rockies
     60

     65

     50

     45

     40

     35

     .30

     25 k_ -e-

     20

     15

     10

      6
                                                                            \
              30      60      90     120

           PM10_AVG UGJCU METER (26 C)
      1385   M»r Apr May Jun Jul  Aug B*p Oct Nov D*c


  -B  PM10        -+ PM2.6      \-tr- PM2.8-10~   |
Figure 6-20.        AIRS PM10 and PM2 5 concentration pattern for east of the Rockies.
April 1995
6-40
DRAFT-DO NOT QUOTE OR CITE

-------
 1      that the PM10 concentrations in urban-industrial "hot-spots" exceed their rural surrounding
 2      by less than a factor of two.
 3            The spatial variability of PM10 over the eastern United States is driven primarily by
 4      the varying aerosol emission density.  This can be deduced from the coincidence of high
 5      concentrations within urban industrial areas.  The atmospheric dilution, i.e. horizontal and
 6      vertical dispersion is not likely to be spatially variable.  Also,  the chemical aerosol formation
 7      and removal processes are likely to have weak spatial gradients when averaged over a
 8      calendrical quarter.  Hence, the main factor that is believed to be responsible for the spatial
 9      variability is the emission field of primary PM10 particles and the precursors of secondary
10      aerosols.
11            PM10 concentration in excess of 30 /*g/m3 is recorded over the agricultural states of
12      Iowa, Kansas, Nebraska,  and South Dakota.  The elevated PM10 concentrations over this
13      region tend to persist over all four seasons.  The spread of yearly average concentrations east
14      of the Rockies is only 28%.
15            The eastern PM10 seasonality (Figure 6-20d) is rather pronounced, with winter
16      concentrations (December through March) of 24 /^g/m3, and July through August peak of 35
17      Mg/m3- The amplitude of the PM10 seasonal concentrations is  about 30%.
18            The scatter chart of PM2 5-PM10 relationship shows significant amount of scatter,  with
19      a slope of 0.58.  The ratio of the overall average PM2 5 and PM10 concentration is 0.6 such
20      that 60%  of PM10 in the sub 2.5 /im size range.  The seasonality of the fine particle
21      concentration  over the East is bimodal with a major peak in July and smaller winter peak in
22      January.  The coarse particle concentration shows a single broad peak over the warm season,
23      April through October (Figure 6-20d). It is therefore evident that fine and coarse particles
24      have different seasonal dynamics in the East.
25
26      6.3.2.3 Western U.S. PM10 Pattern and Trend
27            The mountainous states, west of the  Rockies (Figure 6-21) show high PM10
28      concentrations (>50 jug/m3) at localized hot-spots during the cold season, Quarters 1 and 4.
29      These high concentrations occur over both  metropolitan areas such as Salt Lake City, as well
       April 1995                                6-41      DRAFT-DO NOT QUOTE OR CITE

-------
          Average AIRS PM10 -1985-1993
    AIRS PM10 Concentration Trends
               West of Rockies
   AIRS PM2.5 vs. PM10 - Monthly Avgerages
               West of Rockies
    150
      1985 1986 1987  1988 1989 1990  1991 1992

               -O PM10AVG-SIO  -4- PM10AVG + SIG


AIRS PM10, PM2.5 AND PM2.5-10 MONTHLY CONC.
                West of Rockies
     60
                                                        Mar Apr May Jun Jul  Aug S«p Oet N«v Dec

                                                            4- PM2.5
                                              -B- PM10
            PM10_AVG UGTCU METER (25 C)

Figure 6-21.       AIRS PM10 and PM2.5 concentration pattern for west of the Rockies.
 April 1995
6_42      DRAFT-DO NOT QUOTE OR CITE

-------
  1      as in smaller towns in mountain valleys of Colorado Wyoming, Montana, Idaho, Oregon,
  2      and Washington.
  3            The main geographic features of California PM10 aerosols are the Los Angeles basin
  4      and the San Joaquin Valley.  Both basins show concentrations in excess of 50 /^g/m3.  These
  5      basins are also confined by surrounding mountains that limit the dilution, facilitate cloud
  6      formation, and have emissions that are confined to the basin floor.  Accordingly, they
  7      represent airsheds with characteristic spatial and temporal pattern.  It is likely that the actual
  8      spatial texture of the PM10 concentration field in the mountainous western states has much
  9      higher texture than depicted in Figure 6-2la.
 10            It appears that the spatial pattern of these high concentration hot spots is driven by
 11      both emissions as well as by the restricted wintertime ventilation due to mountainous terrain.
 12      Over the mountainous western states the atmospheric dilution by horizontal  and vertical
 13      dispersion is severely restricted by mountain barriers and atmospheric stratification due to
 14      strong and shallow inversions.  Radiative cooling also causes fog formation which enhances
 15      the production rate of aerosols in the valleys.  As a consequence, mountain tops are generally
 16      extruding out of haze layers.  Emissions arising from industrial, residential, agricultural,
 17      unpaved roadways and other sources are generally confined to mountain valleys.  In the
 18      wintertime the mountain valleys are frequently filled with fog which also influences the
 19      chemical transformation and removal processes.  As a consequence all  three major factors
20      that determine the ambient concentrations, i.e. emissions, dilution, and chemical rate
21      processes are strongly influenced by the topography.  For this reason, many of the maps
22      depicting the regional pattern use shaded topography as a backdrop.
23            The western half of the US,  west of and including the Rockies, show a more
24      pronounced downward PM10 concentration trend (Figure 6-2 Ib).  The reduction between
25      1985 (57 /ig/m3) and 1993 (26 /ug/m3) is a  remarkable 55%.  The reduction between 1986
26      (42 jtig/m3) and 1993 (26 jug/m3) is 38%.  Standard deviation among the western stations of
27      yearly average PM10 concentrations is about 40%.
28            The PM2 5-PM10 relationship (Figure 6-21c) shows that on the average about 50% of
29      the PM10 is contributed by fine particles.  The scatter chart (Figure 6-2Ic) also shows that
30      during high concentration PM10  episodes the fine fraction dominates.
         April 1995                                6-43      DRAFT-DO NOT QUOTE OR CITE

-------
  1            The western PM10 seasonality (Figure 6-2Id) is also rather pronounced, having about
  2      30% amplitude. However, the lowest concentrations (26 /ig/m3) are reported in the late
  3      spring (April through June), while the highest values occur in late fall (October through
  4      January).
  5            The seasonality of PM2 5 west of the Rockies (Figure 6-21d) is strongly peaked in
  6      November through January (40 ^ug/m3). In fact, it is about factor of four higher than the
  7      summertime values.  On the other hand, the coarse fraction shows a broad peak during late
  8      summer, July through October.  It is to be noted that in Figures 6-20 and 6-21, the fine and
  9      coarse particle concentrations do not add up to PM 10, because size resolved samples were
10      only available for tens of sites, while the PM10 concentrations were obtained from hundreds
11      of monitoring stations.
12            In summary, there is a remarkable 40 to 50% reduction of national PM10
13      concentrations between 1985 and 1993.  On the national average the PM10 seasonality is
14      insignificant.  Desegregation of the national averages into east and west of the Rockies,
15      shows that the downward trend west of the  Rockies is more pronounced than over the eastern
16      half of the US.  The east west desegregation also shows that the lack of national PM10
17      seasonality arises from two strong seasonal  signals  that are phase shifted, the eastern United
18      States  has a summer peak, the West fall and winter peak,  and the sum of two signals is a
19      weakly modulated seasonal pattern. Nationally, PM2 5 mass accounts for about 57% of PM10
20      mass.  The East and West show comparable fine fraction (60% in the East and 50% in the
21      West), and fine particles tend to dominate during the winter season particularly in the
22      western US.
23            It is evident that further examination discussed in the next sections will show that the
24      East-West division itself is rather crude and that dividing the conterminous United States into
25      additional subregions is beneficial in explaining the PM10 concentration pattern and trends.
26            A cautionary note on a possible sampling bias is in order.  The national average
27      concentrations were calculated utilizing all of the available data since 1985, when more than
28      200 monitoring stations were operational.  Since that time, the number of monitoring stations
29      has risen to more than 1,300.  The implications of the changing stations density to the above
30      described national PM10 trend is not well studied.  Also, changes in sampling equipment and
         April 1995                               6-44       DRAFT-DO NOT QUOTE OR CITE

-------
  1      monitoring protocols are also possible causes of systematic errors in the reported spatial
  2      pattern and trends.
  3
  4      6.3.2.4  Short-term Variability of PM10 Concentrations
  5            The previous aerosol concentration patterns were expressed as quarterly averages.
  6      However, for health and other effects, the variance of the concentration, in particular the
  7      occurrence of extreme high concentrations is of importance.  The PM10 concentrations
  8      exhibit marked differences in the shape of their distribution functions around the mean
  9      values.  For example in Figure 6-22, the day to day variations of PM10 concentrations in
 10      Knoxville, TN are about 40% of the mean value of 35 jig/m3.  On the other hand, the
 11      concentration time series for Missoula, MT  shows a coefficient of variation  of 60% over the
 12      mean of 34/ig/m3.  During the winter season the coefficient of variation is even higher.  It
 13      is therefore evident, that for comparable mean  concentrations the Missoula,  MT  site exhibits
 14      significantly higher short-term variations.
 15            The variability of concentration is examined spatially and seasonally by computing
 16      logarithmic standard deviation (ratio of 84/50 concentration percentiles) for each monitoring
 17      site.  These deviations were then contoured  for each season.  The results are depicted in the
 18      seasonal maps of the logarithmic standard deviation (Figure 6-23).  The highest
 19      logarithmicstandard deviation is recorded over  the northern and northwestern states during
20      the cold season, Quarters  1 and 4.  Regionally, the logarithmic standard deviation in the
21      north-northwest is about 2.0 with pockets of high winter variability such as Salt Lake City,
22      UT and Missoula, MT. The lowest variability  prevails over the warm season, Quarters
23      2 and 3, covering the southeastern and southwestern states.  Over multistate regions in the
24      southern states the summertime logarithmic  standard deviation is below 1.5.  This means that
25      these areas are covered more or less uniformly by summertime PM10, while the northern
26      states are more episodic.
27
28      6.3.2.5  AIRS PM2 5 Concentrations.
29           The mass concentration of fine particles  in urban areas is not well known.  Sampling
30      and analysis of PM2 5 is limited by small number of stations (<50), sampling period
31      restricted to few years,  and different, non-standard sampling equipment was utilized for

         April 1995                               6-45      DRAFT-DO NOT QUOTE OR CITE

-------
            o
          o in
         Iln
         °- S
           I
                    Mean  :  34
                    CoVa  :  60.64
                    -Min   :  1  !
                    Max   :  239
                    Points:  1660
                  1988       1989      1990      1991

                             —-PM10 300630031 MISSOULA
                                         1992
                                                   1993
200


180


160


140


120


100


80


60


40


20
                    STATISTICS:

                    Mean   : 35
                    CoVa   : 39.92
                   -Min   : 9
                    Max   : 73
                    Points: 258
                 1988       1989       1990       1991

                            —=PM10 470931015 KNOXVILLE
                                        1992
                                                  1993
Figure 6-22.       Short-term PM10 concentration time series for Missoula, MT, adn
                   Knoxville, TN.
April 1995
                           6-46
DRAFT-DO NOT QUOTE OR CITE

-------
                                                                                                 ,  Quarter 2
VO
0\
6
o
2
O

o
c;
o
                                                                              PM10 Log Standard Deviation
PM10 Log Standard Deviation -source IMPROVE and NESCAU




                                 3
                                                                                               ^ -v -.  Quarter 4
•lav:.:-".-.
                       PM10 Log Standard Deviation
                                                 ' SJUICB IMPROVE and NESCAU
                                                       PM10 Log Standard Deviation
n
    Figure 6-23.      Logarithmic standard deviation AIRS PM10 concentrations.

-------
 1     PM2 5 monitoring.  Consequently, it is not possible to perform a detailed spatial and
 2     temporal mapping and trend analysis for AIRS PM2 5.
 3           The yearly average AIRS PM2 5 concentrations are shown in Figure 6-24.  Figure
 4     6-24 also shows the location and magnitude of PM2 5 concentrations arising from the
 5     IMPROVE/NESCAUM monitoring networks.  The fine particle data from the
 6     IMPROVE/NESCAUM shows a smooth pattern with uniformly high concentrations
 7     (> 15 /xg/m3) occurring over the eastern United States and uniformly low concentrations of
 8     <5 jug/m3 between Sierra and Cascade Mountains. This pattern of non-urban fine particle
 9     concentrations was discussed in Section 6.3.1.
10
11     6.3.2.6 Other National Surveys
12           A summary of urban PM10, PM2 5, PMCoarse at eight urban areas, Birmingham, AL,
13     Buffalo, NY, Houston, TX, Philadelphia, PA, Phoenix, AZ, Pittsburgh, PA, Rubidoux, CA
14     and Steubenville, OH was reported by Rhodes and Evans (1985).  The overall ratio of the 10
15     jum fraction to Total Suspended Paniculate (TSP) was 0.486.  The relationships between
16     PM10 and the 15 /on fraction (IP) are very linear for all sites. With exception of Phoenix,
17     AZ and Houston, TX, PM2 5 exceeded the PMCoarse mass concentration in all six urban
18     areas.
19           Spengler and Thurston (1983) reported inhalable paniculate matter (IP) concentrations
20     in six United State cities, Portage, WI Topeka, KS, Kingston, TN, Watertown, MA St.
21     Louis, MO and Steubenville, OH using dichotomous  virtual impactors in the two size ranges,
22     PM2 5  , having dp<2.5 /xm and PMCoarse with 2.5 
-------
  c
  o
  w

  •g
  0.


  o
  o
  ft
 I

o
                                                                 g
                                                                 a:
                                                                 a.
                                                                         I
                                                                        T3


                                                                         CS
                                                                       U
O

85
                                                                         O

                                                                        ^

                                                                        •a
                                                                         
-------
      30
      20
   1
      10
          • IP mats
          • Fine mats
          • CoirM man
          • Total aulfaw man
                        Portage, Wl
                    .*"•>
                                                                   Topeki, KN
                         i   i   «,
                         I   /   •
                     *-\  I  I

                     •  > "
               yVvfxVH
       J FMAMJ JASONOJ FMAMJJASONOJ FMAMJJASONO

            1979           1980           1901
                                                                             • IP mats
                                                                             * Fine mass
                                                                             * Coarse man
                                                                             * Total sulfate mas*
                                                   JFMAMJJASONDJFMAMJJASONDJ FMAMJ J ASOND

                                                        1970          1900           1981
      60
      60
      3°
      20
      10
                        Harriman, TN
                             • IP mass
                             * Fine mass
                             * Coarse mass
                             • Total tulfaie mass
       J FMAMJ  JASONDJ  FMAMJ J ASONO

                1980                  1981
                                                       60
                                                                       Walertown, MA
                                                       30
                                                       20
                                                       10
      • IP mass
      * Fine mass
      • Coarse mass
      * Total sulfatc mass
                                                         .ANv\"'X/
                                                   JFMAMJJASONOJFMAMJJASONDJFMAMJJASONO

                                                        1979          1980           1981
70



60



50







30



20



10
                       Si. Louis, MO
         • IP mass
         « Finf mass
         • Coarse mass
         * Total sulfato mass
                                                                       Steutxjnvillr. OH
   80


   70


   60



1  5°

*  40


   30


   20


   10
• IP mass
* F me mass
» Coarse mass
* Totat sulfalc mass
       J FMAMJJASONOJ FMAMJJASONOJ FMAMJ JASONO

            1978           1980           1981
                                                   J FMAMJJASONDJ FMAMJ JASONDJ FMAMJJASOND
                                                         1979           1980          1981
Figure 6-25.        Monthly mean concentrations in (/tg/m3) of IP, Fine Fraction, and S
                     as (NH4)2SO4 in Portage, WI; Topeka, KS; Harriman, TN;
                     Watertown, MA; St. Louis, NO; and Steubenville, OH.
Source:  Spengler and Thurston, 1983


April 1995
                                          6-50        DRAFT-DO NOT QUOTE OR CITE

-------
  1      sites, an average of only 4.9 weight per cent of the coarse particle mass that was measured
  2      by the dichotomous samplers consists of quartz and 0.4 weight per cent as quartz in fine
  3      mass.  Continental interior sites show the highest average quartz content as well as the
  4      greatest variability.  The coastal regions and eastern interior sites reveal the lowest quartz
  5      concentrations. The complete X-ray spectra from some samples in Portland, OR,  show that
  6      Si comes primarily from minerals such as feldspars, where the Si in the Buffalo, NY
  7      aerosols  comes from quartz.
  8
  9      6.3.3  Comparison of Urban and Non-Urban Concentrations
 10            Seasonal maps of the AIRS PM10-IMPROVE/NESCAUM PM10 spatial concentrations
 11      are given in Figure  6-26.  The AIRS PM2 5 concentrations everywhere exceed their adjacent
 12      IMPROVE/NESCAUM concentrations. The highest AIRS PM2 5 are  reported over the
 13      eastern urban industrial centers, such as Philadelphia and Pittsburgh, where the
 14      concentrations of 50 /xg/m3 exceed their non-urban counterparts by a factor of 2 to 3.
 15      However, the excess urban PM2 5 concentrations are evidently confined to the immediate
 16      vicinity of urban centers.  The PM2 5 concentrations at remote New England, over the
 17      southeastern US, and over the upper Midwest are within about 50% for AIRS PM2 5 and
 18      IMPROVE/NESCAUM PM2 5.  This indicates that over the eastern United States a
 19      regionally homogeneous background of PM2 5 concentration exists that has smooth spatial
20      gradients. Superimposed on the smooth regional pattern are local hot-spots with excess
21      concentrations of factor of 2 to 3 that are confined  to few miles of urban  industrial centers.
22      The regional homogeneity is an indication that the eastern U.S. PM2 5 is composed of
23      secondary aerosols that is produced several days  after the emission of  its gaseous precursors.
24      The excess PM2 5 concentration in urban centers  suggests that primary emissions such as
25      automobile exhaust, heating furnaces, and are responsible for much the urban PM2 5
26      hot-spots.
27           The reported  AIRS PM2 5 concentrations over the Pacific states  are generally higher
28      and average at 20 to 50 /-tg/m3.  This is 5 to 10 times higher than their companion
29      IMPROVE PM2 5 concentrations.  The dramatic difference is attributable to the pronounced
30      concentration differences between urban-industrial-agricultural centers  that occur in
       April 1995                              6-51       DRAFT-DO NOT QUOTE OR CITE

-------
SO
to
H
6
o
z
o
H
O
d
o
                                                                           PM10 AIRS - PM10 IMPROVE
                       PM10 AIRS - PM10 IMPROVE
                                                                           PM10 AIRS - PM10 IMPROVE
                       PM10 AIRS - PM10 IMPROVE
*   Figure 6-26.
n    6
H
W
                    Spatial maps of PM10 concentration difference between AIRS and IMPROVE/NESCAUM networks.

-------
  1      mountainous air basins and the concentrations monitored at remote national parks and
  2      wilderness areas that are generally at higher elevations.  However, it is fair to presume that
  3      the AIRS and IMPROVE PM2 5 data sets represent the extreme of aerosol concentration
  4      ranges that exist over the western US.  The challenging task of filling in the details,
  5      i.e. spatially and temporally extrapolating the aerosol concentrations over the rugged western
  6      United States is discussed in further detail in later regionally and locally focused sections
  7      below.
  8           It is also instructive to compare the seasonality of the urban (AIRS) concentrations to
  9      the non-urban (IMPROVE/NESCAUM) data.  In Figure 6-27 the difference in PM10, PM2 5,
 10      and PMCoarse between AIRS and IMPROVE/NESCAUM sites, using all available data,  is
 11      used to indicate the urban excess particle  concentration compared to the rural concentration.
 12           Nationally, the urban excess fine particle concentration ranges between  18 /-tg/m3 in
 13      December through February and 10 /ug/m3 in April through June (Figure 6-27a).  The urban
 14      excess coarse mass  concentration is less seasonal  ranging between 10 to 7 />tg/m3.  The sum
 15      of the fine and coarse national urban excess mass concentration is about 25 /zg/m3  in the
 16      winter season, and  18 ^g/m3 during spring season.  Hence, the nationally aggregated urban
 17      and non-urban data  confirm that urban area have  excess concentrations on the  order of 20
 18      Mg/m3, and well over half is due to fine particles, particularly in the winter season.
 19           The urban excess (AIRS-IMPROVE/NESCAUM difference) over the eastern United
 20      States (Figure 6-27c) shows fine particles excess of 8 to 12 /*g/m3, with higher value
 21      occurring during both winter and summer. The urban excess coarse mass in the eastern
 22      United States is only 5 to 8 /ig/m3, peaking during spring and summer.  The sum of fine  and
 23      coarse urban excess is 15 to 18 jug/m3 throughout the year.
 24           The excess urban (AIRS-IMPROVE/NESCAUM) aggregated over the western United
 25      States is  much more pronounced in magnitude and seasonality.   The urban excess fine mass
 26      is about 30 /-tg/m3 in November through January and drops to 8 to 10 /xg/m3 in April through
27      August.  The urban excess coarse mass is less in  magnitude and seasonality 15 to 18  /-ig/m3
28      in July through December, and 10 to 12 /ng/m3 in March through May.  The sum of the
29      urban excess fine and coarse mass is 40 to 50 /ng/m3 in November through January and about
30      20 fJLg/m3 in the spring March through June.  The urban AIRS  and non-urban IMPROVE)
31      networks in the western United States clearly monitor distinctly different aerosol  types, as

        April 1995                               5.53       DRAFT-DO NOT QUOTE OR CITE

-------
              US urban excess
                                            Eastern US urban excess
                                                          Western US urban excess
         60
 a
D
O

Z
O
H
O
a
o
H
W

O
>0
n
h-H
H
W
       )
      3

      O
      r™
      s
      a.
         50 --
         40 --
     30
         10 -
                                -a
                          H	h
                                               60
TO    ™    CO"
->    5    5
     •Fine+Coarse Mass
                          CL
                          (1)
                          CO
      >
      O
                                               50 -
                   40

                n


                B)

                3.  30 -
                o
                r-
                5
                Q.
  c
  03
H	1	1	1	h-
•3    Q.    >
•2    m    o
     co    z
           c
           ro
Fine - a - Coarse
• Fine+Coarse Mass
Fine - o - Coarse •
                                                                                  • Fine+Coarse Mass
                                                                             Fine	a - Coarse
Figure 6-27.      Urban excess concentrations (AIRS minus IMPROVE) for the U.S., Eastern U.S., and Western U.S.

-------
  1      stated by their respective goals and mandates.  The urban non-urban difference is such that
  2      the western non-urban concentrations are virtually irrelevant for the much higher urban
  3      values, particularly in the winter season. On the other hand, the eastern urban sites are
  4      greatly influenced by the non-urban,  regionally representative concentrations, particularly in
  5      the summer season.
  6
  7
  8      6.4  Regional Patterns and Trends
  9            This section describes the spatial, temporal, size, and chemical characteristics of seven
 10      aerosol regions of the conterminous US.  The  size and location of these regions were  chosen
 11      based mainly on the characteristics of their aerosol pattern.  The main criteria for delineating
 12      a region were 1) the region had to posses some uniqueness in aerosol trends, seasonality, size
 13      distribution,  or chemical composition;  2) each territory of conterminous United States had to
 14      belong to one of the regions;  3)  for  reasons of computational convenience the shape of the
 15      regions were  selected  to be rectangular on unprojected latitude longitude maps. The resulting
 16      criteria yielded seven  rectangular aerosol regions as shown in Figure 6-28.  It is recognized
 17      that this selection is arbitrary  and for future analysis additional regional  definition criteria
 18      would be desirable.
 19            For sake of  consistency and intercomparisons each region is described using maps
20      delineating the spatial  pattern and the sampling locations (Figure section a).  Monthly
21      concentrations for a given region were computed by averaging all the available data for the
22      specific month. In case of non-urban aerosol chemistry some regions only had 2 to 4
23      monitoring stations. The monthly PM2 5, PMCoarse and PM10 (Figure section b) over
24      regions illustrate the relative  seasonality of each aerosol type.  The non-urban regional
25      average chemical composition is  presented as seasonal charts of chemical aerosol components
26      as a fraction of the fine mass  concentration (Figure section c).  The role of some primary
27      sources, such as coal and fuel oil combustion is indicated through seasonal charts of selenium
28      (coal) and vanadium (fuel oil) trace metals (Figure section d).
29            In addition, for each region figures will  be provided  showing short term variability of
30      PM10 concentrations and PM10 urban excess.
        April 1995                                6-55       DRAFT-DO NOT QUOTE OR CITE

-------
                           PM10     =     01
                           PM2.5    =      12.
                           PM2.5/10 =  Q« 38
               17
       Figure 6-28 Aerosol regions of the conterminous US.
 1     6.4.1  Regional Aerosol Pattern in the Northeast
 2           The Northeast aerosol region covers the New England states, including eastern
 3     Pennsylvania and eastern Virginia to the south (Figure 6-29a).  In the Northeast, terrain
 4     features that significantly influence regional ventilation occur over the mountainous upstate
 5     New York, Vermont and New Hampshire.  Throughout the year, the Northeast is influenced
 6     by Canadian as well as Gulf airmasses.   The region includes the Boston-New York
 7     megalopolis, as well as other urban-industrial centers. It is known that the Northeast is
 8     influenced by  both local sources, as well as long range transport of fine particle haze from
 9     other regions.
10           Data from a two year fine particle network in the Northeast (Bennett et al., 1994)
11     yielded a geometric mean concentration of PM2 5 of  12.9  and paniculate sulfur (1.4
       April 1995
6-56
DRAFT-DO NOT QUOTE OR CITE

-------
  1      equivalent to 4.2 ptg/m3 of sulfate), which is somewhat lower than other comparable rural
  2      data.
  3
  4      6.4.1.1  Non-urban size and chemical composition in the Northeast
  5            The summary of the non-urban aerosol chemical composition in the Northeast is
  6      presented in Figure 6-29c. The region has 14 monitoring sites, 7 of which are part of
  7      NESCAUM in the New England states.
  8            The PM10 concentration exhibits a factor of two seasonal amplitude between 12 /ug/m3
  9      in the winter, and 25 /xg/m3 in June and July  (Figure 6-29b).  About 60% of PM10 is
 10      contributed by fine particles throughout the year, and also contribute to the summer-peaked
 11      seasonality.
 12            Sulfates are the most important contributors of the fine particle mass in the Northeast,
 13      particularly in the summer season when they account for half of the fine mass.  The organics
 14      account for 30 to 40%, with  the higher fractions occurring in the fall and winter, September
 15      through January.  In fact, during the late fall the sulfate and organic contributions are
 16      comparable at 40%.  Fine particle soil is remarkably unimportant throughout the year
 17      (<5%).  Soot on  the other hand, is more significant particularly during the fall when it
 18      contributes about 10% of the fine mass.  The sum of the above four non-urban fine particle
 19      aerosol components, account  for over 90% of the measured fine particle mass throughout the
 20      year.  This indicates that nitrates, trace metals and sea salt are of minor importance in the
 21      northeastern U.S.  fine particle chemical mass balance.
 22            The seasonality of both selenium and vanadium indicates a winter peak (Figure 6-29d).
 23      In particular, the vanadium concentration increases by factor of two for December and
 24      January compared to the summer values. Also, the  V concentration is  higher than over any
 25      other region indicating the strongest contribution of fuel oil  emissions.  The S/Se ratio is
 26      strongly seasonal with the winter value of 1,000 and the summer peak of 2,000 to 2,500.
27
28      6.4.1.2 Urban aerosols in the Northeast
29            The Northeast region shows a substantial (39%) decline in PM10 concentration, from
30      36 /ig/m3 in 1985 to 22 jug/m3 in 1993 (Figure 6-30b).  The standard deviation among the
       April 1995                               6-57       DRAFT-DO NOT QUOTE OR CITE

-------
             PM2.5 - Nonurban PM Monitoring Sites
      PM10, PM2.5 and PMC Monthly Avg.
         Northeast - IMPROVE/NESCAUM Networks
        60,000,	.	,	.	.	,	,	,	.	.	.	.
                                                     55,000

                                                     50,000

                                                     45,000

                                                     40,000

                                                     35,000

                                                     30,000

                                                     25,000

                                                     20,000

                                                     15,000

                                                     10,000

                                                      5,000
                                                  , -e- PMIO
           1989 Feb Mar Apr May Jun Jul Aug Sep Oct  Nov Dec

                 |~f-PM2.5       -A- PMCoffrs*
                     Northeast
                                                                  Northeast
     u.
     •8
     S
     •8
0.9


0.8


0.7


0.6


0.5

0.4


0.3


0.2


0.1


0.0	
 1989 Feb Mai Api May Jun Jul  Aug Sep Od Nov Due

Sulfate           -& Organics

Soil            ;V" SooT'"_'_  ~~_"

Sulf^Qra+Soil+Soot
                                                        1989 Feb Mat Apr  May Jun Jul Aug Sep

                                                   ', -A- Sulfur   i -3 Selenium   -I- Vanadium
                                              Scale   0-4000
                                                                0-4
                                                                          0-10
                                     Oa  Nov Dec

                                      SJSe


                                     0-4000
Figure 6-29.       IMPROVE/NESCAUM concentration data for the Northeast.
                    (a) Monitoring locations,  (b) PM10, PM2.5, and PMCoarse. (c)
                    Chemical fraction of sulfate, soil, organics, and soot,  (d) Tracer
                    concentrations.
April 1995
6-58
                                              DRAFT-DO NOT QUOTE OR CITE

-------
  1      monitoring stations for any given year is about 30%.  The map of the Northeast shows the
  2      magnitude of PM10 concentrations in proportion of circle radius.  It is evident, that the
  3      highest AIRS PM10 concentrations generally occur in urban centers (Figure 6-30a).
  4            The seasonality of the Northeast PM10 concentration (Figure 6-30d) is a modest 20%,
  5      ranging from 25 to 31 /xg/m3.  There is a summer peak in July, and a rather uniform
  6      concentration between September and May  showing only  a slight winter peak.
  7            The PM2 5-PM10 relationship in the scatter charts (Figure 6-30c) show that on the
  8      average 62% of PM10 is contributed by  fine particles.  During high levels of PM10, fine
  9      particles dominate some of the months,  while coarse particles during other months. It is
 10      likely that urban and rural sampling locations exhibit different PM2 s-PM10 relationship.
 11            In general, the regional  scale emissions are not expected  to vary significantly from one
 12      day to another. However, both meteorological transport, i.e. dilution, as well as aerosol
 13      formation and removal processes are important modulators of daily aerosol concentration.
 14      Daily concentration of particulate matter exhibits strong fluctuation from one day to another,
 15      mainly due to  the role of the meteorological transport  variability.  The AIRS PM10 database
 16      reports the concentrations every sixth day, synchronously over the entire country.  The
 17      sample duration is one day  which, over  the long run provides the  concentration distribution
 18      function of daily samples.   For determination of the effects (human health, visibility, acid
 19      deposition) the concentration has to be known at the specific location where the sensitive
 20      receptors reside.  Also the concentrations have to  be known at a short, e.g. daily time scale,
 21      as well as over the long-term.
 22            In order to characterize  the one day-scale temporal variation over a given region, the
 23      entire available data aggregated over the entire region for each monitoring day are plotted  as
 24      time series.  It is recognized that during the other five non-monitored days, the
 25      concentrations may be different from the reported value.  The six day sample increment
26      ensures that both weekday and weekend data are properly taken into account.   The physical
27      interpretation of regionally averaged daily concentration is a measure of the regional scale
28      meteorological ventilation.  High regionally averaged concentrations indicate poor
29      ventilation, i.e. combination of low wind speeds and low  mixing heights and the absence of
30      fast aerosol removal rates, such as cloud scavenging and precipitation.  Low regional
31      concentrations, on the other hand, represent either strong  horizontal transport, deep mixing

        April  1995                                 6-59       DRAFT-DO NOT QUOTE OR CITE

-------
              Average AIRS PM10 -1986-1993
                                                  AIRS PM10 Concentration Trends
                                                              Northeast
                                                  1985 1986  1987  1988  1989 1990  1991  1992


                                                           -5 PM10AYG-SIG --,, PM10AVG + SIG
   AIRS PM2.5 vs. PM10 - Monthly Avgerages
                  Northeast
     ISO

     140
        AIRS PM10, PM2.S and PM Coarse Cone.
                    Northeast
        Corr Coeff
        -Slope
        Y offset
        -Data Points
                                             u
                                             
                                             Si.
       60

       55

       60

       45

       40

       35

       30

       25

       20

       15

       10

       £
              30      60      90

           PM10_AVG UG/CU METER (25 C)
       1986   Mar Apr May Jun Jul  Aug Sep oct Nov Oec

       !T!r PM10    -3- Fine     -+ Coarse
Figure 6-30.       AIRS concentration data for the Northeast, (a) Monitoring locations.
                   (b) Regional PM10 concentration trends,  (c) PM10, PM2-5
                   relationship,  (d) PM10, PM2>5, and PMCoarse seasonal pattern.
April 1995
6-60
DRAFT-DO NOT QUOTE OR CITE

-------
  1     heights, or high regional removal rates.  Advection of high aerosol content air masses from
  2     neighboring regions may also be a cause of elevated concentration in a given region.
  3           The daily variation of the regional averaged urban PM10 concentration for the
  4     Northeast is shown in Figure 6-31.  As a guide to the eye the single day concentration data
  5     for every sixth  day are connected by a line between the data points, although five in-between
  6     days are not monitored. The lowest  regionally averaged daily urban PM10 is about 10 /xg/m3,
  7     while the highest is about 55 /ig/m3, with a regional average of 25? /ig/m3.  The highest
  8     concentrations (> 40 /ng/m3) occur primarily in the summer season.  The tune series also
  9     indicate that the high concentration episodes do not persist  over multiple six day periods.
 10     This is consistent with the notion that the regional ventilation that is caused by synoptic scale
 11     airmass changes, which typically occur every four to  seven days over eastern US.  The daily
 12     time series also convey the fact that day  to day variation in PM10 is higher than the seasonal
 13     amplitude.  In fact, visual inspection of Figure 6-31 the concentration seasonally is barely
 14     discernible.  It can be stated, therefore, that the PM10 concentration in the Northeast is
 15     highly episodic, i.e. the temporal concentration variation is both substantial and irregular.
 16     The excess urban PM10 (AIRS-IMPROVE) is shown in Figure 6-32.
 17
 18     6.4.2 Regional Aerosol Pattern in the Southeast
 19           The Southeast  rectangle stretches from North Carolina  to eastern Texas (Figure 6-33).
 20     From the point  view of regional ventilation the Southeast terrain is flat, with the exception
 21     ofmildly rolling the southern Appalachian Mountains. The region is known for increasing
 22     population over the past decades, high summertime humidity, and poor regional ventilation,
 23     due to stagnating high pressure systems.
 24
 25     6.4.2.1 Non-urban Size and Chemical  Composition in the  Southeast
 26           The non-urban PM10 concentration in the Southeast (Figure 6-33b) is roughly
27     comparable to the  Northeast, exhibiting about factor of two seasonal concentration amplitude
28     between 12 /*g/m3 in the winter,  and 25 /ig/m3 in the  summer.  An anomalous high PM10
29     concentration is recorded for July which is contributed exclusively by excess coarse particle
30     concentrations of about 10 /ig/m3. With  exception of July, the fine particle mass accounts
       April 1995                                6-61      DRAFT-DO NOT QUOTE OR CITE

-------
              o
           O UJ
              D
              O
              o
                              Northeast
                            Every Sixth Day
                            1991
 1992
                                             1993
Figure 6-31.      Short term variation of PM10 average for the Northeast. Data are
                reported every sixth day.


                       Northeast urban excess
                     Jan   Mar   May   Jul   Sep   Nov
Figure 6-32.      Urban excess concentration (AIRS minus IMPROVE) for the
                Northeast.
April 1995
6-62
DRAFT-DO NOT QUOTE OR CITE

-------
  1      for about 70% of the non-urban PM10, leaving the coarse mass of 30% or less throughout the
  2      year (Figure 6-33b).
  3            The most prominent aerosol species in the Southeast are sulfates contributing 40 to
  4      50% of the fine mass (Figures 6-33c).  The  anomalous sulfate fraction (35%) coincides with
  5      the high (20%) soil contribution during July. During other months, soil contribution is
  6      <5%  of the fine mass.  The relative role of the organics in the non-urban Southeast is most
  7      pronounced during the winter (40%), but declines to 25%  during the summer months. The
  8      soot contribution varies between 2% in the summer to 6% in the winter months.
  9            The trace element concentrations of selenium and vanadium (Figure 6-33d) are
10      constant throughout the year, implying that the combined role of emissions and dilution is
11      seasonally invariant.  The concentration of sulfur, on other hand shows a  definite summer
12      peak, that is 2 to 3 times higher than the  winter concentrations.   Consequently, the S/Se ratio
13      is strongly seasonal.  In fact, the warm season S/Se ratio of 2,500 is higher than over any
14      other region of the country.  If Se-bearing coal combustion is the dominant source of sulfur
15      in the  Southeast, than the high S/Se ratio implies that the sulfate production per coal
16      production in the summer is 2.5 times that in the winter.
17
18      6.4.2.2  Urban Aerosols in the Southeast
19            There is evidence of significant (34%) PM10 decline over the past decade (Figure
20      6-34b).  By 1993, the average PM10 was  24 t*g/m3. It is worth noting that this value is
21      higher than the corresponding 1993 concentration for the Northeast (22 /ig/m3).  It is also
22      remarkable that the Southeast concentration trends and patterns most closely resemble the
23      industrial Midwest described  below.  The unique feature of the Southeast  is  the uniformity of
24      the aerosol concentration among the monitoring stations.  In fact the 17%  station to station
25      standard deviation is by far the lowest among the aerosol regions (Figure  6-34b).
26            The Southeast is  also characterized by high seasonal amplitude of 37 %, ranging
27      between 22 /xg/m3 in December through February and 35 ptg/m3 in July through August
28      (Figure 6-34d).  There is no evidence of a winter peak for the southeastern US.
29            The scattergram of PM2 5-PM10 for the Southeast (Figure 6-34c) shows an average of
30      58% fine particle contribution, with considerable scatter.  It should be  noted, however, that
       April 1995                               6-63       DRAFT-DO NOT QUOTE OR CITE

-------
           PM2.5 - Nonurban PM Monitoring Sites
                                          PM10, PM2.5 and PMC Monthly Avg.

                                            Southeast - IMPROVE/NESCAUM Networks
                                           60,000 r  .   .	.	.	,	.	,	.	,—,   ,
                                                     55,000


                                                     50,000


                                                     45,000


                                                     40,000


                                                     35,000


                                                     30,000


                                                     25,000


                                                     20,000


                                                     15,000


                                                     10,000


                                                      5,000
                                                        1983 Feb Mat Apr May Jun Jul Aug Sep Oct  Nov Dec


                                                 i -B- PM10      I -I- PM2.5       -A- PM CoQrse
                    Southeast
                                                                  Southeast
0.9


0.8


0.7


0.6


0.5


0.4


0.3


02


0.1


0.0
                                  *-.i --
                                  -''"    '-
           15(19 Feb Mar Apr May Jun Jul Aug Sep Oc» Nov Dec


        -6- Sulfate          43- Organics

        4 Soil            |~-r- Soot       	  !

        -c. Stiif+Oro+Soll+Soot
                                                     4,000



                                                     3,600



                                                     3,000



                                                     2,500
                                                  •| 2,000
                                                  c
                                                     1,500



                                                     1,000



                                                      600
                                              Scale    0-4000
                                              1909 Feb Mar Apr  May Jun  Jul Aug  Sep Oct  Nov Dec


                                            Sulfur   I -Q- Selenium  -4-  Vanadium -9- S/S«

                                                       n-4        0-10      0-4000
Figure 6-33.        IMPROVE/NESCAUM concentration data for the Southeast.
                     (a) Monitoring locations, (b) PM10, PM2 5, and PMCoarse.  (c)
                     Chemical fraction of sulfate, soil, organics, and soot,  (d) Tracer
                     concentrations.
April 1995
                                     6-64
DRAFT-DO NOT QUOTE OR CITE

-------
               Average AIRS PM10 -1986-1993
      AIRS PM2.5 vs. PM10 - Monthly Avgerages
                    Southeast
       160

       140 -
                30     60     90

             PM10_AVG UG/CU METER (25 C)
                                                   60
        AIRS PM10 Concentration Trends
                   Southeast
                                                    1985  1986 1987 1988 1989  1990  1991  1992
                                                ,-A- PM10AVG    I -O- PM10AV6-SIC  -4- PM10AVG + SIG
         AIRS PM10, PM2.5 and PM Coarse Cone.
                    Southeast
                                               o

                                               a:
                                               |

                                               y
         1*88   Mar Apr May Jun Jul Aug s«p Get Nov D«c

         I-A- PM10   I -a- Fine    -+ Coarse
Figure 6-34.       AIRS concentration data for the Southeast, (a) Monitoring locations.
                   (b) PM10, PM2 5, and PMCoarse.  (c) Chemical fraction of sulfate,
                   soil, organics,  and soot,  (d) Tracer concentrations.
April 1995
6-65
DRAFT-DO NOT QUOTE OR CITE

-------
 1      size segregated samples were only available briefly for two monitoring sites which may not
 2      be representative for the large southeastern region.
 3            The regionally averaged daily PM10 concentrations over the Southeast (Figure 6-35)
 4      shows a clearly discernible seasonality.  The concentrations during the winter months are
 5      about factor of two lower than during the  summer.  Overall,  the lowest concentrations are
 6      about 14 /xg/m3, and the highest about 50  /xg/m3, which is about factor of three.  However,
 7      seasonality of the temporal signal accounts to about half of the variation.  Hence, within a
 8      given season the sixth day to sixth day variation is only about 50%.  It can be concluded that
 9      the PM10 concentration over the southeastern United States region is quite uniform in time,
10      although it exhibits a substantial seasonality.   It is to be noted that the southeastern United
11      States also exhibits the highest spatial  homogeneity, i.e. the average deviation of average
12      concentrations between the stations. A further index of the short-term temporal variation is
13      given in the seasonal maps of concentration standard deviation (Figure 6-34b).  The PM10
14      urban excess (AIRS-IMPROVE) for the  southeast region is given in Figure 6-36.
15
16      6.4.3  Regional Aerosol Pattern in the Industrial Midwest
17            This aerosol region stretches between  Illinois and western Pennsylvania,  including
18      Kentucky on the south (Figure 6-37a). The industrial  Midwest is covered by flat terrain with
19      the exception of the central Appalachian Mountains in the Virginias. In the winter the region
20      is under the influence of cold Canadian air masses, while during the summer moist airmass
21      transported from the Gulf Coast prevail.  This region includes the Ohio and Mississippi
22      River Valleys that are known for high sulfur emission densities.  The region also includes
23      major metropolitan areas.
24
25      6.4.3.1 Non-urban Size  and Chemical Composition in the Industrial Midwest
26            The seasonal pattern of the non-urban aerosol in the Industrial Midwest is shown in
27      Figure 6-37b. It is worth noting that  the regional observations are based on a few
28      monitoring sites and their representativeness  is questionable.  The PM10 concentrations range
29      between 10 and  22 /zg/m3, comparable to  the non-urban levels in other eastern  U.S. regions.
30      It is quite remarkable that 70 to 80%  of PM10 is contributed  by fine particles throughout the
31      year.  In fact, the coarse particle concentrations are  4  to 5 /ng/m3, which is lower than over

        April 1995                               6-66      DRAFT-DO NOT QUOTE OR CITE

-------
              0!
           O UJ
           is
           0- 5
              3
              O
              O
              3
                               Southeast
                             Every Sixth Day
                           1991
                                    1992
          1993
Figure 6-35.     Short term variation of PM10 average for the Southeast. Data are
               reported every sixth day.
                        Southeast urban excess
                     0 H	1	1	1	1	1
                      Jan   Mar   May   Jul    Sep   Nov

Figure 6-36.     Urban excess concentration (AIRS minus IMPROVE) for the
               Southeast.
April 1995
6-67
DRAFT-DO NOT QUOTE OR CITE

-------
          PM2.5 - Nonurban PM Monitoring Sites
                                                 PM10, PM2.5 and PMC Monthly Avg.
                                                   Industrial - IMPROVE/NESCAUM Networks
                                                  80,0001	.	.	.	.	,	,	,	,	,	,	
                                                     1983 Feb Mai Apt May Jun Jul  Aug Sep Oct Nov Dec


                                                  PM10      ] -4- PM2.5       -£- PM CoQrs«
               Industrial Midwest
                                                            Industrial Midwest
 S
 c
0.9


0.8


0.7


0.6


0.6

0.4


0.3


0.2


0.1 -
        0.0
           I- -Q.
          1989 Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

      -A- Sulfate          -a- Organics

      •f Soil           tj:_???!.	~Z	:

      -(v SulftOrg+SoiHSoot
                                            1983 Feb Mar Apr  May Jun Jul Aug Sep Oct Nov Oac

                                        -A- Sulfur   -3. Selenium  -4- Vanadium ;_-3_&fSe    ;

                                  Scale    0-4000      O^t        0-10     0-4000
Figure 6-37.       IMPROVE/NESCAUM concentration data for the Industrial
                    Midwest,  (a) Monitoring locations,  (b) PM10, PM2 5, and PMCoarse.
                    (c) Chemical fraction of sulfate, soil, organics, and soot.  d)Tracer
                    concentrations.
April 1995
                                     6-68
DRAFT-DO NOT QUOTE OR CITE

-------
  1     any other region of the US.  Hence, the contribution of wind blown dust, fly ash, or other
  2     man-induced dust entrainment is not a significant factor in the non-urban areas of the
  3     Industrial Midwest.
  4           The chemical mass balance (Figure 6-37c) shows that sulfates are 45 to 55 %  of the
  5     fine mass which is higher than the sulfate fractions in other regions. Organics exhibit a
  6     variable contribution that is high (40%) during the cold season (October through February)
  7     and remarkably low (20%) in July and August.  The strong winter peak for the organic
  8     fraction differs markedly from the Northeast where the organics are aseasonal.  Another
  9     unusual feature of the chemical mass balance is that the sum of sulfate, organics soil and soot
 10     is about 75 % during the summer and 95 % in the winter.  It is not known what is the
 11     composition of the missing 25% during the summer time.
 12           Chemical tracer data is shown in Figure 6-37d.  The chemical tracer for coal
 13     combustion, selenium ranges between  1,000 and 1,500 ng/m3, which is higher than  in any
 14     other region.  There is a  sizeable month to month variationin Se concentration (partly due to
 15     a small number of data points) and the seasonality is not appreciable.  This means that the
 16     combined effects of coal combustion source strength and meteorological dilution is seasonally
 17     invariant over  the industrial Midwest.  The concentration of vanadium, which is a tracer for
 18     oil combustion is low  throughout the year. The concentration of fine particle sulfur exhibits
 19     random monthly variation but indicates a summer peak.  The  S/Se ratio is a rather smooth
 20     seasonal curve ranging between 1,000  in the winter and 2,000 during the summer months.
 21      Hence, the sulfate yield is about twice  during the summer than during winter months.  For
 22     comparison both the Northeast and Southeast exhibit higher seasonality (factor of 2.5) in
 23      S/Se ratio.  A  possible explanation for this change in S/Se ratio is that over the industrial
 24      Midwest the average age of the  aerosol producing emissions is less than over the Northeast
 25      or Southeast.  Alternatively, the sulfate formation rate may be higher over the Northeast and
26      Southeast.
27
28      6.4.3.2 Urban Aerosols  in the Industrial Midwest
29           The PM10 concentration trends for the industrial Midwest (Figure 6-38b) show a
30      decline of 34% (from 38 to 25 /zg/m3)  between 1985 and 1993. There is also a 28%
       April 1995                               6-69       DRAFT-DO NOT QUOTE OR CITE

-------
                Average AIRS PM10 -1985-1993
                                                    AIRS PM10 Concentration Trends
                                                                 Industrial
                                                    60
                                                     1985 1986 1987 1988  1989  1990 1991 1992
       AIRS PM2.5 vs. PWI10 -Monthly Avgerages
                  Industrial Midwest
        160
                                                    PM10AVG    i -Q- PM10AVC-SIG  -+ PM10AVG + SIC
                                   AIRS PM10, PM2.5 and PM Coarse Cone.
                                           Industrial Midwest
                                                LU
                                                tn
                                                CJ
                                                s
                 30     60      90     120

               PM10 JVVG UG/CU METER (25 C)
                                 1986   Mar Apr May Jun Jul Aug 8«p Oct Nov Dec

                                 j-A- PM10   i -a- Fine     -+• Coars*
Figure 6-38.
AIRS concentration data for the Industrial Midwest,  (a) Monitoring
locations, (b) PM10, PM2 5, and PMCoarse. (c) Chemical fraction of
sulfate, soil, organics, and soot,  (d) Tracer concentrations.
April 1995
                        6-70
DRAFT-DO NOT QUOTE OR CITE

-------
  1     deviation among the stations within the region.  As in the Northeast, the higher
  2     concentrations occur within urban-industrial areas (Figure 6-3 8a).
  3           The PM10 seasonality (Figure 6-38d) is virtually  identical (37%  amplitude) to the
  4     seasonality of the Southeast: the lowest concentrations (25 /-tg/m3) occur between November
  5     and February, while the highest values are recorded in June through August (40 jug/m3).  It
  6     is quite  remarkable that throughout the 1980s of PM10,  the trends and the seasonality of the
  7     midwestern PM10 aerosols is comparable to that of the Southeast.  At this time, the only
  8     rationale for separating into a region is the belief that the source types  and the aerosol
  9     composition may possibly be different from the Southeast.
 10           Fine particles contribute 59% of the PM10 concentration on the average (Figure
 11     6-38c),  and high PM10 can occur when either fine or coarse particles dominate.  It is notable
 12     that size segregated samples are available primarily from urban-industrial sites over the
 13     industrialized Midwest.
 14           Daily concentration over the industrial Midwest (Figure 6-39) varies between 50 and
 15     75 />ig/m3. The lowest regional concentrations occur during the winter  months, while the
 16     highest values (in excess of 40 jug/m3) occur during the summer.  It is  evident, that
 17     seasonality is an important component of the time series, accounting for about half of the
 18     variance.  The elevated concentrations occur only  one sixth day observation at the time,
 19     indicating  general absence of prolonged episodes that last 12 days or more. The industrial
 20     Midwest also show substantial spatial variability.  The urban excess PM10 (AIRS-IMPROVE)
 21      for the industrial midwest is given in Figure 6-40.
 22
 23      6.4.4  Regional Aerosol Pattern in the Upper  Midwest
 24            The upper Midwest covers the agricultural heartland of the country (Figure 6-40). The
 25      region is void of any terrain features that would influence the regional ventilation.  Industrial
 26      emissions and the population density are comparatively low. However, the relatively high
 27      PM10 concentrations in this region warrant a more detailed examination.  In the winter, the
28      region is covered by cold Canadian airmasses, while in the summer moist Gulf air alternates
29      with drier  Pacific airmasses.
       April 1995                                6-71       DRAFT-DO NOT QUOTE OR CITE

-------
          O UJ

          it
          0- 5
             D
             O
             O
                              Industrial Midwest
                               Every Sixth Day
                           1991
 1992
 1993
Figure 6-39.     Short term variation of PM10 average for the Industrial Midwest.
               Data are reported every sixth day.
                       In. Midwest urban excess
                     Jan   Mar  May   Jul   Sep   Nov



Figure 6-40.     Urban excess concentration (AIRS minus IMPROVE) for the

               Industrial Midwest.
April 1995
6-72
DRAFT-DO NOT QUOTE OR CITE

-------
  1      6.4.4.1  Non-urban Size and Chemical Composition in the Upper Midwest
  2            Over the non-urban areas of upper Midwest the PM10 concentration is about 8 /ug/m3
  3      during November through April winter season, and increases to 15 /xg/m3 during the
  4      summer.  Fine and coarse particles have a comparable contribution to the PM10 mass (Figure
  5      6-41b).
  6            The chemical mass balance  (Figure 6-4 Ic) indicates that during the March through
  7      May spring season  sulfates dominate, but during July through October season organics
  8      prevail.  This is a rather unusual pattern not observed over any other region. The
  9      contribution of fine particle  soil exceeds 10% in the spring as well as in the fall season.
10            Chemical tracers are  shown in Figure 6-41d. Selenium  concentration is low
11      throughout the year (400 to  600 ng/m3), but the highest concentrations are observed during
12      the summer.   This suggest that either the Se sources or the Se transport into the Upper
13      Midwest from other regions is strongest in the summer.  The concentration of the fine
14      particle sulfur is  < 500 ng/m3 throughout the year, but somewhat higher during March and
15      April. The spring peak for  fine particle sulfur has not been observed in any other region.  It
16      is also worth noting that S/Se ratio is the highest during the spring and lowest in July
17      through September.  This hints on the possibility that over the Upper Midwest additional
18      sources of fine particle sulfur are present for which Se is  not a tracer.  Here again, it needs
19      to be pointed  out that the above chemical patterns are based on only two monitoring stations.
20
21      6.4.4.2  Urban Aerosols in the Upper Midwest
22            The agricultural upper Midwest (Figure 6-42) shows the smallest decline  among  the
23      aerosol regions.  Over the past decade the region average PM10 concentration ranged
24      between  25 and 31  /-ig/m3.  Some reduction (19%) is evident since 1989.  As over the
25      eastern US, the highest concentrations occur in the vicinity of urban areas. Some of the
26      station to station concentration spread arises  from low concentrations over western North
27      Dakota.  On the average,  the deviation among the stations over the region is a moderate 30%
28      (Figure 6-39).  The upper Midwest is also unique in that it shows the regionally lowest
29      seasonal  amplitude of 19%,  with the  slightly lower concentrations occurring in December and
30      January.  The sparse size  segregated  data indicate that only 38% of PM10 is contributed by
       April 1995                               5.73       DRAFT-DO NOT QUOTE OR CITE

-------
            PM2.5 - Nonurban PM Monitoring Sites
                             PM10, PM2.5 and PMC Monthly Avg.
                             Upper Midwest - IMPROVE/NESCAUM Networks




1
O)
c



OU,VW
55,000
50,000
45,000
40,000
35,000
30,000
26,000
20,000
15,000
10,000
6,000
n
-
-
-
-
-
-
/"-%^^
£^-*~-3>^
                                                   1989 Fob Mar Apr May Jun Jul Aug Sep Oct Nov Dec

                                                        JH PM2.5      -&- PM Coarse
                   Upper Midwest
                                         Upper Midwest




10
s
5
V
\L
•6
g

C
""











03
0.8

0.7
0.6

05

0.4

03


02

0.1

-
_•' * ' '-•
-^ :- ~
''
-

/'X &-~a
/ ^\^ /
(^^ 	 &/ Ch- B ^s— ^
x°\ 	 ' \/f^
H' X^6 ^a^is- -^ —
^-•s- a-' *
T3
-
.-4-- a. •-'*'
.- ''•+--. 4. ___..-' '^..4.
>"-"«-" 1^ - 0- . <: . _r,_ ^ _..- -Q- -* -9 -'"?- "
3,500

3,000

2,500

1 2,000
c

1,600


1,000


500

0.0' 	 • 	
inn1! Fpb Mar Anr Uav Jun Jul Aurj Spp r^ri Nnw npr A
_

-

-

-
..,

-
:•- —
^
'-'-
a •' — •> •'
A-^^^^ „ 	 '^
~"~~&-& -^'"^^^'^^^^^'^sfe^i^^^
^-.+- -M — h---1--.^.-^" ^~a»:-«-i;:S'-

1989 Feb Mar Apr May Jun Jul Aug Sep Ocl Hov Dec
-&- Sulfate -B- Organics 	 __
i — -— 	 ] -A- sum.
+ Soil 1-0- Soot J
r -a Selenium -4- Vanadium I -B S/Se

-0 SurfVOrg+Soll+Soot Scale 0-4000 O^t 0-10 0^000
Figure 6-41.
EMPROVE/NESCAUM concentration data for the Upper Midwest.
(a) Monitoring locations, (b) PM10, PM2 5, and PMCoarse.
(c) Chemical fraction of sulfate, soil, organics, and soot,  (d) Tracer
concentrations.
April 1995
                      6-74
                                                   DRAFT-DO NOT QUOTE OR CITE

-------
  1      fine particles.  This is  and indication that wind blown dust from natural or man-induced
  2      sources prevails.  In this sense, the region is similar to the Southwest.
  3            The daily regionally averaged PM10 concentrations  in the upper Midwest (Figure 6-43)
  4      range between 15 and 45 ^g/m3.  The highest values (>40 jug/m3) generally occur in the
  5      summer season, while the low regional concentrations occur mainly in the cold season, but
  6      low values also occur  in the summer.  It is interesting that the lowest PM10 concentrations
  7      over the upper Midwest (15 /ug/m3) are comparable to the Southeast and the industrial
  8      Midwest, but differ from these regions by the absence of high concentration events or
  9      episodes. In fact, the  PM10 "episodes" over the upper Midwest are all in 40 to 45 /ug/m3
 10      concentration range, compared to 50 to 75 /*g/m3 in the Midwest.  The seasonality is barely
 11      discernible from the time series confirming that the day  to day variation exceeds the seasonal
 12      modulation.  The urban excess PM10 (AIRS-IMPROVE) for the upper midwest is  given in
 13      Figure 6-44.
 14
 15      6.4.5  Regional Aerosol Pattern in the Southwest
 16            The Southwest covers the arid states from western Texas to Arizona  (Figure 6-45a).
 17      The Southwest is characterized by mountainous terrain features between the southern Rockies
 18      and the Colorado Plateau.  The industrial activity and  agriculture is minor compared to other
 19      regions.  Major population centers are El Paso, Phoenix, and Tucson.   The meteorology of
 20      the region is characterized by  low annual precipitation, except during the  summer monsoon,
 21      July through September, when moist air penetrates from the Gulf of Mexico toward the
 22      western states, bringing moisture  and precipitation.
 23            The geographic  pattern  Figure 6-45a of hours of blowing dust shows that western
 24      Texas and southern California deserts are the dustiest regions of the country. A forty-year
 25      trend of dust hours shows that the 1950s had several times the dust occurrence of the 1980s
26      (Patterson et al., 1994). It is  likely, that during the severe droughts of the 1930s the dust
27      frequency was even higher.
28            Wind erosion is linked to relative humidity as well as wind speed, the combined
29      changes in both variables can  sometimes trigger dust storms.  An example is the dust storm
30      in November, 1991 that caused the severe accident on Interstate 5 in California. Seventeen
31      people died in this 164-car accident with low visibility from dust (Gregory et al., 1994).

        April 1995                                6-75       DRAFT-DO NOT QUOTE OR CITE

-------
               Average AIRS PM10 -1986-1993
                                                   AIRS PM10 Concentration Trends
                                                             Upper Midwest
      AIRS PM2.5 vs. PM10 - Monthly Avgerages
                   Upper Midwest
          0     30     60      90     120
              PM10.AVG Ufi/CU METER (25 C)
                                                   1985 1986 1987  1988 1989 1990 1991  1992

                                                           j -£] PM10AVG-SIG -4- PM10AVO + SIO
   !-&- PM10AVG
         AIRS PM10, PM2.5 and PM Coarse Cone.
                   Upper Midwest
                                               u
                                               5
        60

        55

        50

        46

        40



        30



        20

        15

        10

         6
         1886   Mir Apr May Jun Jul Aug s«p Oct Nov 0«c

         I-A- PM10   I -a- Fin*     -4- Coarw
Figure 6-42.       AIRS concentration data for the Upper Midwest, (a) Monitoring
                   locations, (b) PM10, PM2.5, and PMCoarse.  (c) Chemical fraction of
                   sulfate, soil, organics, and soot,  (d) Tracer concentrations.
April 1995
6-76
DRAFT-DO NOT QUOTE OR CITE

-------
           o
           m
           CM
         O UJ
         It;
            =>
            O
            O
                           Upper Midwest
                          Every Sixth  Day
                          1991
                                  1992
                                          1993
Figure 6-43.     Short term variation of PM10 average for the Upper Midwest. Data
               are reported every sixth day.


                      Up. Midwest urban excess
                     Jan   Mar   May   Jul   Sep   Nov
Figure 6-44.      Urban excess concentration (AIRS minus IMPROVE) for the Upper
               Midwest.
April 1995
6-77      DRAFT-DO NOT QUOTE OR CITE

-------
                 PM2.S . Nonurfaan PM Monitoring Sites
                                                        PM10, PM2.5 and PMC Monthly Avg.
                                                         Southwest - IMPROVE/NESCAUM Networks
                                                         60,000
                                                            1389 Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

                                                                    I- PM2.5       -A- PM CoQrse
                        Southwest
                                                                     Southwest
        u.
        •8
        §
  0.9

  0.8

  0.7

  0.6

  0.5

  0.4

  0.3

  02

  0.1

  0.0
    1989 Feb Mai Apr May Jun Jul Aug Sep Oct Nov Dec

-A- Sulfate          -a Organics

-+• Soil

-«•• SulfHDrg+Soil+Soot
                                                         4,000


                                                         3,600


                                                         3,000


                                                         2,600
                                                         1,600


                                                         1,000


                                                          600
                                                            1983 Feb Mar Apr May Jun Jul  Aug Sep Oct Nov Dec

                                                       i-A- Sulfur  | -a- Selenium  -4- Vanadium -e- S/Se

                                                    Scale   0-4000      0-4        0-10     0-4000
Figure 6-45.        IMPROVE/NESCAUM concentration data for the Southwest, (a)
                    Monitoring locations, (b) PM10, PM2 5, and PMCoarse.  (c) Chemical
                    fraction of sulfate, soil, organics, and soot,  (d)  Tracer
                    concentrations.
April 1995
                                 6-78
DRAFT-DO NOT QUOTE OR CITE

-------
  1           Gillette and Sinclair, 1990 estimated that dust devils (strongly spiraling updrafts) in
  2     arid regions of southwestern United States have comparable contribution to dust
  3     re-suspension as wind-driven soil erosion.
  4
  5     6.4.5.1  Non-urban Size and Chemical Composition in the Southwest
  6           The PM10 concentrations at non-urban southwestern sites show a double peak,  one
  7     during the late spring (April through July), and another in October.  This bimodal seasonality
  8     is imposed by the coarse particle mode. The PM2 5 mass concentration is unimodal with a
  9     summer peak.  Overall, the non-urban PM10 concentrations are comparatively low (8 to
 10     15 pig/m3) and over 60% contributed by coarse particles (Figure 6-45b).
 11           The chemical mass balance (Figure 6-45c) shows sulfates to be the main contributors
 12     during the winter (December through March) as well as in late summer (July through
 13     October).  However, sulfate and organics contributions are comparable during March through
 14     June as well as during November through December.   Fine particle soil plays a prominent
 15     role in the spring fine particle  chemical mass balance reaching 25%. However,  the role of
 16     fine particle soil dust during December through February dwindles to below 10%.
 17           The selenium and vanadium trace elements concentrations (Figure 6-45d) are very low
 18     and rather invariant throughout the year.  The fine particle sulfur concentration is low and
 19     exhibits a peak during August, which is the period of the summertime monsoon, when air
 20     masses of Gulf of Mexico penetrate deep into the southwestern US.  The S/Se ratio is
 21      comparatively low and bimodal, with peaks in April through May as well as August through
 22      October.
 23
 24      6.4.5.2 Urban Aerosols in the Southwest
 25            The downward PM10 trend of the Southwest is a remarkable 50% between 1985
 26      (52 /*g/m3) and 1993 (26 /xg/m3).  The decline was quite steady throughout the period.
 27      Another notable feature of the  Southwest is the large concentration spread of 45% among the
 28      monitoring sites (Figure 6-46). Sites with low concentrations (<20 ng/m3) occur adjacent to
29      high concentration sites (>50 /xg/m3).
30            Seasonally, the Southwest PM10 concentration shows two peaks, one in late spring
31      April through June, and another during the fall  October through November. The

        April 1995                               6.79      DRAFT-DO NOT QUOTE OR CITE

-------
 1     concentration dip in August and September has not been observed for any other region. The
 2     late summer concentration drop coincides with the occurrence of the moist monsoonal air
 3     flows from the Gulf of Mexico.  It is also notable, that the rather high 1993 PM10
 4     concentration of 26 /ng/m3 is only exceeded by the southern California region.
 5           The size segregated aerosol samples from the Southwest clearly show that coarse
 6     particles dominate the PM10 concentration, fine particles contributing only 37%.   The scatter
 7     chart also reveals that high PM10 concentration months occur without the presence of fine
 8     particles.  Itis evident, therefore, that in the Southwest natural and man-induced coarse
 9     particle dust is the dominant contributor to PM10 aerosols.
10           The short term PM10 concentration over the Southwest (Figure 6-47) exhibits a highly
11     irregular pattern, that ranges between 12 to 52 /xg/m3 regional average for any given day.
12     Both the lowest (10 to 15 /xg/m3) as well as the highest values are dispersed throughout the
13     year. The seasonality is virtually indiscernible, being  much smaller than the  sixth day to
14     sixth day variation. It is worth noting (Figure 6-46b) that the southwestern region has rather
15     low  logarithmic  standard deviation compared to other regions.
16           The urban excess PM10 (AIRS-IMPROVE) for the Southwest is given  in Figure 6-48.
17
18     6.4.6  Regional Aerosol  Pattern  in the Northwest
19           The Northwest is defined to cover the bulk of the western United States north of the
20     Arizona border (Figure 6-49a). It is covered  by  mountainous terrain of the Rockies, as well
21     as the Sierra-Cascade mountain ranges.  It is clear that the Northwest is actually a collection
22     of many aerosol subregions.  The meteorology is highly variable between the Pacific
23     Northwest and the Rocky Mountains with prevailing winds generally from the west.  The
24     main feature of the Northwest is pronounced elevation ranges between mountain tops and
25     valleys, and the resulting consequences on emission pattern (confined to the valleys)  and
26     limited ventilation. The  mountainous Northwest has also industrial population centers, such
27     as Seattle, Portland,  Salt Lake City and Denver.
28            Examining the carbonaceous particles and regional haze in the western and
29     northwestern US, White and  Macias, 1989 concluded  that in the rural areas the
30     concentrations of paniculate carbon are comparable to those of sulfate. Examining
31     paniculate nitrate (White and Macias, 1987) showed that the paniculate nitrate concentration

       April 1995                               6-80      DRAFT-DO NOT QUOTE OR CITE

-------
                     Average AIRS PM10 -1985-1993
                                                         AIRS PM10 Concentration Trends
                                                                     Southwest
                                                         1985 1986 1987  1988 1989 1990 1991 1992


                                                        PM10 AVO    "I -Q. PM10AVO-SIO  -+ PM10AVG + SIC
           AIRS PM2.5 vs. PM10 - Monthly Avgerages
                         Southfetst
            150

            140 -
             6

             65

             60

             45

             40

             35

             30

            25

            20

            15

            10

             6
              AIRS PM10, PM2.5 and PM Coarse Cone.
                         Southwest
                                                                       -Q— i
                     30     60      90

                  PM10_AVG UGTCU METER (25 C)
                                         120
            1886   w»r Apr May Jun Jul Aug Sep Oct Nov D«c

                    _J -B- Fin*     -4- Gears*
Figure 6-46.       AIRS concentration data for the Southwest, (a) Monitoring locations.
                    (b) PM10, PM2 5, and PMCoarse.  (c) Chemical fraction of sulfate,
                    soil, organics, and soot,  (d) Tracer concentrations.
April 1995
6-81
DRAFT-DO NOT QUOTE OR CITE

-------
             o
             m
             e^

             oT
          o 111

          i&
          Q- 5
             D
             O
             O
                                Southwest
                             Every Sixth  Day
                            1991
                                     1992
                1993
Figure 6-47.     Short term variation of PM10 average for the Southwest. Data are
               reported every sixth day.
                        Southwest urban excess
                CO


                "5)
                3
                   40
                   35
                   30 <
                   25
                   20
                   15
H	1-
                     Jan   Mar   May    Jul    Sep   Nov



Figure 6-48.      Urban excess concentration (AIRS minus IMPROVE) for the

                Southwest.
April 1995
     6-82
DRAFT-DO NOT QUOTE OR CITE

-------
             PM2.5 - Nonurban PM Monitoring Sites
                                                    PM10, PM2.5 and PMC Monthly Avg.

                                                      Northwest - IMPROVE/NESCAUM Networks
                                                     60,000,	,—-,	,	,	,	     .
                                                        1989 Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec


                                                  •&PM10      .; H-  PM2.5       -i- PM Coarse
                     Northwest
                                                                  Northwest
    c:
    •6
0.9


0.8


0.7


0.6


0.5


0.4


0.3


02


0.1


0.0
                     -
                  13 —-— e-
            1989 Feb Mai Apr May Jun Jul Aug Sep del  Nay Dec


         -&- Sulfate           -e Organics

         -+ Soil            !'-5-' Soot          '

         -r> Sulf+Org+SolH-Soot
a,wu
3,500
3,000
2,500
2,000
1,500
1,000
500
n

-
-

-

-
, -*--' - -
^^^-^^-^^^^-^
                                               Scale
                                             1389 Peb Uar Apr  May Jun Jul Aug Sep Oct  Nov


                                          ts- Sulfur   -3- Selenium  -+• Vanadium -Q- S/Se

                                           0-4000      0-4       0-10      0-tOOO
Figure 6-49.       IMPROVE/NESCAUM concentration data for the Northwest.
                    (a) Monitoring locations, (b) PM10, PM2-5, and PMCoarse.
                    (c) Chemical fraction of sulfate, soil, organics, and soot,  (d) Tracer
                    concentrations.
April 1995
                                 6-83
DRAFT-DO NOT QUOTE OR CITE

-------
 1     in the northern states, (MT, ID, WY) were substantially higher, both in absolute terms and
 2     relative to sulfate concentrations.  Aerosol particulate nitrates over rural mountainous West
 3     were also episodic,  i.e. few samples contributed a large fraction of the fine particle
 4     integrated dosage.
 5
 6     6.4.6.1  Non-urban Size and Chemical Composition in the Northwest
 7           The non-urban PM10 concentrations show low values ranging between 7 to 14 /^g/m3
 8     in the northwestern US.  The seasonally shows a peak in the summer which is contributed
 9     by both fine and coarse particles.  Coarse particles account for more than half fine mass,
10     particularly during March through June spring season (Figure 6-49b).
11           The chemical mass balance (Figure 6-49c) shows roughly comparable contributions
12     from sulfates and organics, but their  seasonality is phase shifted. Sulfates prevail during the
13     spring season  while organics dominate during late fall (October through January).  Fine
14     particle soil dust contributes 20%  during April and May, but decline well below 10% during
15     the winter months (November through February).   Overall, about 80% of the fine mass is
16     accounted for  by  the sulfates organics, soil  and soot.
17           Both  selenium and vanadium concentrations  (Figure 6-49d) are low in the Northwest,
18     but there is  an indication of a summer peak of Se.  The S/Se ratio is between 500 to 1,000,
19     which is the lowest among the regions.  It is interesting that this ratio has both spring peak
20     as well as fall peak, similar to the pattern observed for the southwestern United States.
21
22     6.4.6.2  Urban Aerosols in the Northwest
23            The time trend of aggregated PM10 concentrations conveys a significant decline of
24     50% from  1985 to  1993  (from 50 to 25 /xg/m3).  However, from 1986 to 1993 the decline
25     was only 37%, which is  believed to be more  representative for the region because of the low
26     station density in the early years.  Once again, the average 1993 concentration is 25 pig/m3
27     which is comparable to the 1993 concentrations of the eastern U.S. regions.  The spread of
28     concentration  among the Northwest stations is large, with standard deviation of
29     45%  (Figure 6-50b). This spread in the concentration values  is also evident from the various
30     circle sizes  of the Northwest map. It is believed but  not formally confirmed that low
31     concentrations occur primarily at high elevation sites that are  above  the daily reach of surface

       April 1995                                6-84      DRAFT-DO NOT QUOTE OR CITE

-------
  1     based haze and smoke layers.  Inspection of the circles in the northwestern map also reveals
  2     that the highest PM10concentrations in the Northwest occur at more remote mountainous
  3     valleys, rather than in the center of urban-industrial areas.
  4           The seasonality of the northwestern United States has an amplitude of 36% which is
  5     comparable to the strong seasonality of the eastern  US. However, the shape of the seasonal
  6     signal is phase shifted to a winter peak.  The lowest concentration occurs during March
  7     through May and gradually increases to a peak in December through January, falling sharply
  8     between January and March.
  9           The limited PM2 5-PM10 data for the  Northwest indicate that on the average 59% of
 10     PM10 particles are PM2 5. The scatter chart also indicates that the extreme PM10
 11     concentrations are contributed mainly by fine particles.  Furthermore, the extreme PM10
 12     concentration also occur in the winter season.
 13           The daily concentration when averaged over  the large and heterogeneous northwestern
 14     region exhibits a remarkably small sixth day to sixth day variation (Figure 6-51).
 15     Furthermore, there is clear seasonality with  a strong winter peak.  Within a given season, the
 16     regionally averaged  concentrations only vary by 20  to 40% from one sixth day to another.
 17     This low temporal variance would indicate homogeneity of aerosol over the region.
 18     However,  examination of the logarithmic standard deviation (Figure 6.3.12) shows that the
 19     Northwest is spatially more heterogeneous and has the highest logarithmic standard deviation
 20     among all regions.   This apparent discrepancy shows the advantages of aggregations
 21      performed on multiple spatial and temporal scales.  Evidently, in the  Northwest high
 22     concentration PM10 pockets, in topographically  confined airsheds result in strong spatial and
 23      temporal variations.  However, the sensory evidence suggests that large scale elevated PM10
 24      concentrations that cover the entire Northwestern region do not exist because high
 25      concentrations are not "synchronized" between  the different airsheds.  In this sense, the
 26      Northwest differs markedly from  the eastern US,  where large regional scale airmasses with
27      elevated PM10 determine the regionally averaged values.  The urban excess PM10 (AIRS-
28      IMPROVE) for the Northwest is given in Figure 6-52.
       April 1995                               6-85       DRAFT-DO NOT QUOTE OR CITE

-------
               Average AJRS PM10 • 1985-1993
       AIRS PM10 Concentration Trends

                   Northwest
       60,	p	•	•	.	1	.	r
      AIRS PM2.5 vs. PM10 -Monthly Avgerages
                    Northwest
                                                   1986  1985  1987  1988 1989 1990 1991  1992
                                              \-tr PM10AVG    i -Q- PM10AVO-SIO  -+ PM10AVG + SIC
          AIRS PM10, PM2.5 and PM Coarse Cone.
                    Northwest
                                               u
                                               &
          0      30     60     90

              PM10JWG UGJCU METER {26 C)
          1S8«   M»r Apr M»y Jun Jul AUJ S«p 0«t Ncv Dee


          \-ts- PM10   I -B- Fin*     -+• Coars*
Figure 6-50.       AIRS concentration data for the Northwest, (a) Monitoring locations.
                   (b) PM10, PM2 5, and PMCoarse.  (c) Chemical fraction of sulfate,
                   soil, organics, and soot,  (d) Tracer concentrations.
April 1995
6-86
DRAFT-DO NOT QUOTE OR CITE

-------
             o
          O UJ
          IS
          OL 5
             D
             O
             O
                               Northwest
                            Every Sixth Day
                           1991
 1992
 1993
Figure 6-51.      Short term variation of PM10 average for the Northwest.  Data are
               reported every sixth day.
                       Northwest urban excess
              "5)
               » 20
              O
              i
              a.
                   Jan    Mar   May   Jul    Sep   Nov


Figure 6-52.      Urban excess concentration (AIRS minus IMPROVE) for the
               Northwest.
April 1995
6-87
DRAFT-DO NOT QUOTE OR CITE

-------
 1     6.4.7  Regional Aerosol Pattern in the Southern California
 2           The region covers California, south of San Francisco Bay (Figure 6-53a).  It was
 3     declared as a separate region primarily because of the known high aerosol concentrations in
 4     the Los Angeles and San Joaquin basins.  Meteorologically the region is exposed to the air
 5     flows from the Pacific that provide the main regional ventilation toward the south and
 6     southeast.  The precipitation in the region occurs in the winter season,  with the summer
 7     being hot and dry. The regional ventilation of the San Joaquin Valley is severely restricted
 8     by Sierra Mountain ranges.  Also, the San Gabriel  Mountains constitute an air flow barrier
 9     east of the Los Angeles basin. Both basins have high population, as well as industrial and
10     agricultural activities. Hence, human activities are believed to be the main aerosol sources
11     of the region.
12
13     6.4.7.1  Non-urban Size and Chemical Composition in the Southern California
14           The PM10 concentration at the non-urban sites over southern California ranges
15     between 10 /ig/m3 during December through February, and 20 to 25 /*g/m3 in April through
16     October.  Coarse particles contribute more than 50% of PM10 during the warm season May
17     through October. Both the fine and coarse aerosol fractions are lowest during the winter
18     months (December through March).  The summer peak fine particle seasonality at non-urban
19     southern California sites is in marked contrast to the strongly winter peaked urban fine
20     particle concentrations (Figure 6-53b).
21           The chemical mass balance (Figure 6-53c) of non-urban southern California aerosol is
22     clearly dominated by organics, which contribute 30 to 40%  throughout the year.  Sulfates
23     account for only 10 to 15% of the fine mass in the winter, and about 20% in the summer
24     months.  The sulfate  fraction of the non-urban southern California fine  mass is the lowest
25     among the regions.   Fine particle soil  dust is about 10% between April through November
26     and drops to 5% during the winter months. A notable feature  of the southern California
27     chemical mass balance is that 45% of the winter, and 35% of the summer fine mass
28     concentration is not accounted by sulfates, organics, soils, and soot. Undoubtedly nitrates
29     are the major contributors to the southern California aerosols.
30           Both selenium and vanadium (Figure 6-53d) show low values throughout the year
31     without significant seasonality.  On the other hand  the fine particle  sulfur concentration

       April 1995                               6-88      DRAFT-DO NOT QUOTE OR CITE

-------
             PM2.5 . Nonurban PM Monitoring Sites
       PM10, PM2.5 and PMC Monthly Avg.
        Northwest - IMPROVE/NESCAUM Networks
        60,000,	.	.	.	.	.	.	.	,	,	,	,
                                                    55,000

                                                    50,000

                                                    45,000

                                                    40,000

                                                    35,000

                                                    30,000

                                                    25,000

                                                    20,000

                                                    16,000

                                                    10,000

                                                     5,000
                                        •*---
                                                       1989 Feb Mar Apr May Jun Jul Aug Sep Od Nov Dec

                                                 :~iEHPM?0; -4- PM2.5       -A  PMCoQrse
                    S. California
                                                                 S. California





vt


Uu

g
1
LL






4000, 	 r- 	 , 	 r— -, 	 . 	 . 	 . 	 . 	 • 	 • 	 . 	 1
OS
0.8

0.7


0.6


0.5


0.4

0.3

02
0.1

-
-

•'''""" ,-i*-
f- ""• '''' "" •*"•
~" "^ *- -
^ ''


-


"-• r^'B^Q~L
'TZL £3r~ ~~^~^ ~l-J^-~ __ ^.Q 'U
A.
^ ^
— &^~^ ^^\- _^A~~^~^
^, ft\ 	 _f* — *>" ^-j
fc...-^--~*- o- , 4,. -i- 0- - -S- -S- •'
3,500

3,000


2,500


n
I 2,000
c

1,500


1,000

600

0.0 ' 	 ' 	 ~ 	
*««.* i- t u_- A 	 ••_.. !..» lul Ann Can Art Nnw Hpr n
-

-


_



-




.-, - '" ""

N
.-•'" .,,---1^-*^^-^,^^
'^Sr^ ^ ^
.
          4- Soil            l-o- Soot
          -o- Sulf+Org+SoiH-Soot
            1989 feb Mar Apr May Jun Jul Aug Sep Ocl Nov Dec



   Scale    0-4000      0^       0-10     0-4000
Figure 6-53.        IMPROVE/NESCAUM concentration for the Southern California, (a)
                    Monitoring locations, (b) PM10, PM2 5, and PMCoarse.   (c) Chemical
                    fraction of sulfate, soil, organics, and soot,  (d) Tracer
                    concentrations.
April 1995
6-89
DRAFT-DO NOT QUOTE OR CITE

-------
 1     shows a definite summerpeak at 500 ng/m3, compared to 200 ng/m3 during the winter.
 2     Consequently, the S/Se ratio increase from 500 in the winter 1,000 to 1,500 in the summer.
 3
 4     6.4.7.2  Urban Aerosols in the Southern California
 5           The downward concentration trend of southern California between 1985 and 1993 was
 6     41%. However, between 1986 and 1993 the reduction was 27%.  By 1993, the southern
 7     California annual average PM10 concentration was reduced to 32 ^g/m3.  However, there is a
 8     sizable concentration spread among the stations (40% standard deviation). Inspection of the
 9     circle sizes in the map points to uniformly high concentrations in the San Joaquin Valley as
10     well as in the Los Angeles basin.  The low concentration sites are located either on the
11     Pacific coast or in the Sierra Mountains.  Thus there are clear patterns of basin-wide elevated
12     PM10 concentrations and lower values in more remote areas (Figure 6-54b).
13           The seasonality of the PM10 pattern in southern California is significant at 27%.
14     Furthermore the seasonal pattern is unique that the highest concentrations occur in
15     November, while the lowest in March.  However, it is a see saw rather than a sinusoidal
16     pattern.
17           On the average, about half of southern California PM10 is contributed by fine particles
18     as shown in PM2 5-PM10 scattergram.  Most of the high PM10 concentration months are
19     dominated by fine particles  and tend to occur in the fall, season.
20           The sixth day average time series for the southern California region (Figure 6-55)
21     shows remarkably high sixth daily variance, between 12 and 75 /ig/m3.   The lowest values
22     tend to occur between January and April, while the highest concentrations (>50 /ig/m3) tend
23     to occur during October through December.  Concentration excursions factor of two are
24     common between two consecutive six day time periods. However, visual inspection of the
25     sixth daily signal also reveals a substantial seasonality that peaks in the fall September
26     through December and lowest in the spring.
27           The urban excess PM10 (AIRS-IMPROVE) for Southern California is given in
28     Figure 6-56.
        April 1995                               6-90      DRAFT-DO NOT QUOTE OR CITE

-------
                  Average AIRS PM10 -1985-1993
          AIRS PM10 Concentration Trends
                      S. California
          60
                                                       1985  1986 1987 1988  1909 1990 1991 1992


                                                                -Q- PM10 AVO - SIO  -+- PM10 MG + SIO
        AIRS PM2.5 vs. PM10 - Monthly Avgerages
                     S. California

             CO&RELATiOH'
            AIRS PM10, PM2.5 and PM Coarse Cone.
                       S. California
             Avg X
                Y
             Avg Y/Avg X
             Corr Coeff
             Slope
             Y offset
             Data Points
             0      30      60      90     120

                 PM10JVVG UGJCU METER (25 C)
                                                  I
           1986   Mar Apr Mjy Jun Jul Aug S«p Oct Nov

           I -4- PM10   I -Q- Fine     -+• Coarse
                                                                                       D«c
Figure 6-54.        AIRS concentration for the Southern California, (a) Monitoring
                    locations, (b) PM10, PM2 5, and PMCoarse.  (c) Chemical fraction of
                    sulfate, soil, organics, and soot,  (d) Tracer concentrations.
April 1995
6-91
DRAFT-DO NOT QUOTE OR CITE

-------
               o
               m
            O UJ
            it
            o- 5
               ^
               O
               o
                                         Southern
                                         California
                                        Every Sixth
                                            Day
                             1991
   1992
  1993
Figure 6-55.      Short term variation of PM10 average for the Southern California.
                Data are reported every sixth day.

                        S. California urban excess
                    10
                      Jan   Mar   May   Jul    Sep   Nov
Figure 6-56.      Urban excess concentration (AIRS minus IMPROVE) for the
                Southern California.
April 1995
6-92
DRAFT-DO NOT QUOTE OR CITE

-------
  1     6.5  SUB-REGIONAL AEROSOL PATTERNS AND TRENDS
  2           The health and other effects of aerosols are imposed on individuals, and the density of
  3     population varies greatly in space.  Consequently, the evaluation of effects requires the
  4     knowledge of aerosol concentrations over specific locations where sensitive receptors reside.
  5     The purpose of this section is to characterize the aerosol pattern at specific sites, small
  6     airsheds or sub-regions.  The discussions is  organized by region and then by monitoring site
  7     within a region. Most urban aerosol sampling is confined to PM10 or in some instances to
  8     PM2 5 and PMCoarse.  However,  detailed chemical composition data are reviewed for
  9     several urban areas.
 10
 11     6.5.1  Sub-regional Aerosol  Pattern  in  the  Northeast
 12           In the northeastern region, the Shenandoah National Park and Washington, DC
 13     constitute a natural urban-non-urban pair of  size and chemically resolved aerosol data.  New
 14     York City and Philadelphia are also major metropolitan areas with substantial aerosol data.
 15     Whiteface Mountain site distinguishes itself from its background by high elevation.
 16
 17     6.5.1.1  Shenandoah National Park
 18          The PM10 concentration at the Shenandoah National Park IMPROVE site
 19     (Figure 6-57a) exhibits a pronounced summer peak (27 /zg/m3), which is factor of three
 20     higher than the low winter value of 9 /xg/m3. The strong seasonality is driven by the
 21      seasonal modulation of the fine mass which accounts for 70 to 80% of PM10  mass
 22      (Figure 6-57a). The coarse particle concentration ranges between 3 to 6 /ig/m3, which is
 23      small compared to the fine particle mass, particularly in the summer season, when it  accounts
 24      to <  25% of the PM10.  It is clear that at this remote  site, in the vicinity of industrial source
 25      regions, fine  particles determine the magnitude of PM10.
 26           The chemical mass balance for the Shenandoah IMPROVE monitoring site (Figure
 27      6-57b) clearly documents the dominance of sulfate aerosols which contributes about 60% of
 28      the fine mass during April through September, and about 50% during the winter months.
29      Organics, on  the other hand,  range from 20% in summer to 30% in the winter months.  The
30      contribution of fine particle soil and soot is well below 5%.  Throughout the year about 90%
31      of the fine mass is accounted for by these measured substances.

        April 1995                               6.93      DRAFT-DO NOT QUOTE OR CITE

-------
u.
•8
|

1
                           PM10, PM2.5 and PMC Monthly Avg.
                                      Shenandoah NP
                            60,000
                            65,000

                            60,000

                            45,000

                            40,000

                            35,000

                            30,000

                            25,000

                            20,000

                            16,000

                            10,000

                             6,000
                               1969   Mar   May   Jul    S«p   Nov

                           -B- PM10      H-- PM2S     I-A- PMCoare*  I
                Shenandoah NP
                                                              Shenandoah NP
         OJ

         o^

         0.7

         0.6

         0.6

         0.4

         03

         02

         0.1

         0.0
          I--&-Q
                                 B'
                                   -Q.
                                                    4,000
                                                 3,600
                                                 3,000
2,000


1,600


1,000


 600
           1989 Fed Mar Apr M*y Jan M Aug Sap Od Hov Dec

       -A- Sulfato          -& Orfltnlcs
                                                    1S89 F«b Hw Apr Miy J« M  A«| S*p Oct No* 0*c
       .+. SOD            1-0- SOOt

       -cv- SulfKJro+SoiHSoot
                                     J
                                         Scale
0-4000
       -Q- 6*)«flhjRi  --»- V«nt5, and PMCoarse.
                    (c) Chemical fraction of sulfate, soil, organics, and soot,  (d) Tracer
                    concentrations.
April 1995
                                         6-94
   DRAFT-DO NOT QUOTE OR CITE

-------
  1            Chemical Tracer data is shown in Figure 6-57rc. The concentration of coal-tracer
  2      selenium shows a mild double peaks, one during December through March, and another in
  3      June through September.  Vanadium is relatively constant throughout the year. The fine
  4      particle sulfur concentration is almost factor of five higher in August (3,300 ng/m3) than the
  5      December values(700 ng/m3).  This extreme sulfur seasonality  is unique to the Shenandoah
  6      site.  The S/Se ratio has a remarkably smooth but highly  seasonal variation that varies by
  7      about factor of four between the winter  (700) and summer (2,600) values.  If  Se-bearing coal
  8      combustion was the exclusive source of sulfur at the Shenandoah National Park ,  then the
  9      sulfate yield per emission would be 3 to 4 higher in the summer than in the winter.
10            An examination of the nature and sources of haze in the Shenandoah Valley/Blue
11      Ridge Mountains area (Ferman et al., 1981) showed that  sulfate aerosols were the most
12      important visibility reducing species.  Averaging 55% of  the fine particle mass, sulfates (and
13      associated  water) accounted for 78% of the total light extinction.  The second  most abundant
14      fine particles, accounting for 29%  of the fine mass, was organic carbon.  The remaining
15      particle mass and extinction were due to crustal materials.
16            Using an in-situ rapid response measurement of H2SO4/(NH4)SO4 aerosol in
17      Shenandoah NP,  VA, Weiss et al, 1982, found that the summer sulfate and ammonium ions
18      average 58% of particle mass smaller than 1 mm.  The particle composition in terms of
19      NH4+/SO42" molar ratio ranged from 0.5 to  2.0 with strong diurnal  variation.   The particles
20      were most acidic at 1500 EOT and least acidic in the period 0600 to 0900 EDT.  The water
21      contained in ambient aerosol particles was more strongly  associated with sulfate and
22      ammonium ions than with the remainder of fine particle mass.
23
24      6.5.1.2 Washington, DC
25            The PM10 concentration at Washington DC (at the  top of the National Park Service
26      Headquarters building) is virtually  constant over the seasons at  25 to 30 /*g/m3.  Fine
27      particles contribute over 70% of PM10 throughout the year (Figure 6-58a). The lack of
28      seasonality in the fine particle mass is in sharp contrast to the factor of three seasonal fine
29      mass modulation at the Shenandoah National Park.   The coarse particle concentration in
30      Washington, DC is 8 to 10 ^g/m3 throughout the year, exhibiting  virtually ho  seasonality. In
       April 1995                                6-95       DRAFT-DO NOT QUOTE OR CITE

-------
 1      fact, compared to Shenandoah National Park, the excess coarse particle concentration in
 2      urban Washington, DC is only 5 j*g/m3 throughout the year.
 3           The chemical mass balance at the urban Washington, DC site (figure 6-5 8b) is
 4      dominated by sulfates during the summer months (50%  of fine mass), which declines to 30%
 5      in January.  Organics, on the other hand, are more important during October through
 6      January (40%)  but only 30% during May through August.  Thus, the relative roles of
 7      organics and sulfates at the Washington, DC urban site  is fully phase shifted by half a year.
 8      Soot is a substantial (9 to 12%) contributor to the fine particle mass, particularly during
 9      October through December. Fine particle soil contributes a remarkably low 2 to 5% of the
10      fine particle mass at this urban site.  Hence, neither coarse particle, nor fine particle dust is
11      an important factor in Washington, DC.
12           The chemical tracer species are shown in Figure  6-58c.  The coal tracer, selenium
13      concentration ranges between 1.5 to 2.0 ng/m3 without appreciable seasonality.  The urban
14      Se is also higher than the Se at any of the non-urban site,  including the industrialized
15      Midwest.  Vanadium, the tracer for fuel oil, varies by factor of two between the high winter
16      values (> 8 ng/m3) and low summer values (3 ng/m3).  The pronounced V concentration
17      seasonality is a clear indication of that the emissions from fuel oil and other vanadium
18      sources are seasonal. The fine particle sulfur concentration varies by about factor of two
19      between 1,400  ng/m3  winter concentration, and about 3,000 ng/m3 summer peak. The
20      seasonal modulation of sulfur in Washington, DC is only factor of two compared to the
21      factor of four fine sulfur modulation at Shenandoah National Park.  The difference is
22      primarily due to the elevated winter  sulfur in Washington, DC. The S/Se ratio is about 0.6
23      in the winter and about 1.5 in the summer. It differs from Shenandoah by the lower summer
24      S/Se ratios.
25
26      6.5.1.3  Comparison of Non-urban (Shenandoah) to Urban (Washington, DC) Aerosols
27           The Washington, DC urban site and the companion non-urban Shenandoah monitoring
28      site constitute a rare data pair that allows the quantification of urban-rural differences in fine
29      and coarse particle concentration, and chemical composition.  Within Washington, DC,
30      industrial emissions are moderate compared to the industrial midwestern cities.  However,
31      both automobile emission density and emissions from winter time heating are expected to  be

        April 1995                                6-96       DRAFT-DO NOT QUOTE OR CITE

-------
PM10, PM2.5 and PMC Monthly Avg.
60,000
65,000
60,000
45,000
40,000
35,000
J 30,000
c
25,000
20,000

0.9
0.8
0.7
tn
Fraction of Fine Mi
o o o o o
to b !>. bi b»
0.1
0.0
19
-Ar SM
-+• Soil
15,000
10,000
5,000
0
19
-B- PM10
Washington DC
Washington DC
-
-
-
-

ja-e" H\_
J3— S¥^ '^St ^V
i— e— B— s-^1 Nj^/N
: _^ x^^^v ,J
-,.-*---.
^
i










89 Mar May Jul Sep Nov
-(-- PM25 i-A- PM Coarse |

a t> '*' ' ~~ w
.- '%'" '^— *-'
/ ^


*r *^ ^Q
^< /^X*
/^"^B-e X ^^
•0"
t ^ > - -o. ^ _4
j,--'"*"— «f.".I^~— 4--.".'°' 4,
39 F«b Hw Apt Hay Jim Jul Aug Sep Oct Hov Dec
ate -B- Organic:
)-o- Soot |
4,000
3,600
3,000

2,600
c
1,000
500
0
m
-frSulft«
Washington DC

-
-'••• --f A f'
/ 15 \ /
/ * \ * •' '"k
_B */P;"' \ •' "• ; ^
^ \ y / v' \ ,°A-e-v' ~
v^-/\ / / ^\^// \ ^
>r • ' /• <& / \
- >r x -S" ;•' \ 'V \.
J-~ - i'v
\>
:-* "^^.:
19 Felt Mar Apr
r -a- Selenl
May Jun Jul Aug Sep Oa HOY Dec
urn -•)- Vanadium | -e- SISt }
    -0- SulffOrg+SoBfrSoot
                                     Scale
                                            0-4000
                                                     0-4
                                                             0-10
                               0-4000
Figure 6-58.      BMPROVE/NESCAUM concentration for Washington, D.C. (a)
                 Monitoring locations, (b) PM10, PM2i5, and PMCoarse. (c) Chemical
                 fraction of sulfate, soil, organics, and soot,  (d) Tracer
                 concentrations.
April 1995
6-97
DRAFT-DO NOT QUOTE OR CITE

-------
 1      high. In this section the excess aerosol concentrations in Washington, DC over the
 2      Shenandoah site are examined to elucidate the urban influence.
 3            The Washington, DC excess PM10 concentration (Figure 6-59a) ranges between 15-20
 4      Mg/m3 in the winter, and  <3 ^cg/rn3 in the summer.  Hence, there is almost an order of
 5      magnitude higher urban excess during the winter, compared to the summer.  The seasonality
 6      of the excess PM10 is driven by the winter peak excess fine particle concentration of 10-12
 7      Mg/m3-  The excess coarse particles hover in the 3 to 6 pig/m3 range throughout the year.
 8      Thus, the urban Washington, DC concentration exceeds its non-urban regional aerosol values
 9      during the winter season,  and the excess winter time urban  aerosol is largely contributed by
10      fine particle mass. This is remarkable, and indicates the diminishing role of coarse particle
11      fly ash, road dust resuspended by automobile or construction, road salt and all other sources
12      of urban coarse particles.
13            The chemical composition of the excess fine particle  concentration over the
14      Shenandoah non-urban background is also shown in Figure  6-59b.  Fine particle organics
15      dominate the urban excess chemical mass balance,  ranging between  1 ^g/m3 during the
16      summer, and 5.5 jug/m3 during the winter.  The seasonality of excess organics also drives the
17      seasonality of excess fine  mass.  There is an excess sulfate concentration of 1 to 2 /ug/m3 in
18      Washington, DC, except during July, August, and September.  In fact, in August in
19      Washington, DC sulfate concentration is about 0.3 /xg/m3 below the Shenandoah values.  The
20      urban excess soot concentration is  1 to 2 /ig/m3 throughout  the year.  The  soil contribution to
21      the fine particle mass is identical to the values of the Shenandoah National Park,  yielding
22      virtually no excess fine soil contribution in the urban area.  It is worth noting, however, that
23      there is a modest excess coarse concentration of 2 to 5 /*g/m3 in Washington, DC
24      (Figure 6-59).  This indicates that  soil dust aerosol components that make up the coarse mass
25      are large particles without appreciable mass below 2.5 /mi.
26            The short-term fine mass concentration at Washington, DC and  Shenandoah National
27      Park for the year 1992 is  shown in Figure 6-60. Although the sampling is conducted
28      Wednesdays and Saturdays for 24-hours, the data points are connected as a guide to the eye.
29      The figure also compares  the daily (Wednesdays and Saturdays) fine particle  sulfur
30      concentrations at the two monitoring sites.  The fine mass concentration time series shows
31      elevated concentrations (>30 /xg/m3) that occur throughout the year. On the  other hand,

        April 1995                                6-98      DRAFT-DO NOT QUOTE OR CITE

-------


fl>
o
1
1
Lai
o.
z
«
o
c
ro
c
«
£
CO
1
U
o
c
o
ri
O)
"?-
— i 	 1 	 : 	 1 I i i TV i 	 na — i — *n 	
' X I
n 4 '• -
* q V -
v n ii -i
'"--. \ '/^
* - ^ p-j \ j
- ^;. |
\& -
' \
f 'ffl'
* ' ^-f
'* ' ' !
' , '
£- 0 < ..• -
-' / ' ('
<> ffl < f .
/ ' [ / ">
* m «/•••'' -
/ f
/ ? fH
1 1 1 1 1 1 1 l.-x" 1 In,' 1 -JJ I
ooooooooooooc
oooooooooooo
oooooooooooo

U
0)
0
g
z
ts
o
g^
U
M
B)
3
_
3
c
3
"5
S1


a.
<
b.
(0

^
0)
LL
S
> 5
(/)

I
                                                    M
                                                    to  *-

                                                    ?  §
                                                    O  CO

                                                    q:  9
                                                   M

                                                   M

                                                   (0
                                                         r
                                £


                                3  "5



                                -h  4-
                                                                 I
                                                          I
                                                           «
                                                           o
                                                          "O
                                                           c
                                                           «
                                                           c
                                                           a>
                                                                   -a
                                                                   c
                                                                 •a s
                                                                 a>  o
                                                                 S  "3
                                                                 «  ^



                                                                 il
                                                                 «  So

                                                                 U  fe

                                                                 P  ,T
 O)

 o

 I
        S
        10
           0>
           U

           C

           %

           I
           o
           O

           c
U)
                 10
                 10
o  o
o  o
o  o.
o
\o
                 .
                10
                ^
   §o
   o
o  o

8  8
                                    o
                                    o
                                    o
                                                                 .5
                                                                 J3
                                                         18!
                                                   a.


                                                   t
                                                          a.
                                                             g
                                                          *j PS
                                                          t« ^
                                                          W ^j


                                                          II
                                              C ^
                                              o
                                              o ^



                                              11
                                              « U
                                                             ej
April 1995
                           6-99
                      DRAFT-DO NOT QUOTE OR CITE

-------
 1      high fine mass levels at Shenandoah are recorded only during the summer season.
 2      Paniculate sulfur concentration at the urban and non-urban site are comparable and co-vary
 3      throughout the year.  This indicates that paniculate sulfur is part of the regional airmass that
 4      at any given day influences both Washington, DC and Shenandoah.  Fine particle mass, on
 5      the other hand, shows and excess concentration at Washington, DC, particularly during the
 6      winter months.  Visual inspection of the fine mass daily time series  clearly indicates that the
 7      concentration change from one daily sample to another can be an order of magnitude
 8      different. Consequently, most of the concentration variance is due to random synoptic
 9      airmass changes, and to a lesser degree due to periodic seasonal  variations.  Further
10      concentration variance would exist if hourly data were available.
11
12      6.5.1.4  New York City, NY
13            The New York City metropolitan area is characterized by high population density,
14      moderate industrial activity, and relatively flat terrain. The PM10 concentration over the
15      metropolitan area is shown in Figure 6-6la. The circles  in the map show the locations of the
16      monitoring sites and the magnitude of each circle is proportional to the average PM10
17      concentration at that site using all available data. The observed average concentrations
18      change by about of factor of two to three from one location to another.  Higher average
19      concentrations tend to occur near the center of the metropolitan area.
20            The long-term trend of PM10 averaged over the New York City metropolitan area
21      (Figure 6-61b) shows a decline from about 35  /*g/m3 in 1986 to 25 jiig/m3 in 1992.  This
22      PM10 reduction is comparable to the reduction over the entire Northeastern region. The
23      average seasonal pattern over  the New York City metropolitan area  (Figure 6-61c) is 25 to
24      30 /ig/m3 throughout the year, but rises to about 40 /ig/m3 in July.
25            The seasonal pattern at three different individual monitoring sites in the New York
26      City metropolitan area is shown in Figure 6-62. The three sites  all  show similar seasonally
27      with a summer peak, but with elevated concentrations closer to the city center.
28            Size segregated aerosol samples in New York City (Figure 6-62e,f) show that at both
29      sites, PM10 concentrations are contributed primarily by fine particles.  Furthermore, the
30      seasonality at the Manhattan site shows two peaks, the summer peak and the winter peak.
31      Based on the discussion of the more extensive Washington, DC (Section 6.5.1.2)

        April 1995                                6-100      DRAFT-DO NOT QUOTE OR CITE

-------
                                                                    M
                                                                    re
                                                                    •
                                                                       •K
                                                                       re
                                                                       a.


                                                                       >a
                                                                       c
                                                                       o
                                                                       £.
                                                                        re
                                                                        o
                                                                       -

                                                             o £:
                                                                              13

                                                                              §
                                      •a
                                      o
                                      •a


                                      I
                                      O)
                                                                              T3
                                      O

                                      |D




                                      1
                                                                              I

                                                                              C/J

                             U  4J
                             o  re


                                §-c
                                re


                             i1  I
                             i-  re

                             in  a)
                                                                       •p
                                                                              V}
                                                                              en


                                                                              i


                                                                              c
                                                                              _o
                                                                              '•£

                                                                              2
                                                                              I

                                                                              >>
                                                                              ^
                                                         U)
                                  £tu/6u
                                                                              r,
April 1995
6-101      DRAFT-DO NOT QUOTE OR CITE

-------
                Average PM10-1985-1993
                             AIRS PM10 Concentrantion Trends
                                         New York
                              60,-	.	.	.	.	,	.	,_
                                                 1985 1986 1987  1988  1989 1990  1991  1992
                                                AIRS PW110 SEASONAL CONCENTRATION
                                                     Washington-New York Megalopolis
                                            P )r
                                            5H!
                              60

                              55

                              50

                              45

                              40

                              35

                              30

                              25

                              20

                              16

                              10

                              6
                                                 1985   Mar Apr May Jun Jul Aug 8«p Oet N«v D«c
Figure 6-61.
Aerosol concentration map, trend and seasonality in the New York
City region.
April 1995
                       6-102      DRAFT-DO NOT QUOTE OR CITE

-------
                                                          Mar   May   Jul   Sep   Nov
          NEW YORK CITY, NY - Manhattan
            £]- =PM10_AV360610010 NEWYORK CITY
            - =PM10_AV360610069 NEWYORK CITY

           NEW YORK CITY, NY
90
60
70
60
50
40
30
20
10
19
-dr
-^ 90
80
70
60
50
A
^A / v, -. 40
f" "S ** Q - J"
., ------ -'• "' -'- - - ,. -4___. 10
34 Mar May Jul Sep Nov 19
=PM10_AV360610056 NEW YORK CITYj [^r
-
-
-
-
' /A '
^A A^/a-QX\X^\ /PN
^ -B- -ef XQ^ B^ H
34 Mar May Jul Sep Nov
=PM10_AV 360610079 NEWYORK CfTYj
           -Q- =PM26_AV 360610056 NEW YORK CITY
            "  -PMC AVG36061005fiNPWYnRkrm
       -Q- "PM26JW360610079 NEW YORK CPTY
         =PMC AVR 3«ofiion7a NFw YORK rrn
Figure 6-62.       Fine, coarse and PM10 particle concentration near New York City.
April 1995
6-103      DRAFT-DO NOT QUOTE OR CITE

-------
 1      measurements it can be inferred that the summer peak in the fine mass is mainly due to the
 2      regional haze aerosol, while the winter peak is contributed by the local sources, confined to
 3      the inner metropolitan area.
 4            As part of the  New York Summer Aerosol study (Leaderer et al., 1978) continuous
 5      size monitoring confirmed the expected bimodal volume distribution with modes between
 6      0.1 to 1.0 /mi and over >3.0 /mi. A number  of interesting patterns were observed when the
 7      size distribution was averaged by hour of day. The diurnal average total number
 8      concentration showed a pattern which  corresponded closely with the normalized diurnal
 9      traffic pattern.  Particles  <0.1 /mi showed the most marked diurnal variation, following the
10      total number curve.  Moreover, particles in size ranges >0.1 /mi showed little variation in
11      the diurnal pattern. Analysis of samples processed by the diffusion battery indicated that
12      approximately 54%+ 18% of the sulfate  measured was in the suboptical range (approximately
13      0.04 /mi to 0.3  /mi) with the remainder above 0.3 /mi. Little sulfate mass was found in
14      particles in the nuclei range (<0.04 /mi).  Analysis of impactor samples for sulfates
15      consistently showed that more than 85%  of all water soluble  sulfates were  <2.0 jam in size.
16      Virtually  no nitrate was present in the nuclei size range while the suboptical size range
17      accounted for approximately 30%  of the total  nitrate.  70% of the total nitrate was found in
18      the size range >0.3 /mi.  Analysis of large stages of Anderson impactor  showed that
19      approximately 50% of particulate nitrate was greater than 5.5 jum in size.
20            Urban and rural particulate sulfur monitoring near New York City in the  summer
21      (Leaderer et al., 1982) indicated that sulfate concentration distributions were regionally
22      homogeneous and increased with increasing ozone levels and covariant with several other
23      pollutant and meteorological parameters.  Sulfate concentrations correlated  strongly with
24      ammonium and  strong acid at all sites.  No significant diurnal patterns for sulfate or
25      ammonium were seen at any site for low and high ozone levels.  Strong  acid concentrations
26      were highest at  the rural and semi-rural sites,  lowest at the urban sites, increased with
27      increasing ozone levels and exhibited diurnal patterns which matched the ozone diurnal
28      patterns.
29            Sulfate acidity  measurements (Waldman et al., 1991) at Chestnut Ridge, PA (east of
30      Pittsburgh) suggest that higher acidity  occurred in the overnight period (0000 to 0800) in the
31      late fall, while sulfate had its highest levels in the morning to afternoon period.  Size

        April 1995                               6-104      DRAFT-DO NOT  QUOTE OR CITE

-------
  1      dependent,  mass and composition of New York aerosol for low, medium, and high visibility
  2      levels was reported by Patterson and Wagman,  1977.  At all levels of visibility, bimodal or
  3      multimodal particle size distribution were observed for total mass and for individual
  4      components.  Decreased visibility corresponded to increased particle mass concentrations
  5      especially in the fine particle fraction (ranging in size from about 0.1 to 1.00 MM).
  6      Increases in the proportion of particulate sulfate and to a lesser extent of nitrate, chloride,
  7      ammonium, and carbon  were  also associated with decreased visibility.
  8            Aerosol pattern analysis of a major wintertime (1983) pollution episode in northern
  9      New Jersey (Lioy et al., 1985) revealed that the intensity of the episode was the greatest in
10      the area of highest commercial, residential and industrial activities, and that the atmospheric
11      stagnation conditions resulted  in the significant accumulation of aerosol mass.  The aerosol
12      mass was primarily fine aerosols, and  the extractable organic matter compromised about 50%
13      of the particle mass.
14            An analysis of the winter aerosol chemistry data in the Northeast (Poirot et  al., 1990)
15      found that the winter PM2 5 apportionment consisted of 48% sulfates, 23 % organic matter,
16      15% soot, 4% soil, and  10% unexplained.
17
18      6.5.1.5  Philadelphia, PA
19            The metropolitan  area of Philadelphia includes urban-industrial emissions over flat
20      terrain.  The map of the area  shows relatively uniform PM10 concentrations throughout the
21      metropolitan area, with the exception of a single site (AIRS #421010149) in the middle of the
22      urban area.
23            The 1985  to 1992 trend of PMi0 averaged over the metropolitan area is somewhat
24      downward but not significantly.  The seasonal concentration of PM10 (Figure 6-63a) is about
25      30 to 35 /ig/m3 throughout the year, except during the summer months when it rises above
26      40 jug/m3.
27            The seasonal average PM10 concentration for three sites near the center of Philadelphia
28      is shown in Figure 6-64.  One of the sites is the high concentration site noted on the
29      metropolitan map.  The average PM10  concentration at that site ranges between 100 to 150
30      Mg/m3 which is a factor of 2 to 3 higher than the concentration at the neighboring sites.   The
31      daily concentration at this anomalous monitoring site (AIRS #421010149) correlates poorly

        April 1995                                6-105      DRAFT-DO NOT QUOTE OR  CITE

-------
                  Average PM10 -1985-1993
                               AIRS PM10 Concentrantion Trends
                                          Philadelphia
                                                   1986 1986 1987 1988 1989 1990 1991 1992
                                                 AIRS PM10 SEASONAL CONCENTRATION
                                                              Philadelphia
                                                   60

                                                   66

                                                   60

                                                   46

                                                   40

                                                   36

                                                   30

                                                   25

                                                   20

                                                   16

                                                   10

                                                   6
                                                   1985   Mar Apr May Jun Jul Aug Eep Oct Nsv 0«c
Figure 6-63.
Aerosol concentration map, trend and seasonality in the Philadelphia
region.
April 1995
                       6_106      DRAFT-DO NOT QUOTE OR CITE

-------
o\
                100
                90
                80
                70
                60
                50
                40
                30
                20
                10
                      PHILADELPHIA, PA
 1984   Mar   May   Jul   Sep   Nov
|-A- -PM10_AV 421010004 PHILADELPHIA |
 d- -PM25_AV 421010004 PHILADELPHIA
 -4- -PMC JWG 421010004 PHILADELPHIA
                                            250

                                            200

                                            150

                                            100

                                             50
                                                      Philadelphia - Suburban 15 km Separation
                                ilL
                                                       50
                                                              100
                                                                       150
                                                                               200
                                                                                                       PhiladelpNa - Urban, 4km Separation
200

150 ••

100 ••

 50

 0
                                                     4PMKM210100382
                                                                                                                         200
                                                                                                                                250
    Figure 6-64.
         Fine, coarse and PM10 particle concentration near Philadelphia.

-------
 1      with an adjacent site (AIRS #4210104492) which is only 4 km apart (Figure 6-64e).  This is
 2      an indication that the concentration at the anomalous site is under the influence of strong
 3      local source of PM10.  In contrast, two sites in suburban Philadelphia that are 15 km apart
 4      (Figure 6-64d) show a strong correlation of daily measurements indicating that spatially
 5      uniform regional haze influences the daily values at both sites.
 6             Size segregated aerosol samples (Figure 6-64a)  at Philadelphia (AIRS #421010004)
 7      show the fine paniculate mass is the main contributor  of PM10 in this city.  It is possible,
 8      however, that at other sampling sites, e.g. the anomalous #4210104492 site, coarse particles
 9      prevail.
10             Outdoor summertime sulfate (SO42~) concentrations  were found to be uniform within
11      metropolitan Philadelphia (Suh, et al 1995).  However, aerosol strong acidity  (H+)
12      concentrations were found to vary spatially.  Also, the wintertime sulfate pattern was likely
13      to be more heterogeneous in space and time as discussed for Washington, DC  This variation
14     generally were independent of wind direction, but were related to local factors, such as the
15      NH3 concentration, population density, and distance from the center of the city.
16
17     6.5.1.6 Whiteface Mountain, NY
18            The AIRS sampling location at the Whiteface Mountain in Upstate New York, is a
19     high mountain top site, elevated from the surrounding terrain.  The monitoring site offers the
20     possibility of comparing mountain tops concentrations to the surrounding lower elevation
21      sites.
22            The seasonal pattern of PM10 concentration  for Whiteface Mountain  and the
23     surrounding low elevation sites,  Saranac Lake, and Saratoga Spr. are shown in Figure 6-65.
24     The concentration at the three sites  is virtually identical during June through September.
25     However, during the winter the mountain top  site at Whiteface has a PM10  concentration
26     which is only one third of the surrounding sites. This indicates that during the winter, the
27     Whiteface mountain top is above the surface-based aerosol layer, while during the summer
28     the height of the well mixed aerosol layer rises above the  mountain top producing are
29     resulting in a uniform concentration at all sites.  At Whiteface Mountain, NY (Webber et al.,
30      1985) have found direct microscopic evidence of flyash particles during a summertime (June
31     23,  1983) sulfate episode.

       April 1995                               6-108       DRAFT-DO NOT  QUOTE OR CITE

-------
            14
            13C
            12(
            110
            100
            90
            80
            70
            60
            50
            40
            30
            20
            10
                                               PM10 -Average
& B  s-'s-^a-
            0
            1984
                 Mar   May  Jul   Sep   Nov
            £1 I=PM10_AV360330002SARANACLAKE
            H =PM10_AV 360910002 SARATOGA SPR
        Figure 6-65.       PM10 concentration seasonality at Whiteface Mountain and
                          neighborhing low elevation sites.
 1      6.5.2  Sub-regional Aerosol Pattern in the Southeast
 2      6.5.2.1 Winston-Salem, NC, Florida
 3            Winston-Salem is a small urban area imbedded in the relatively homogeneous region
 4      of the Southeast. In fact, the main emphasis of the illustrations below is on the remarkable
 5      uniformity of aerosols in that region.
 6            Comparison of three AIRS PM10 monitoring sites in North Carolina, Winston-Salem,
 7      Greensboro, and Raleigh-Durham (Figure 6-66) shows virtually identical concentrations
 8      (within 10%), both in absolute magnitudes, and in the seasonality.  This is an indication that
 9      these sites are exposed to the same regional airmass throughout the year.  It also indicates
10      that the excess PM10 concentrations due to local urban sources are insignificant.
       April 1995
                              6-109      DRAFT-DO NOT QUOTE OR CITE

-------
o
150
140
130 -
120 -
110 -
100
 90
 80
 70
 60
 50
 40
 30
 20
 10
     1984   Mar    May   Jul    Sep   Nov
     -A- -PM10JW 120950007 ORLANDO
     -H - "PM10_AV 120570095 TAMPA
    [-4. -PM10_AV 120251016 MIAMI      ~~|
                                                         -tr -PM10.AV 370670009 WINSTON-SALEf
                                                         •3 -PM10_AV370810009 GREENSBORO
                                                         -- •PMib_AV37f830003~RALEIGH~ ~
                                                          W1NSTON-SALEM, NC
                                                     100

                                                      90

                                                      80)-

                                                      70

                                                      60

                                                      SO

                                                      40

                                                      30

                                                      20

                                                      10
                                                      1984   Mar  May  Jul   Sep   Nov
                                                     \-tr 'PM10JVV 370670009 WINSTON-SALE(
                                                      -B- -PM2S.AV 370670009 WINSTON^ALEI
                                                      - -PMC.AVG 370670009 WINSTON-SALE
Figure 6-66.
              Aerosol concentration pattern at North Carolina and Florida sites.
April 1995
                                         6-110       DRAFT-DO NOT QUOTE OR CITE

-------
  1           Size segregated monitoring data for Winston-Salem show that fine particles contribute
  2     70 to 80%  of the PM10 mass of 25 to 35  /*g/m3.  Coarse particles are seasonally invariant at
  3     about 10 /-ig/m3 which is typical for eastern US.
  4           The  PM10 concentration at monitoring sites in Florida (Orlando, Miami, Tampa) show
  5     virtually identical concentrations ranging between 25 to 30 pig/m3 throughout the year,
  6     without appreciable seasonality.
  7
  8     6.5.2.2 Large Southeast Metropolitan Areas
  9           The  relative homogeneity of southeastern aerosol concentrations is consistent with the
 10     observations at the regional perspective.  However, the concentrations in large metropolitan
 11     areas such as Atlanta, GA, Birmingham, AL, Mobile,  AL, New Orleans, LA,  and Houston,
 12     TX,  Dallas-Fort Worth, TX, show somewhat elevated  values.  It is evident,  however, that
 13     these southeastern urban areas do not have high concentration hot spots.
 14           The  seasonal PM10 concentration at sites in New Orleans, LA, Mobile and
 15     Birmingham, AL show uniformity (20 to 40 jug/m3) with modest seasonality.  At sites in
 16     Houston, Austin, and San Antonio, TX show similar values, but more spread in
 17     concentration.
 18           The size segregated aerosol samples collected in the cities of the Gulf states, Corpus
 19     Christi,  Forth Worth, Houston, TX and New Orleans,  LA all show that fine particle
 20     concentrations  are relatively low (10 to 20 /ig/m3) and  are seasonal  (Figure 6-67). Coarse
 21      particle concentrations, on the other hand, account for more than half of the  PM10 mass and
 22      their contribution is most pronounced during the summer season.
 23            In Houston,  TX Dzubay et al., 1982 found that in summertime fine particle mass
 24      contained 58% sulfate and 18  % of carbonaceous  material.  They  also found that the coarse
 25      fraction (2.5 to 15 /*m) consisted of 69% crustal matter, 12 %  carbon, and 7 % nitrate
26      species.
27            Characterization of the Atlanta area aerosol (Marshall et al., 1986) show that
28      elemental carbon and paniculate sulfur represent,  respectively 3.1  to 9.9% and 1.9 to 9.4%
29      of the total  suspended particulate mass. The concentration of elemental carbon, sulfur, and
30      TSP exhibit strong  seasonal variations, with elemental carbon decreasing from winter to
31      summer, and sulfur and TSP increasing. Elemental carbon appears to be statistically separate

        April 1995                               6-111       DRAFT-DO NOT QUOTE OR CITE

-------
 1     from sulfur, indicating that the sources for elemental carbon and paniculate sulfur are
 2     distinct.
 3
 4     6.5.2.3 Great Smoky Mountains
 5           Size segregated fine and coarse aerosol concentrations were measured at the Great
 6     Smoky Mountains National Park in September of 1980 (Stevens et al., 1980).   Sulfate and its
 7     associated ions contributed to 61% of the fine particle mass, followed by organics (10%) and
 8     elemental carbon (5%).
 9
10     6.5.3  Sub-regional Aerosol Pattern  in  the Industrial Midwest
11           The  chemical composition of summertime (July, 1981)  fine particles in  Detroit (Wolff
12     et al.,  1982) was found as 52% sulfates, 27% organics, 4% elemental carbon, 8% soil dust
13     Nitrate was found to absent from fine mass. Fine particles, themselves contributed about
14     64% of the aerosol mass balance.
15           Since the  turn of the century, the major cities in the industrial midwestern states had
16     air pollution problems due to smoke and dust.   Pittsburgh, St. Louis, Chicago, and Detroit
17     were among the  formerly notorious air pollution hot spots. The recently acquired PM10
18     database  now allows the re-examination of these metropolitan areas in the industrial Midwest
19     for their  concentration pattern in the 1990s.
20
21     6.5.3.1 Pittsburgh, PA
22           The  average PM10 concentration over the extended metropolitan area is shown in
23     Figure 6-68a.  The map also includes the industrial cities in the eastern Ohio,  Steubenville,
24     OH, and Weirton, OH, located on the Ohio River.  The average PM10 concentration at the
25     80 sites shown on the map varies only by about 20% from site to site.  Outstanding high
26     concentration hot spots are also absent.  It is thus evident that during the 1985 to 1993
27     period, the PM10 concentrations in the Pittsburgh sub-region was spatially  rather uniform.
28           The PM10 concentration trend shows  declining values from about 40 /ig/m3 to 30
29     Mg/m3 (Figure 6-68b) over the 1985 to 1993 period. Figure 6-68b also marks  the
30     concentration standard deviation among the  monitoring sites for each year, which is about
31     15 to 20%  and shrinking over time.

       April 1995                              6-112     DRAFT-DO NOT QUOTE OR CITE

-------
13
 2.
 vo
 0\
  H
  m
   n
   i—i
   s
160
140
130
120
(10
100
9C
80
70
60
60
40
30
20

1

	 . 	 • 	 • 	
-
-
-





A
-_^_V \.
' JJr--"^ "^ a •1-VX*\__A_^
- ' ' V ^^ * i-"*""1
i B a' B """* a


. . . _. —
16O
140
IX
120
110
100
90
80
70
60
60
40
30
20
10

0


•
•
-
•
.

•

•
«- 4_ *"*\' •*
i__J,*^M*-*r" ^=^^

.


                                                                                                            1984   M«r   M«y   Jul   !«P


                                                                                                            4r -PM10.AV 482011038 HOUSTON


                                                                                                            -O -PM10.AV4846300W AUSTIN
                                                                                                                                              1984
                                                                                                           HOUSTON, TX
                                                                                     Mir  Miy  Jul   Sep


                                                                                A- -PWtO_AV220710010 NEW ORLEANS

                                                                                a -PM10.AV010970016 MOBILE

                                                                               ' H -PMiojWOIoYjjOMBiRMINCJHAM  \




                                                                              NEWOLRLEANS, LA
                          1M4  M.r   M«y   Jut   S.p   Nov
                         |T!r -PM10JW493660012 CORPUS CHRIS1


                          •El- -PM2t^V 483660012 CORPUS CHRIS"


                          .+ -PMCJW6483S60012 CORPUS CHRIS
•-tr -PM10JW 464390060 FORT WORTH  .


 D "PM26_AV 4843*0060 FORT WORTH


 •I -PMC AVG 484390060 FORT WORTH
                                                                                                              M«V  Jul
-6- -PM10_AV 462010024 NOT IN A CITY


0  -PM2S_AV 482010024 NOT IN A OTY


-,  -PMC_AVG 482010024 NOT IN A CfTY
1984   Mir   M»y  Jul    Scp   Nov



-ft- -PM10_AV220TiboiO NEW ORLEANS


-f3 -PM26.AV 220710010 NEW ORLEANS


   "PMC_AVG 220710010 NEW ORLEANS
         Fisure 6-67.          Aerosol concentration pattern in Texas and  Gulf states.
            6

-------
                                           AIRS PM10 Concentration Trends
                                                      Pittsburgh
w
55
60

45
0 «°

O ^ 36
'Sis 30
;is
kg 26
3 20
16
10
5
_
-.---•K
' " " , - +-
'"*"''
' 	 *\ ' ""- '*
\»_ __ A "*''' "'-.
>-~D "X. _-^^
^ ^ 	 ^^^A— -^>
\ /
^-^t/ VV^"Q^--Q--:I
-
-
-
-
1985 1986 1987 1988 1989 1990 1991 1992
j"-A- PM10AVG -Q- PM10AYG-SIG -+ PM10AVG + BIG
                                          AIRS PIU110 SEASONAL CONCENTRATION
                                                      Pittsburgh
                                            1986   Mar Apr May Jun Jui Aug Sep Oct Nov Dec
Figure 6-68.       Aerosol concentration pattern and trends in the Pittsburgh subregion.
April 1995
6-114      DRAFT-DO NOT QUOTE OR CITE

-------
  1           The seasonality of the PM10 pattern (Figure 6-68c) is dominated by a summer peak
  2     (45 pig/m3),  which is about 50% higher than the winter concentrations (30 jug/m3).  The
  3     seasonality at sites in Pittsburgh, PA, Weirton, OH, and Steubenville, OH show comparable
  4     values and a seasonal pattern that is slightly above the sub-regional average.  Hence, these
  5     formerly highly polluted locations are now virtually identical to their rural background.
  6           Size segregated aerosol samples in Pittsburgh, PA and Steubenville, OH (Figure 6-69)
  7     show  that fine particles contribute 70 to 80% of the PM10 mass, and also dictate the summer-
  8     peak seasonality of the PM10 concentrations. As in other urban monitoring sites in the
  9     eastern US, the coarse particle concentration in Pittsburgh is about 10 pig/m3 and seasonally
 10     invariant.  The size segregated seasonal data for Steubenville, OH exhibit more random
 11     fluctuations as well as discrepancy between the sum of fine and coarse at one hand,  and
 12     PM10 on the other.  The discrepancy is attributed to the small number of size segregated
 13     aerosol samples.
 14           The remarkable uniformity of fine particles mass and elemental composition from site
 15     to site in the Ohio River Valley was also shown by Shaw and Paur, 1983. Sulfur was the
 16     predominant element in fine particles.  Factor analysis of element concentrations indicated
 17     three clusters throughout the year (1) coarse particle crustal  elements (2) fine particle sulfur
 18     and selenium (3) fine particle manganese, iron and zinc.
 19           The chemical mass balance of Weirton-Steubenville aerosol was examined by
 20     Skidmore  et  al., 1992.  Primary motor vehicle and secondary ammonium sulfate were
 21     dominant contributors to the PM2 5 aerosol.  Steel emissions were also significant
 22     contributors to PM2 5. Wood burning and oil combustion were occasionally detected.
 23     Geological material was the major contributor to the coarse aerosol fraction.  Primary
 24     geological material, primary motor vehicle exhaust, and secondary sulfate were the major
 25     contributors to PM10 at all five monitoring sites.
 26           The composition of size-fractionated summer aerosol in nearby Charleston, West
 27     Virginia was reported by Lewis and Macias, 1980.  Ammonium sulfate was  the largest single
28     chemical component (41%)  of the fine aerosol mass. Carbon was also a large component of
29     both fine  and coarse particle mass constituting 16%  and 12% respectively.  Factor analysis
30     indicated that four factors were sufficient to satisfactorily represent the variance of
       April 1995                               6_H5      DRAFT-DO NOT QUOTE OR CITE

-------
VO
                                                                                      -PM10_AV 420030031 PITTSBURGH



                                                                                   -U  -PM10_AV 640291004 WEIRTON




                                                                                   -H^PMlb~AV''39081l6l2STEUBENVILLE!
                       STEUBENWILLE, OH
PITTSBURGH, PA
PITTSBURGH, PA -2
o\
1
^ 90
80
70
S «o
S
£ 60
H
i 40
O
O 30
o
H
^ 10
0
(~~^ Q


-
-
-


.
tf^eS*~~*~~\
~^* \
ft~~T5r"^~~^iSx^*' \ /
a, ^«^-
i / \ A i
' ^ - + ^Q •"*" ^4. - *^ • -•• \y •* Vi ''
- B • ^S— / ' Q , ' *
+ ""'"•--f'^Q 4 V
IUU

90
80
70
60

50

40

30
20

10
A









/ q \
y / \ v.
»^^^ 'B \ ^N/^
i. ^~* • , » n -£
TS . n n- H /
s- B , \ / -
.(• - t^
-1- •+ '• +
••-!+• ' s...-*-' -
1 UU

90
80
70
60

60

40

30
20

10
n



-
-


_
/^-

X ^^
^^~S ^*~^ ^^^^
^^ ^ ^ "^
'- -0- » B' ^ & -e -*
. -4-..4. -1- -->•-+-•+• -4..^..
--•+ 4 -S--.+ -'

O 198'1 Mar May Jul Sep Nov 1984 Mar May Jul Sep Nov 1984 Mar MaV J"1 Sep Nov
H
t"d . -6- -pMio_AV39oeiioi2 STEUBENVILLE; -A- -PMIO_AV 420030021 PITTSBURGH . -A- »PMIO_AV 420030027 PITTSBURGH '
O Q "PM2S_AV 390811012 STEUBENVILLE -Q- -PM2S_AV 420030021 PITTSBURGH -B =PM25_AV 420030027 PITTSBURGH
_ H -PMC AVG 39081 1012 STEUBENVILLE -4- -PMC AVG 420030021 PITTSBURGH -+ =PMC Av(5 420030027 PITTSBURGH
n ~
W Figure 6-69. Fine, coarse and PM10 concentration near Pittsburgh.

-------
  1     26 measured parameters.  The factors were characteristic of crustal material, ammonium
  2     sulfate, automotive emissions, and unidentified anthropogenic sources.
  3
  4     6.5.3.2 St. Louis, MO
  5            Historically, the St. Louis metropolitan area has been known for high particulate
  6     concentrations, particularly on the  east side of the Mississippi River.  The map of the
  7     metropolitan area (Figure 6-70a) shows about factor of 2 to 3 concentration differences
  8     among the PM10 monitoring stations.  The monitoring sites east of the river tend to be higher
  9     than the western sites of  this sub-region.
 10           The average PM10 in the St. Louis metropolitan area (Figure 6-70b) has been
 11     declining from 40  to 45 ng/m3 to 25 to 30 jug/m3 by 1993. This decline is comparable to the
 12     average reductions over the industrial midwestern region.  The seasonality of the
 13     sub-regionally averaged concentrations  (Figure 6-70c) shows the summer peak with 40 to 50
 14     Mg/m3 which is about 50% higher than the winter averages.
 15           Seasonal comparison of the individual monitoring sites in the area shows that Granite
 16     City, IL and East St. Louis, IL have higher PM10 concentrations throughout the year
 17     compared  to western St. Louis, MO sites.
 18           Size segregated aerosol samples  at three sites west of the Mississippi River (Ferguson,
 19     MO, Affton, MO and Clayton, MO) show that fine particles are mostly responsible for
20     PM10, including the seasonality (Figure 6-71).  Coarse particles contribute 10 jig/m3 or less
21      throughout the year, although corresponding size segregated aerosol data for more polluted
22      eastside of the Mississippi River are not available.
23            Monitoring the diurnal and seasonal patterns  of particulate sulfur and sulfuric acid in
24      St. Louis (Cobourn and Husar, 1982) noted an afternoon increase in particulate sulfur
25      concentration of about 20%.  For the summertime,  particulate sulfur concentration was
26      higher than the annual mean by 40%.
       April 1995                               6_117       DRAFT-DO NOT QUOTE OR CITE

-------
                Average PM10 -1986-1993
                             AIRS PM10 Concentrantion Trends
                                         St Louis
                                                     1986 1987 1988  1989 1990  1991  1992
                                                AIRS PWI10 SEASONAL CONCENTRATION
                                                  1985   Mir Apr M«y Jun Jul  Aug S«p Oct N»v Dec
Figure 6-70.
Aerosol concentration pattern and trends in the St. Louis subregion.
April 1995
                      6-118
DRAFT-DO NOT QUOTE OR CITE

-------
  1      6.5.3.3  Chicago, IL
  2            Chicago has also been known for industrial dust, smoke,  and haze, particularly in East
  3      Chicago and Gary, IN.  The average PM10 concentrations over  the Chicago sub-region
  4      (Figure 6-72a) shows that concentrations vary by a factor of two or less throughout the
  5      subregion.  The downward trend (from 40 to 25 /ig/m3) is comparable to the PM10
  6      reductions over the industrial Midwest. The seasonality of PM10 is also typical with the
  7      summer peak of 40 /-ig/m3 and winter values of 20 to 30 />ig/m3.
  8            Superposition  of seasonal PM10 data at Chicago, IL, East Chicago, IL, and Gary, IN
  9      conveys a remarkable spatial uniformity, as  well as comparatively low PM10 concentrations
 10      in this area that has historically been  a smoky and dusty industrial sub-region.
 11            Chemical composition measurements in Chicago (Lee  et al., 1993) showed that me for
 12      an concentrations for SO42' (5.55 /*g/m3), NH4+ (2.74 ^g/m3), NH3 (1.63 /xg/m3),
 13      HNO3(0.81 ng/m3),  HNO2 (0.99 pig/m3), for SO2 (21.2 pcg/m3), for NO3- (4.21  /ig/m3), and
 14      for H+ (7.7 nmol/m3).  The highest values occurred in the summer, except for HNO2 and
 15      NO3" which had the highest values in the winter.
 16            Comparison of atmospheric coarse particles at an urban and non-urban site near
 17      Chicago, IL show that the concentration were 50% higher during mid-day than at night. Dry
 18      ground samples were 30 % higher than wet ground and 90% higher than frozen ground
 19      samples. (Noll et al., 1985).
 20            The analysis of coarse particles in Chicago, IL (Noll et al., 1990) show that the coarse
 21      particle mass could be divided  into two categories: material that was primarily of crustal
 22      origin (Al, Ca, Fe, and  Si) and material that was primarily of anthropogenic origin (Cd, Cu,
 23      Mn, Ni, Pb, and Zn).  The mass of crustal material varied between 15 and 50%  of the total
 24      coarse mass, while the mass of anthropogenic material was < 1%.
25            The composition of atmospheric coarse particles at urban  (Chicago, IL) and non-urban
26      (Argonne, IL) were reported by Noll  et al.,  1987.  Limestone and silicates were  the main
27      source of material at  the non urban site.  Anthropogenic sources, represented by  fly ash and
28      coal,  were present in the industrial sector sample and rubber  tire was present in the
29      commercial sector sample.  The mass median diameters (MMD) for different components
30      were  as follows: limestone (20/mi, silicates 12jum, coal, flyash and iron oxide (12^m) and
31      rubber tire (25/mi).

        April 1995                              6-119      DRAFT-DO NOT QUOTE OR CITE

-------
                                                        150
                                                        140
                                                        130
                                                        120
                                                        110

                                                        90
                                                        80
                                                        70

                                                        60
                                                        40
                                                        30
                                                        20
                                                        10
                                                         1984  Mar  May  Jul   Sep
                                                         -&- -PM10_AV29S100080STLOUIS
                                                     Nov
                                                           -=PM10_AV171S30010 EAST ST LOUIS
         ST LOUIS, MO
            ST LOUIS, MO -Clayton
                                                                         ST LOUIS, MO -Affion
 1984  Mar
^A- -PM10_AV29189500l"FERGUSON
-B- »PM25_AV291895O01 FERGUSON
-4- -PMC.AVG 291895001 FERGUSON
                                      -PM10JW 291692003 CLAYTON
                                    -Q- -PM25_AV 291892003 CLAYTON
                                    •+ "PMC.AVG 291892003 CLAYTON
                                        [-A- •PM10_AV29189000fgFTQJr
                                         -Q- -PM26_AV291890001 AFFTON
                                         •4- "PMC_AVG 291890001 AFFTON
      Figure 6-71.
Fine, coarse, and PM10 concentration pattern near St.  Louis.
April  1995
                    6-120       DRAFT-DO NOT QUOTE OR CITE

-------
               Average PM10 -1985-1993
  o
^ 01
< ui
O UJ
s*
O- P
150
140
130
120
110
100
 90
 80
 70
 60
 60
 40
                                                  AIRS PM10 Concentrantlon Trends
                                                                Chicago
                                                   1985 1986  1987 1988 1989  1990 1991  1992
                                                 AIRS PM10 SEASONAL CONCENTRATION
                                                              Chicago-Gary
                                                  60,	.	.	,	.	.	.	-.	r-	.	,	,
 1984  Mar   May   Jul   Sep  Nov
 -A- =PM10_AV 170310014 CHICAGO
 -fl "PM10_AV 180890006 EAST CHICAGO
i + -=PM10_AV 180890022 GARY       |
55
50
45
40
35
30
26
20
15
10
6
                                                   1985   Mar Apr May Jun Jut Aug Sep Oet Nov D«e
    Figure 6-72.
                   Aerosol concentration pattern and trends in the Chicago subregion.
     April 1995
                                          6-121
        DRAFT-DO NOT QUOTE OR CITE

-------
 1     6.5.3.4 Detroit, MI
 2            Detroit is the center of the automotive industry and has been studied by several
 3     investigators.
 4            In Detroit, July, 1981 (Wolff and Korsog, 1985) fine mass average was found to be
 5     42.4 /ig/m3. A major contribution (50%) of sulfate source which appears to be coal
 6     combustion was identified. The coarse fraction which averaged as 25.8 /ig/m3 was
 7     dominated by  crustal material which accounted for about two-thirds of the coarse material.
 8     Significant contributions were also identified from motor vehicles (mostly due to reintrained
 9     road dust) and iron and steel industry emissions.
10            The seasonal variations in nitric acid, nitrate, strong aerosol acidity, and ammonia in
11     urban  are, Warren, MI was examined by Cadle, 1985.  The greatest variations was for
12     ammonia, which was 8.5 times higher in summer than winter.  The least variation was for
13     paniculate nitrate which had a summer maximum only 1.8 times higher than in spring
14     minimum. It  was  noted that ammonium nitrate volatilization from filters and impactors can
15     cause large errors  in summertime measurements, but the errors  are not significant during the
16     winter.
17            The influence of local and regional sources on the concentration of particulate matter
18     in urban and rural sites near Detroit, MI was investigated by Wolff et al., 1985.  Analysis of
19     spatial variations of the various particulate components revealed: (1) at all four sites the
20     PM2 5  was dominated by regional influences rather than local sources  . The site in industrial
21     sector  had the largest impact of local sources, but even at his site the local influences appears
22     to be smaller than he regional ones.  (2) the regional influences were most pronounced on the
23     sulfate levels which accounted for the largest fraction (40-59=0%) of the PM2 5. (3) organic
24     carbon compounds were the second most abundant PM2 5 species accounting for 20 to 49%
25     of the  mass.  Organic carbon seems to  be controlled by both local and regional organic
26     carbon influences.  Vehicular emissions and possibly secondary reactions appear to affect the
27     organic carbon concentrations   (4) elemental carbon appears to be dominated by local
28     emission (5) PMCoarse was dominated by local sources, but at the industrial site unknown
29     non-crustal elements were significant components of coarse mass.
        April 1995                               6-122      DRAFT-DO NOT QUOTE OR CITE

-------
  1     6.5.5 Sub-regional Aerosol Pattern in the Southwest
  2           Southwest is a dusty region and much of the discussion below pertains to coarse
  3     particles and soil dust.
  4           The arid southwestern U.S. metropolitan areas (El Paso, TX, Phoenix-Tucson, AZ)
  5     has modest industry and national parks (grand Canyon) where the prevention of visibility
  6     degradation has been stated as a national goal.
  7
  8     6.5.5.1  El Paso,  TX
  9           The PM10 concentration in the El Paso, TX subregion shows that the high and low
 10     concentration sites can be found in close vicinity of each other (Figure 6-73a).  This is an
 11     indication that local  sources  of PM10 with limited range of impact are important. The PM10
 12     trend since 1985 (Figure 6-73b) shows a remarkable reduction from 60 /xg/m3 to 30 /xg/m3,
 13     although the downward trend is not  monotonic.  This substantial reduction parallels the factor
 14     of two PM10 decline over the entire southwestern region.
 15           The seasonality of PM10 over the El Paso, TX sub-region (Figure 6-73c) is bimodal
 16     with peaks in the spring time, March through July, as well as another peak, October through
 17     November. This double peak seasonality at El Paso, TX also parallels the seasonality of the
 18     entire  region.  The double peak is further illustrated by superimposed seasonal charts  for
 19     Tucson, AZ, Albuquerque, NM and El Paso, TX.  It is evident, that these three widely
 20     spaced sampling sites all show the concentration reduction in August which coincides  with
 21      the arrival of moist "monsoon" flow from the Gulf of Mexico toward Arizona.
 22           Size segregated aerosol samples for El Paso, TX (AIRS #481410037) shows that
 23      coarse particles dominate the PM10 concentrations, accounting for about 70% of the PM10
 24      mass (Figure 6-74a).  This is consistent with the important  role of coarse particles over the
 25      arid Southwest. In comparison, size segregated data for San Antonio, TX (Figure 6-74b)
 26      located closer to the  Gulf Coast in Texas, shows that fine and coarse mass  have comparable
 27      contributions, similar to Houston,  TX.
28
29      6.5.5.2 Phoenix and Tucson, AZ
30            The Phoenix-Tucson sub-region (Figure 6-75a) shows substantial PM10 concentration
31      range.  Samplers within Phoenix or Tucson area indicate 2 to 3  times higher concentrations

        April 1995                               6_123      DRAFT-DO NOT QUOTE  OR CITE

-------
         Average PM10 -1985-1993
   AIRS PM10 Concentrantion Trends
                 El Paso
                                                1986 1387 1988 1989  1990  1991 1992
                                          AIRS PWI10 SEASONAL CONCENTRATION
                                                        El Paso
                                           80,	.	.	.	,	.	.	,	,	.	,
                                      3!
                                       |
    75
    70
    65
    60
    55
    60
    45
    40
    35
    30
    25
    20
    15
    10
     5
                                            1985   Mar Apr May Jun Jul Aug s«p Oet Nov D«e
Figure 6-73.       Aerosol concentration pattern and trends in the El Paso subregion.
April 1995
6-124      DRAFT-DO NOT QUOTE OR CITE

-------
       100
        90 -

        80 -

        70

        60
        50
        40

        30
        20
        10
                  EL PASO, TX
                                              SAN ANTONIO, TX
H --Q
                         3"
                            5-
        1984   Mar    May   Jul    Sep   Nov
        -A- =PM10_AV481410037ELJPASO
        -O- =PM25_AV 481410037 EL PASO
        ~4- =PMC AVG 481410037 EL PASO
100
 90

 80

 70

 60
 50

 40

 30
 20

 10
                                       1984   Mar   May    Jul    Sep   Nov
                                      -A- =PM10_AV 480290036 SAN ANTONIO
                                      -ED- =PM25_AV480290036 SAN ANTONIO
                                       -(- =PMC_AVG 480290036 SAN ANTONIO
      Figure 6-74.      Fine coarse, and PM10 concentration pattern near El Paso.


1     than the more remote sites, particularly the ones in the mountains.  There is a general decline
2     of PM10 level between 1985 and 1993, but it is in the presence of substantial year to year
3     variation (Figure 6-75b).  The average PM10 seasonality of the Phoenix-Tucson sub-region
4     (Figure 6-75c) shows the bimodal spring and fall peak pattern which is characteristic for the
5     entire Southwest region.
6           The wintertime aerosol chemical pattern in Phoenix was reported by Chow et al.,
7     1990.  Solomon and  Moyers  (1986). They found fine particle crustal species, sulfates,
8     nitrates, and organic  and elemental carbon to be at least five times higher hi concentration
9     when comparing samples during a haze episode to samples taken during good visibility.
      April 1995
                                 6-125
     DRAFT-DO NOT QUOTE OR CITE

-------
       Average PM10 -1986-1993
                                        AIRS PM10 Concentrantion Trends
                                                  Phoenix-Tucson
                                          1985  1986  1987 1988 1989 1990  1991 1992
                                       AIRS PM10 SEASONAL CONCENTRATION
                                                   Phoenix-Tucson





cr
10_AVG
/1ETER[25
1






Ol/
75
70
65
60
55
50
45
40
35
30
25
20
15
10
5
0
_
-
-
-
-
.
-
x/^^Xv>/'^
~
~
.
.
_
	
                                         1888   Mar Apr May Jun Jul  Aug S«p Oct Nev Dee
Figure 6-75.       Aerosol concentration pattern and trends in the Phoenix-Tucson
                  subregion.
April 1995
6-126
DRAFT-DO NOT QUOTE OR CITE

-------
  1      During the Phoenix Urban Haze Pilot Study during the winter 1988 to 1989 (Frazier
  2      C.A., 1989) a definite diurnal cycle in PM2 5 concentrations was observed.  The maximum,
  3      generally but not always,  occurred at night, which is consistent with the meteorological
  4      observations or poor dispersion and dilution.
  5            A chemical characterization of wintertime fine particles in Phoenix, AZ (Solomon and
  6      Moyers, 1986) showed a dominance of organic and nitrate aerosols.  The composition of the
  7      Phoenix wintertime haze is most like that of Denver, CO, a city which also experiences
  8      wintertime inversions (Pierson and Russell, 1969; Countess et al., 1980;  Groblicki et al.,
  9      1981). In both cities the average measured  NO3" concentrations was  about 1 to 2 times that
 10      of average SO42" concentration. In addition, the average SO42"concentration measured in
 11      Phoenix was much lower than those observed at other locations throughout the US, but
 12      similar to the regional values observed in the Southwest (Moyers,  1981).
 13            Wintertime PM10 and PM2 5 chemical compositions and source contributions in
 14      Tucson,  AZ  (Chow et al.,  1992) show that the major contributors to the highest PM10
 15      concentrations were geological material (>50%) and primary motor vehicle exhaust
 16      (> 30%) at  three urban sampling sites.  Secondary ammonium sulfate, secondary ammonium
 17      nitrate, and copper smelter aerosols were found to contribute less than 5% to elevated PM10
 18      concentrations.
 19            It is instructive to compare chemical concentrations in Phoenix with wintertime values
20      in other mountainous, arid communities (Denver (commercial), CO,  Reno (commercial),
21      NV, and Sparks (residential), NV.  Organic carbon (OC) at the Phoenix site was twice the
22      elemental carbon at the Denver and Sparks sites, while the OC/EC ratio was one to one  at
23      Phoenix and  Reno sites.  Wood burning contribution at Reno site was very low.  The
24      average arsenic concentrations in Phoenix was  four times higher than observed in other
25      cities, which indicates the potential influence of Arizona smelters located within 100 miles  of
26      Phoenix.  Average sulfate levels in Phoenix were higher than they were in Denver, which
27      has less local emissions of SO2.  Nitrates were major source of visibility  reduction in
28      Denver. In Phoenix, nitrates were significant, but the carbonaceous species appear  to have
29      much larger role in visibility impairment.  The average light absorption (babs in Phoenix  was
30      nearly a factor of three higher than the averages obtained in Denver.
31

        April 1995                               6-127      DRAFT-DO NOT QUOTE OR CITE

-------
  1      6.5.6  Sub-regional Aerosol Pattern in the Northwest
  2            The mountainous northwestern United States has many aerosol regions with different
  3      characteristics.  The discussion below will examine South Lake Tahoe, as a case study for
  4      mountain-valley difference, Salt Lake City, UT, Denver, CO, Idaho-Montana sites, and
  5      several Washington-Oregon sites.
  6            Dresser 1988 investigated the winter PM10 concentrations in a small ski resort town,
  7      Telluride, CO and found that the street dirt and sand are major contributors, particularly
  8      during the dry post snow period.
  9            Wintertime aerosol characterization and source apportionment was also conducted for
 10      San Jose, CA, attributing 45% of the PM10 mass to residential wood combustion (Chow
 11      etal.,1995).
 12            In Portland, OR, carbonaceous aerosol was found to account  for about 50% of fine
 13      aerosol mass (Shah et al., 1984)
 14
 15      6.5.6.1 South Lake Tahoe
 16            South Lake Tahoe IMPROVE  monitoring site is located in a in a populated area on the
 17      south shore  of Lake  Tahoe. The Bliss State Park IMPROVE monitoring site is to the
 18      northwest, elevated (700ft) and removed from the populated areas.   The pair of sites
 19      illustrates the  populate-remote difference in aerosol pattern. The aerosol  and visibility at the
20      two lake Tahoe sites were also examined (Molenar et al., 1994)
21           The concentration of all aerosol components is substantially higher on the south lake
22      shore compared to the more remote site.  The seasonality and chemical composition is also
23      substantially different.  The excess PM10 concentration at the S. Lake Tahoe site compared
24      to Bliss State Park (Figure 6-76) is about 5 jLtg/m3  during the warm season, May through
25      September, and it climbs to 28 /^g/m3 excess in January. The factor of five seasonal
26      modulation for valley excess PM10 is likely contributed by winter time emission sources,
27      poor dispersion compared to the summer, as well as fog, that tend to enhance the aerosol
28      formation.  Fine and coarse particles contribute roughly equally to excess PM10 mass
29      concentration.  However, fine particles contribute about 60% during the  fall season and
30      coarse particles prevail (>60%) during the spring.  Both fine and coarse particles  show a
       April 1995                              6-128      DRAFT-DO NOT QUOTE OR CITE

-------



03
0
C
£
^B
J£
5
_^
0.
«
s
(A

S
o
(0
^
09
5
to

— 1 	 1 	 1 	 1 "---I. 	 1 J-.J 	 1 	 1 	 1 	 1 — • 1 ' '
~ - • ^*
" '- "E3 •-•' +
•'-,._ tk^ .1 \

" ^^~ ^ra
• >-. 13 , • "^x^ I
.X*"' 0 ^«
^..y^eT ^.
. -.;-• ' ' ^O^ 4 +«
---' — ^" ' i
i i i i -_ -r i —i, -f"" i i i i v i ,'
ooooooooooooc
oooooooooooo
oooooooooooo
 O co u.
4 a 9 ^
i n
s
«
0
1 2
2 «
m i i o>
§i i ^
s T- £\ o
i£ i — i
3 ! O =>
CO i CO CO
4> ! t t
                              9U\J JO UOflOBJJ
                                                                     •8
                                                                     8
                                  •S
                                  H


                                  I
                            Oi
                                                               a.


                                                               t
                                                               o
                                                               r-


                                                               a.
                                                                     §

                                                                     !
                                                                     I
                                                                     "3
                                                                     3
                                   £UJ/6ll
                                  3,
April 1995
6-129     DRAFT-DO NOT QUOTE OR CITE

-------
 1      winter peak concentration. The chemical composition of the coarse mass is not known but
 2      both soil dust or the fine particle species are plausible.
 3            The chemical composition of the valley excess fine particle mass concentration also
 4      shows a strong seasonality for organics and soot. In fact, the excess organics concentration
 5      in the winter (13 /xg/m3) is almost an order of magnitude higher than the summer values.
 6      The seasonal concentration of excess fine particle soot is similar to that of the organics.
 7      However, the relative magnitude of winter organics compared to soot is higher in the winter
 8      (factor of five) than in the summer (factor  of two).   The concentration of fine particle sulfate
 9      is virtually identical for South Lake Tahoe and Bliss State Park. This implies that the South
10      Lake Tahoe aerosol sources do not contain sulfur.  It is also worth noting that the excess fine
11      particle soil at South Lake Tahoe is below  1 /jg/m3, which is a  small fraction of the coarse
12      mass.  Thus, the crustal component of the  South Lake Tahoe aerosol contributes to the
13      coarse mass but not appreciably to  the fine mass concentration.
14            In summary, there  is a significant excess  PM10 aerosol concentration S. Lake Tahoe
15      compared to the adjacent Bliss State Park remote site, particularly during the winter season
16      (28 jug/m3).  The excess mass is about equally distributed between fine and coarse particles.
17      The fine mass is largely composed  of organics.
18
19      6.5.6.2 Salt Lake City, UT Sub-region
20            Salt Lake City, Ogden, and Provo, UT are part of an airshed that is confined by tall
21      mountains to the East, limiting the  dispersion by westerly winds.
22            The seasonal average PM10 concentration at three AIRS sites in Salt Lake City,
23      Ogden, and Provo, UT is shown in Figure 6-77b.  All three sites show virtually identical
24      seasonality, having peak concentrations during December through January.  This confirms
25      that the three sites belong to the same airshed with  similar source pattern, meteorological
26      dispersion and chemical transformation and removal processes.
27            The size segregated fine and coarse  concentration data exhibit a dynamic seasonal
28      pattern.  Fine particles clearly dominate the high winter concentrations reaching 40 to
29      50 pig/m3, compared to 10 /ig/m3 of summer values.  This magnitude of fine mass
30      concentration is among the highest  recorded in the AIRS data system.  Coarse particles are
31      less seasonal and they are more important during the dry summer season. The formation of

        April 1995                                6-130      DRAFT-DO NOT QUOTE OR CITE

-------
                                                                1984   Mar   May   Jul    Sep   Nov
                                                                -6- =PM10_AV 490350012 SALT LAKE CITY
                        100
                         90
                         80
                         70
                         60
                         60
                         4O
                         30
                        20
                        10
                               SALT LAKE, UT
 100
      J3 -PM1 0JW 490S700(H OgDENI ~     I
      -4- «PM10_AV490490002 PROVO

       SALT LAKE, UT - 2
                        1984   Mar  May  Jul   S«p   Nov
                       | -&- -PM10_AV 490360003 NOT IN A CHY^
                        -B- -PM26_AV 490360003 NOT IN A COY
                        -+ -PMC.AVG 490360003 NOT IN ACHY
 1984  Mar   May   Jul    Sep  Nov
[•A- -PM10.AV490363001 SALTLAKECrg
-O- -PM2S_AV49036300t SALTLAKE CfTY
-+ -PMOVG 490363001 SALT LAKE Crr
Figure 6-77.        Aerosol concentration pattern near Salt Lake City.
April  1995
                                               6-131       DRAFT-DO NOT QUOTE OR CITE

-------
 1      sulfate and nitrate during winter inversion fogs near Salt Lake City, UT were studied by
 2      Mangelson et al., 1994.
 3
 4      6.5.6.3 Denver, CO
 5           The Denver "brown cloud is a manifestation of high wintertime concentration of
 6      particles and gases.  Several recent studies have focused on the characterization of the
 7      Denver brown cloud aerosols.
 8           Size distribution measurements of winter Denver aerosol (Countess et al., 1981) show
 9      that on high pollution days that the MMAD of the accumulation mode aerosol was about
10      0.31 jum with ag±2.0.  Wolff et al.,  1980 found that on the average motor vehicles were
11      responsible for 27%  of the elemental carbon while wood burning was responsible for 39% of
12      the elemental carbon.
13           The chemical composition of wintertime Denver fine aerosol mass (16.4 ptg/m3)
14      (Sloane et al., 1991) shows the dominance of organic (8.1 /wg/m3) and elemental carbon
15      (2.6/ug/m3) oversulfa evidence that the fine particle sulfate and nitrates are bimodal,
16      composed of 0.2 to 0.3/mi and 0.4 to 0.6 /mi modes.
17
18      6.5.6.4 Northern Idaho-Western Montana Sub-region
19           The mountainous northern Idaho and western Montana  is characterized by deep
20      valleys, absence of major  industrial sources, or large urban-metropolitan areas.
21      Nevertheless, PMj0 monitoring sites in northern Idaho and western Montana report
22      concentrations that are among the highest in the nation, as illustrated in  Figure 6-78a, while
23      adjacent sites are among the lowest.  The large spatial concentration variability is evidently
24      related to the rugged terrain.  Most of the monitoring sites are located in the flat valleys.
25           The PM10 concentration trend (Figure 6-78b) is strongly downward, with a factor of
26      two reduction (from 54 to 27 /ig/m3) between 1985 and 1993.  The average seasonality of
27      the sub-region is strongly  winter peaked (Figure 6-78c) with a factor of  two modulation
28      between 25 and 45 Mg/m3.
29           A typical example in northern Idaho (Figure 6-79a), where three  adjacent sites show
30      winter monthly averaged peak concentrations of 50 to 85 jug/m3.  This is higher than the
31      monthly average PM10 concentration anywhere in the eastern US.

        April 1995                               6-132      DRAFT-DO NOT QUOTE OR CITE

-------
           Average PM10 -1985-1993
      AIRS PM10 Concentrantion Trends
              N. Idaho - NW.Montana
       60 P	.	•	•	•	1	.	-
                                               1985 1986 1987 1988  1989 1990 1991 1992
                                             AIRS PM10 SEASONAL CONCENTRATION

                                                       N Idaho-NW Montana
                                              65 I	•	•	•	•	•	•	•	•	•	•	r

                                           o
      60

      55

      50

      45

      40

      35

  ; s  so

   3  26

      20

      15

      10

       6 -
                                               1985   Mar Apr May Jun Jui  Aug s«p Oct Nov D«c
Figure 6-78.      Aerosol concentration pattern and trends in the N. Idaho-NW
                  Montana subregion.
April 1995
6-133      DRAFT-DO NOT QUOTE OR CITE

-------
                                                         S"!
                                                               O
                                                               g
                                                               o
                                                               §
                                                               o
April 1995
6-134    DRAFT-DO NOT QUOTE OR CITE

-------
  1     during December through February the concentrations are elevated to about 50 /-ig/m3. The
  2     third site (AIRS#300630034) shows the highest winter peak (> 100 Mg/m3), but summer
  3     values that are comparable to the other two sites.  The latter site is the closest to the city
  4     center.  It is evident that in Missoula, MT high concentration gradients exist between the
  5     populated areas and remote sites.  Boise and Salmon, ID (Figure 6-79e) also show elevated
  6     PM10 concentrations during the cold season. Idaho Falls, ID on the  other hand, is seasonally
  7     uniform at about 30 yug/m3, which is  comparable to the background (AIRS#300630020)
  8     Missoula, MT site.
  9           Remarkably low PM10 concentrations of 10 jug/m3 are reported at three PM10
 10     monitoring sites near Anaconda-Deer, ID (Figure 6-79f).  This is remarkable because the
 11     sites are in a  valley, and that the characteristic  winter peak is completely absent.  This
 12     suggests that pristine,  low, PM10 sites can exist in the northwestern valleys, and hence the
 13     region is not uniformly covered by wintertime haze or smoke.
 14
 15     6.5.6.5 Washington-Oregon Sub-region
 16           The Pacific Northwest is also a mountainous sub-region that exhibits unique aerosol
 17     characteristics.  PM10 monitoring sites in Seattle, Bellevue, and Tacoma, WA  (Figure 6-80)
 18     show relatively low concentrations between 20 to 40 jiig/m3, with the higher values occurring
 19     during the winter months. A much more pronounced seasonality of PM10 concentrations is
 20     recorded in southern Oregon,  Medford, Grants Pass, and Klamath  Falls, OR.  These sites
 21      evidently belong to an airshed of which emissions, dispersion, and aerosol formation
 22     mechanisms are conducive for the formation of winter time aerosol (60 to 80 /xg/m3).
 23            Fine and coarse particle data collected over limited period in 1987 show that the
 24      winter peak of PM10 is entirely due to the strong winter peak of fine particle mass (50 to 100
 25      Mg/m3). Coarse mass, on the other hand, is seasonally invariant at about 10 to 20 /wg/ni3.
 26      Fine particles clearly are responsible for the winter peak.  This is somewhat different from
 27      the observations at South Lake Tahoe, where the  winter peak was attributed to both fine and
28      coarse particles.
29            The size segregated aerosol data for Bend and Central Point, OR, show diminishing
30      concentrations compared to Bedford, where the reduction of PM10  is  mainly due to the
31      decrease of the fine particle mass during the winter season.

        April 1995                                6-135     DRAFT-DO  NOT QUOTE OR CITE

-------
ON
           160
           140 -
           130 -
           120 -
           110
           100
            90
            80
            70
            60
            60
            40
            30 f=
            20
            10
            1984   Mar   May   Jul   Stp
            -ir -PM10.AV 630330073 SEATTLE
            j -a- •PM10_AV 630330004 BELLJVUE"
                                        Nov
            .4. -PM10_AV 630630021 TACOMA
                                MEDFORD, OR
                                                              100
                    Il-
                   ia:
                                                                         -£r -PM10.AV 410293001 MEDFORD
                                                                         -Q -PM10JW 410330008 GRANTS PASS
                                                                         '.+ 'ipwioJw'41036ob04'KLAM'AT>l FALL.)
                                           BEND, OR
                                                                             CENTRAL POINT, OR
                                                                       100
                                                                        90
                                                                        80
                                                                        70
                                                                        60
                                                                        60
                                                                        40
                                                                        30
                                                                        20
                                                                        10
                        1984   Mar   May  Jul   Sep   Nov
                       |-A- •PWr6_AV'41029300l"MEDFORD'''  j
                        -B- «PM2S_AV 410293001 MEDFORD
                        -I- -PMC_AVG 410293001 MEDFORD
                                 1984   Mar   May   Jul    S.p   Nov
                                                                                                      1984  Mar   May   Jul    S«p   Nov
                                ;-&- -PM10_AV 410170001 BEND
                                 -B- -PM2SJW 410170001 BEND
                                 •+ -PMC.AVO 410170001 BEND
                                                                        hi- •PM10_AV410291001 CENTRAL
-a- "PM26_AV 410291001 CENTRAL POINT
•+ -PMC.AVG 410291001 CENTRAL POIN-
      Figure 6-80.
Aerosol concentration pattern in Washington State and Oregon.

-------
  1     6.5.7  Sub-regional Aerosol Pattern in the  Southern  California
  2           The southern California region has two sub-regions, the San Joaquin Valley and the
  3     Los Angeles-South Coast Air Basin, discussed separately in sections below.
  4
  5     6.5.7.1 San Joaquin basin
  6           The wide air basin between the coastal mountain ranges of California to the west and
  7     the Sierra Nevada Mountains to the east shows a remarkably uniform PM10 concentrations as
  8     indicated on the map (Figure 6-8la).
  9           There  is evidence of PM10 concentration reduction from 55 to 40 pg/m3, but the trend
 10     is not conclusive (Figure 6-81b). The seasonal modulation amplitude  over the San Joaquin
 11     Valley (Figure 6-8Ic) is about factor  of 2.5 between the low spring concentration 30 to 35
 12     Atg/m3, and high fall concentration (60 to 70 /-ig/m3).  The unique feature of this seasonality
 13     is the fall peak which differs from the summer peak  in the eastern United States and winter
 14     peak over the mountainous northwestern states.
 15           The AIRS database contains  valuable size segregated fine  and coarse particle
 16     concentration data within the San Joaquin Valley, as  shown in Figure  6-82 for Fresno,
 17     Madera, Visalia, and Bakersfield, CA.  These monitoring sites show virtually identical
 18     concentration pattern for fine and coarse mass.  Both coarse and fine particles are important
 19     contributors to the San Joaquin Valley PM10 aerosol.  However,  their respective prevalence
 20     is phase shifted.  Fine particles are  most important during November through February
 21      winter season, while coarse particles prevail during June through November.  As a
 22      consequence,  in November, both coarse and fine particles are present  causing the seasonal
 23      peak of PM10.  Conversely, during  March through May, neither fine or coarse particles are
 24      abundant and  the PM10 concentration  is lowest during the spring season.
 25            The temporal dynamics of the emissions, ventilation and aerosol formation in the San
 26      Joaquin Valley was a subject of detailed aerosol monitoring, and source apportionment
 27      studies.
 28            The aerosol composition at non-urban sites (Chow et al.,  1995) provide further
29      characteristics of the central California aerosol pattern (Figure 6-82).   A PM10 aerosol study
30      was carried out at six sites in California's San Joaquin Valley from 14 June 1988 to 9 June
31      1989,  as part  of the 1988 to 1989 Valley Air Quality Study (VAQS).   Concentrations of

        April 1995                               6_137      DRAFT-DO NOT QUOTE OR CITE

-------
              Average PM10 -1985-1993
                                                 60
                                               AIRS PM10 Concentration Trends
                                                         San Joaquin Valley
                                                 1985  1986  1987 1988 1989 1990  1991  1992

                                               AIRS PM10 SEASONAL CONCENTRATION
                                                         San Joaquin Valley
                                                1985   Mar Apr May Jun Jul  Aug S«p oet Nov D«e
Figure 6-81.       Aerosol concentration pattern and trends at San Joaquin Valley.
April 1995
6-138     DRAFT-DO NOT QUOTE OR CITE

-------
              FRESNO, CA
                                                      MADERA, CA
  a
  10

o^
> a:
< LU

o't
100



 90



 80



 70



 60



 50



 40



 30



 20



 10
      1984   Mar   May   Jul   Sep   Nov
                PM10_AV 060190005 FRESNO
               VISALIA, CA
       1984   Mar    May  Jul    Sep   Nov
                "PM10JW 061072002 VISALIA	
o ui
     100



     90



     80



     70



     60



     50
                                               O   40

                                               S

                                                   30



                                                   20



                                                   10
                                              1984   Mar   May  Jul    Sep
                                  Nov
                                                   |-A- ePM10_AV 060390002 MADERA
                                                    BAKERSFIELD, CA
                                                o
                                              <
                                               '
                                                UJ
                                                o
                                              100



                                              90



                                              80



                                              70



                                              60



                                              50



                                              40



                                              30



                                              20



                                              10
                                               1984   Mar   May   Jul    Sep   Nov
                                                        PM10_AV 060290004 BAKERSFIELD
      Figure 6-82.      Fine, coarse and PM10 concentration pattern in the San Joaquin

                       Valley.
      April 1995
                                       6-139     DRAFT-DO NOT QUOTE OR CITE

-------
  1      PM10 and PM2 5 mass, organic and elemental carbon, nitrate, sulfate, ammonium, and
  2      elements were determined in 24-h aerosol samples collected at three urban (Stockton, Fresno,
  3      Bakersfield) and three non-urban (Crows Landing, Fellows, Kern Wildlife Refuge) locations
  4      (Chow et al.,  1993).  The VAQS data indicate the federal 24-h PM10 standard of 150 ^g/m3
  5      was exceeded at four out of the six sites and for reasons which differ by season and  by
  6      spatial region of influence.  The annual average source  contributions to PM10 at Bakersfield,
  7      the site with the highest annual average, were 54% from primary geological material, 15%
  8      from secondary  ammonium nitrate,  10 % from primary motor vehicle exhaust,  8% from
  9      primary construction, the remaining 4% is unexplained.  The results of the source
10      apportionment at all sites show that geological contributions dominate in summer  and fall
11      months, while secondary ammonium nitrate contributions derived from direct emissions of
12      ammonia and  oxides of nitrogen from agricultural activities and engine exhaust are largest
13      during winter months. (Chow et al., 1992).
14
15      6.5.7.2 Los Angeles-South Coast Air Basin
16           The Los Angeles basin is confined by the San Gabriel Mountains which limit the
17      ventilation during westerly winds. Intensive emissions from automotive  and industrial sources
18      produce the notorious Los Angeles smog as a secondary photochemical  reaction product of
19      primary emissions.  The map of the Los Angeles sub-region shows (Figure 6-83a) that the
20      highest PM10  concentrations are measured in the eastern half of the LA basin.
21           There has been a substantial reduction of sub-region average PM10 concentration from
22      60 to 37 pig/m3 from 1985 and 1993 (Figure 6-83b).  The seasonality of the basin averaged
23      PM10 concentration shows a 50% amplitude, with the peak concentration  (60 /-ig/m3) during
24      October and the lowest values (40 /xg/m3) during January through March (Figure  6-83c).
25      Hence, this fall peaked seasonality is similar to the fall peak over the San Joaquin Valley.
26           Some unique characteristics of the Los Angeles basin are depicted in Figure 6-84. It
27      shows that monitoring sites at different parts of the basin have markedly different seasonal
28      concentration  pattern. Hawthorne, near the Pacific Coast and Burbank in an inland valley
29      have the highest concentration in  late fall (November through January).  On the other hand,
30      Rubidoux in the eastern part of the basin exhibits the highest concentration in the  late
31      summer, July  through October.  It is likely, that the  main cause of different seasonalities can

        April 1995                               6-140      DRAFT-DO NOT QUOTE OR CITE

-------
      Average PM10 -1986-1993
                      AIRS PM10 Concentrantion Trends
                                  Los Angeles
                                              1986 1987 1988  1989  1990 1991 1992
                                        AIRS PM10 SEASONAL CONCENTRATION
                                                     Los Angeles
                                         80
                                         75
                                         70
                                         65
                                         60
                                         65
                                      O  60
                                     '5  45
                                         40
                                         35
                                         30
                                         25
                                         20
                                         15
                                         10
                                          5
                                          1985   Mir Apr Mty Jun Jul Aug S*p Oct Nov D*e
Figure 6-83.
Aerosol concentration pattern and trends at Los Angeles.
April 1995
                      6-141      DRAFT-DO NOT QUOTE OR CITE

-------
                                                                                           Nov
 LOS ANGELES, CA • Long Beach
 100
   LOS ANGELES, CA -Azuza
-A- •°PM10_AV 660376001HAWTHORNE
•B- =PM10_AV 060668001 RUBIDOUX
--(- =PM10_AV 060371002 BURBANK
         LOS ANGELES, CA - Rubldoux
        100
  1984   Mar   May  Jul    Sep  Nov
 rJT-PMIOJW 060374002 LONG BEACH
 -Q- -PM26_AV 060374002 LONG BEACH
 •Jr -PMC.AVG 06037*002 LONG BEACH
 1984   Mar   May  Jul   Sep  Nov
I TJr-PM10_AV 060370002 AZUSA      |
 -0- -PM25.AV 060370002 AZUSA
 -4- -PMC.AVG 060370002 AZUSA
        1984  Mar  May  Jul   Sep   Nov
       & -PM10.AV0606S8001 RUBJOOUX
        -Q- "PM26_AV 060668001 RUBIDOUX
        -4- "PMC_AVG 060668001 RUBIDOUX
Figure 6-84.        Fine, coarse and PM10 concentration pattern near Los Angeles.
April  1995
             6-142       DRAFT-DO NOT QUOTE OR CITE

-------
  1      be found in the seasonally varyingmeteorological, transport, and chemical transformation
  2      pattern, rather than of emissions.  The role of coarse and fine particles in the Los Angeles
  3      basin is also illustrated in Figure 6-84.  At Long Beach, near the coast (adjacent to
  4      Hawthorne), the concentration of fine particles dominate the PM10 during the November
  5      through February  winter season (40 to 50 /*g/m3).  Coarse particles at Long Beach are
  6      constant throughout the year at about (20 ^g/m3).  The size segregated aerosol samples at
  7      Azusa and Rubidoux  in the eastern part of the basin, both show a PM10 peak during
  8      September through October, although the concentration are higher at Rubidoux. At both
  9      sites fine and coarse particles contribute  roughly equally to the high PM10 concentrations.
 10      Thus, the PM10 aerosols over the smoggiest parts of the Los Angeles basin  are not dominated
 11      by fine secondary aerosols but contributed by both fine and coarse particles.
 12            The Los Angeles smog has been subject of extensive spatial, temporal, size and
 13      chemical composition studies even before the late 1960s (J. Colloid Interface Sci. (1972)
 14      volume 39, Hidy et al., 1980).  More recently the LA aerosol characteristics have been
 15      further elucidated  by  Southern California Air Quality Study (SCAQS) (Watson et al., 1994;
 16      Chow etal., 1994;   and other SCAQS studies).
 17            PM2.5 constituted one-half to two-thirds of PM10 at all sampling  sites.  PM10 mass
 18      concentrations were highest during the fall and were dominated  by PM25.  Nitrate, sulfate,
 19      ammonium, and organic and elemental carbon were the most abundant  species in the PM2 5
 20      fraction.  The coarse  particle fraction was composed of soil-related elements (e.g.  aluminum,
 21      silicon, calcium, iron) at the inland sites  and with marine-related elements (e.g. sodium,
 22      chloride) at the coastal sits. Average concentrations for most chemical compounds were
 23      higher during the fall than during the summer, except for sulfate which was more abundant
 24      in summer. PM2 5 nitrate  and ammonium concentrations were negatively biased for daytime
 25      samples compared to  nighttime samples,  consistent with diurnal  changes in temperature and
 26      the effect of these changes on the equilibrium between paniculate ammonium nitrate and
27      gaseous ammonia and nitric acid. (Chow et al., 1994).
28            In situ, time resolved analysis for  aerosol organic and elemental  carbon in Glendora,
29      CA (Turpin et al., 1990) showed strong diurnal variations with peaks occurring the daylight
30      hours.  Comparison of the diurnal profile of organic carbon with those  of elemental carbon
31      providedevidence for  the secondary formation of organic aerosol in the  atmosphere.  Turpin

        April 1995                               6-143      DRAFT-DO NOT QUOTE OR CITE

-------
 1      et al., 1991 observed that secondary organic aerosol appears to have contributed roughly half
 2      of the organic aerosol in Pasadena during midday summer conditions.
 3            Turpin and Huntzicker (1991) also found that the organic and elemental carbon
 4      concentrations exhibit strong diurnal variations.  Peak concentrations occur during the
 5      daylight hours in the  summer and at night in the fall.  The maximum concentrations observed
 6      in the fall. The maximum concentrations observed in the fall (maximum total carbon,
 7      88 /ig/m3) were two to three times higher than the summer maxima (maximum total carbon,
 8      36 Aig/m3).
 9            Gaseous nitric  acid and fine particulate nitrate at Claremont, CA (Pierson and
10      Brachaczek,  1988) both showed pronounced (~ 10 fold) diurnal variations, however coarse
11      particles showed little diurnal variations.
12            Wolff et al., 1991 measured the smog aerosol pattern at Claremont, CA and Long
13      Beach, CA, in the eastern and western Los Angles basin, respectively.  Claremont's air
14      quality during the summer was characterized  by high concentrations of photochemically
15      produced pollutants including ozone, nitric acid,  particulate nitrate, and particulate organic
16      carbon (OC). The highest concentration of these species were experienced during the
17      daytime  sampling period (0600 to 1800) and were associated with transport from the western
18      part of the basin.  Long Beach' air quality during the fall was characterized by frequent
19      periods of air stagnation that resulted in high concentrations of primary pollutants including
20      PM10, OC and elemental carbon (EC) as well as particulate nitrate.  Night -time levels of
21      most constituents exceeded day tune levels due to poorer night-time dispersion conditions.  At
22      Claremont, OC and nitrate compounds accounted for 52%  of PM10 , while at Long Beach
23      they accounted for 67% of PM10.  On the average, there appears to be sufficient particulate
24      ammonium to completely neutralize the nitrate and acidic sulfates.
25            Careful size distribution measurements  in the Los Angeles basin (John et al., 1990)
26      shed light on the size spectrum dynamics for ammonium, sulfate and nitrate.  Three modes,
27      two submicron and one coarse, were sufficient to fit all of the size distributions.  The
28      smallest  mode, at 0.2 ±0.1 /ma, aerodynamic diameter, is probably a condensation mode
29      containing gas phase reaction products.  A larger mode 0.7+0.2 /mi, is defined as a droplet
30      mode. Most of  the inorganic particle mass was found in the droplet mode.  The observed
31      condensation and droplet modes characterize the  overall size distribution in the 0.1 to 1.0 /mi

        April 1995                              6-144      DRAFT-DO NOT QUOTE OR CITE

-------
  1      range, previously described by Whitby and coworkers as a single accumulation mode
  2      (Whitby et al., 1972; Whitby, 1978). Wall et al., 1988 also found that in September 1985 at
  3      Claremont,  CA fine particle nitrate was associated with ammonium, while coarse mode
  4      nitrate was associated with both ammonium and sodium. Sulfate was primarily in two
  5      submicrometer modes.  Strong acid was associated with the smaller sulfate mode.
  6            A clear demonstration of the effect of relative humidity and aerosol loading on
  7      atmospheric sulfate size distributions is given by Hering and Friedlander, 1982.   Days of
  8      high relative humidity and aerosol loading correspond to high mass median diameters
  9      (0.54+0.07 jitm) for the sulfate  while low relative humidity and low aerosol loadings
10      correspond to small mass median diameters (0.2±0.02 /mi).  According to their
11      interpretation, the large (0.54 /mi) sulfate particles resulted from aqueous phase  reactions of
12      SO2. The fine (0.2 /mi) sulfate resulted from homogeneous gas phase reactions leading to
13      the nucleation of sulfuric acid particles.
14            McMurry and Stolzenburg, 1989 provide evidence that Los Angeles smog aerosols are
15      externally mixed.  Monodisperse ambient aerosols were often found to split into
16      nonhygroscopic (nowater uptake) and hygroscopic portions when humidified.  An average of
17      30% of the particles in the 0.2 to 0.5 /mi range were nonhygroscopic.  However, the
18      proportion of particles that were nonhygroscopic varied considerably from day to day and on
19      occasions was  70 to 80% of the particles.  The data show that for the hydrophilic aerosol,
20      the larger particles (0.4 to 0.5 /mi) grew more  when humidified than did smaller particles
21      (0.05 to 0.2 /mi).
22            Murray and Zhang, 1989 reported the size distribution of  ambient organic and
23      elemental carbon near the Grand Canyon and in the Los Angeles basin.  Virtually all of the
24      carbon was  found in the submicron range, some below 0.1 /mi.  However, positive sampling
25      artifacts  for sub O.lptm organics were considered  significant.
26           At the Grand Canyon National Park, Zhang et al., 1994, showed that sulfates and
27      carbonaceous particles were the major contributor to PM2 5 particle scattering during the
28      three winter months and that their contributions were comparable. Scattering by  nitrates and
29      soil dust was typically a  factor of five to ten smaller.  The low pressure  impactor
30      measurements also showed that sulfur size distributions varies considerably (0.07 to
31      0.66 /mi).

        April 1995                               6-145      DRAFT-DO NOT QUOTE OR CITE

-------
  1            Size distributions of aerosol phase aliphatic and carbonyl groups at Claremont, CA
  2      (Pickle et al., 1990) showed maxima in the 0.12 to 0.26 mm and the 0.5 to 1.0 mm size
  3      functions.  For the aliphatic carbon absorbency, the ambient samples generally showed
  4      maxima in the 0.076 to 0.12 mm size fraction, the authors attribute the carbonyl absorbance
  5      almost entirely attributed to products of atmospheric reactions and the aliphatic absorbencies
  6      in particles smaller than 0.12 mm to automotive emissions.
  7            Cahill et al.,  1990 found that the sulfate aerosol size at Glendora, CA is smaller,
  8      0.33 /mi (MMD) during clear days  compared to 0.5 /mi on smoggy days.
  9            The size distributions of organic nitrate groups in ambient Los Angeles aerosol were
10      typically bimodal (Mylonas et al., 1991).  During periods of high photochemical activity, the
11      maxima in the mass loadings were in the 0.05 to 0.075 /mi and the 0.12 to 0.26 /mi size
12      fractions.  During periods of low-moderate ozone concentrations in the distributions were
13      shifted to slightly larger sizes, with maxima appearing in the 0.075 to 012 /mi and the 05 to
14      1.0  /mi size fractions.  A principal component analysis of the organonitrate loadings
15      revealed strong correlations with ozone concentrations and with aerosol phase carbonyl
16      loadings.
17            The analysis of coarse particles in Claremont, CA (Noll et al.,  1990) show that the
18      coarse particle mass could be divided into two categories: material that was primarily of
19      crustal origin (Al, Ca, Fe,  and Si) and material that was primarily of anthropogenic origin
20      (Cd, Cu, Mn, Ni, Pb, and  Zn).  The mass of crustal material varied between 15 and 50%  of
21      the total coarse mass, while the mass of anthropogenic material was < 1%.
22            Chow et al., 1992 also conducted a neighborhood-scale study of PM10 source
23      contributions in Rubidoux,  CA elucidating the role of local soil dust.
24            The daily frequency  distribution of the chemical components of the Los Angeles
25      aerosol measured over a 1-year period were approximately lognormal  (Kao and Friedlander,
26      1994).  For nonreactive aerosol components, the geometric standard deviation (GSD) is
27      nearly constant at 1.85±0.14 even for components from different source types.  An apparent
28      bimodal frequency distribution for sulfates probably corresponds to the two differing reaction
29      pathways by  which gas-to-particle conversion occurs.  However, the bimodal sulfate
30      distribution function was not found at other Los Angeles sites (Kao and Friedlander, 1995).
31      The authors suspect a relationship between GSD and the level of complexity of the stochastic

        April 1995                               5.145      DRAFT-DO NOT QUOTE OR CITE

-------
  1      physical and chemical processes affecting the distributions of the individual species.  They
  2      also point out that the chemical composition of the probable Los Angeles aerosol that
  3      corresponded to the peak in the (nearly) lognormal frequency distribution of the total mass is
  4      lower than he simple average chemical composition.
  5           Twenty-four year (1958 to 1982) average elemental C concentrations at seven
  6      monitoring sites in the Los Angeles area are estimated to range from 6.4 /ng/m3 at downtown
  7      Los Angeles to 4.5 /*g/m3 at West Los Angeles.  (Cass et al., 1984). At most monitoring
  8      sites studied, elemental C concentration were lower in recent years than during the late 1950s
  9      and early 1960s.
 10
 11
 12      6.6   CHEMICAL COMPOSITION OF PM AEROSOLS  AT URBAN AND
 13            NON-URBAN SITES
 14           This section summarizes available data for the  composition of atmospheric particles in
 15      suburban, urban, and a few rural areas for comparison purposes.  Emphasis  has been placed
 16      on the Harvard six-city study and the inhalable particulate network (1980-1981).  However,
 17      data for fine particle mass and elemental composition were available from these studies.
 18      Data for sulfate, nitrate, and elemental and organic carbon content are included from other
 19      studies to provide an overview of the chemical composition of the  atmospheric aerosol in the
 20      United States.  Tables, presented in Appendix 6A,  provide relatively detailed representation
 21      of atmospheric properties of aerosols  to which U.S. populations are exposed. Unfortunately,
 22      data this complete are generally collected over limited time periods and are not of sufficient
 23      duration to be useful for most epidemiological investigations. The tables do, however,
 24      provide insights as to the types of information that  could be collected as part of future
 25      monitoring efforts in support of human exposure  investigations.
26           A summary of all the aerosol sampling studies  included in this compilation is given in
27      Table 6A-la and 6A-lb.  Sampling studies have been grouped by geographical region
28      roughly corresponding to the eastern,  central and western U.S.   Data are tabulated for the
29      PM2.5 (d <2.5 pm),  the coarse fraction of PM10  (2.5 ^m  < d < 10 /*m) and PM-10 (d
30      < 10 jum) size fractions of the ambient aerosol in Tables 6A-2a, 6A-2b, and  6A-2c.
31      Compositional data for all size fractions were broken  down into the following major
32      components:  sulfate, represented here as ammonium  sulfate, (NH4)2SO4; carbon, as organic
        April 1995                              6-147      DRAFT-DO NOT QUOTE OR CITE

-------
 1     carbon (Cv) which has been multiplied by a factor of 1.4 to account for the presence of
 2     oxidized species),  and elemental carbon (Ce); nitrate as NO3"; and remaining trace elements
 3     which have been determined by XRF and or INAA.  The masses of the trace elements from
 4     sodium through lead have been calculated by assuming they are in their most stable forms for
 5     conditions at the earth's surface. Reconstructed masses calculated in this way are  shown by
 6     the entry, Sum, along with measured masses, and the ratio of the two are shown  at the
 7     bottom of the individual summaries for each size fraction.  Not all compositional  categories
 8     were measured in  the studies included in the Tables.  The data are shown in graphical form
 9     in Figures 6-85a, 6-85b, and 6-85c.
10            As can be seen from inspection of Figure 6-85a for the eastern U.S., sulfate is the
11     major identified component of mass for fine particles (46.9%), followed by  carbon (24.8%),
12     minerals (4.3%), and nitrate (1.1%). However, this last inference is based on only two
13     studies in which nitrate was measured.  Coarse particles are seen to consist mainly of mineral
14     forming elements (51.8%) and sulfate (6.7%).  Not enough data were available to determine
15     abundances of carbon species and nitrate in the coarse fraction.  A sizable fraction of both
16     the fine (23.0%) and coarse (41.5%) particle mass is shown as unknown.  This unknown
17     mass is assumed to be mainly water, either bound as  water of hydration or associated with
18     hygroscopic particles.  A small fraction of the mass, especially in the coarse fraction, may be
19     present as carbonates. Carbonates are difficult to quantify, in part because of artifact
20     forming reactions on filters with atmospheric CO2. Stable carbonates can be identified by
21     SEM, especially in regions where they are known to  represent a substantial  fraction of soil
22     composition.
23            Fine particles in the central U.S. (Figure 6-85b) are seen to consist mainly of sulfate
24     (37.9%)  and minerals (9.4%) and elemental and organic carbon (66.1%) abundances.  The
25     reconstructed mass percentages sum to more than 100%.  This is probably because  of an
26     overestimation of the carbon content which was based on only a few samples collected during
27     winter in woodsmoke impacted areas.  Coarse particles were found to consist mainly of
28     minerals (62.8%), sulfate (4.2%) and an unknown fraction (33.0%).   No nitrate or carbon
29     data were available for the coarse fraction from the studies in the central U.S.  However,
30     during the spring and fall pollen becomes a significant fraction of the coarse particle
31     composition.

       April 1995                               6-148      DRAFT-DO NOT QUOTE OR CITE

-------
                               EASTERN U.S.



                                PM2.5 Mass Apportionment



                                            -Minerals (4.3%)
                                        • i r
                      Unknown (23.0%)






                        EC (3.9%)





                       OCxl.4 (20.9%)-


                               Nitrate (1.1%)-


                                  Nitrate based on 3 studies.
       •NH42SO4 (46.9%)
                                Coarse Mass Apportionment
                   Unknown (41.5%)
                                                      Minerals (51.8%)
                          NH42S04 (6.7%)-



                            Insufficient Nitrate, OC, and EC data available.
                                 PM10 Mass Apportionment
                     Unknown (29.2%)
                        EC (3.3%)


                       OCx1.4 (8.5%)
 ^-Minerals (19.6%)




       Nitrate (1.2%)
                                                 -NH42SO4 (38.2%)




                                  Nitrate based on 2 studies.
Figure 6-85a.      Mass apportionment:  Eastern U.S.
April 1995
6-149       DRAFT-DO NOT QUOTE OR CITE

-------
                                         CENTRAL U.S.


                                          PM2 5 Mass Apportionment


                                        EC(11.1%)-\ 	^-Minerals (9.4%)
                              OCx1.4 (55.0%)
                                                                NH42SO4 (37.9%)
                                                          -Nitrate (10.0%)


                                           Reconstructed sum =123%
                                          Coarse Mass Apportionment
                             Unknown (33.0%)
                              NH42SO4 (4.2%)-\ffl
                                                                      (62.8%)
                                      Insufficient Nitrate, OC, and EC data available.
                                           PM10 Mass Apportionment


                                        Unknown (1.4%)-i
                                                             V—Minerals (35.8%)
                               OCx1.4 (29.6%)
                                   EC (5.0%)

                                                            -Nitrate (4.5%)
                                                  T	-"
                                       Sulfate (23.7%)-


                                 Nitrate based on 3 studies; OC and EC based on 4 studies.
Figure 6-85b.       Mass apportionment:  Central U.S.
April 1995
6-150       DRAFT-DO NOT QUOTE OR CITE

-------
                                       WESTERN  U.S.




                                        PM2.5 Mass Apportionment




                                     Unknown (1.3%)-
                                   EC (14.7%)
                             OCx1.4 (38.9%)
                                                        Minerals (14.6%)
                                                             NH42S04 (14.6%)
                                                           Nitrate (15.7%)
                                       Coarse Mass Apportionment
                             Unknown (27.0%)
                           NH42SO4(3.1%)
                                                        r--Minerals (70.0%)
                                   Insufficient Nitrate, OC, and EC data available.
                                        PM10 Mass Apportionment




                                       EC (5.0%)-
                           OCx1.4(29.6%)
                                                          V—Minerals (35.8%)
                                 NH42S04 (23.7%)--      ^-Nitrate (4.5%)




                                        Reconstructed sum = 106%
Figure 6-85c.       Mass apportionment:  Western U.S.
April 1995
6-151       DRAFT-DO NOT QUOTE OR CITE

-------
 1            While gross fine particle composition appears to be broadly similar between the
 2      eastern and central U.S. on the basis of the few studies available, the fine particle
 3      composition is seen to be distinctly different in the western U.S. (Figure 6-85c).  Elemental
 4      plus organic carbon species (53.6%) are the major identified component of mass, instead of
 5      sulfate (14.8%), and minerals and nitrate account for a larger fraction of total mass. While
 6      minerals are seen to account for most of the coarse particle mass (70.0%), insufficient data
 7      were available for elemental and organic carbon species to estimate their contributions to the
 8      coarse mass.  Table 6A-3 shows a comparison of selected ratios of mass components for
 9      each of the three geographical regions of the U.S.
10           Many of the  studies listed in Table 6A-3 involved data collected at more than one site
11      within an airshed.  Information about the variability of particle mass within an airshed can
12      yield information about the nature of sources of the particles. The variability of mean
13      concentrations measured at multiple sites within a study area is used as a measure of the
14     intersite variability in fine particle composition and is shown in Table 6A-4.
15            Data for the chemical composition of the ambient aerosol has been summarized from
16     the Harvard six-city study, the inhalable particle (IP) network and for a number of other
17     studies around the country.  As can be seen from  inspection of the tables, data are not
18     available to characterize the carbon or nitrate content of the ambient aerosol for many of the
19     studies listed.  Over the past 15 years woodstove emissions have become a significant
20     contributor to fine particle mass during the winter (Stevens 1990).
21
22
23     6.7   ACID AEROSOLS
24     6.7.1   Introduction
25            Acid aerosols are secondary pollutants formed primarily through oxidation of sulfur
26     dioxide (SO2), a gas emitted by the combustion of fossil fuels.  Oxidation of SO2 forms
27     sulfate (SO4=),  the major component of acid aerosols.  Sulfate is formed to a lesser extent
28     through the oxidation of sulfur species (H2S and CH3SCH3) from natural sources.  The
29     oxidation of SO2 occurs through a series of heterogeneous (gas-particle) or homogeneous (gas
30     or aqueous) phase oxidation reactions that convert SO2 to sulfuric acid (H2SO4) particles.
31     The sulfate species are typically expressed in terms  of total 804, with the acidic fraction

       April 1995                               6-152      DRAFT-DO NOT QUOTE OR CITE

-------
  1      expressed in terms of titratable H+ and referred to as aerosol strong acidity. H+ is found in
  2      the fine particle size fraction (aerodynamic diameter (Dp)  < 1.0 /mi) (Koutrakis and Kelly,
  3      1993).  Although recent research has shown a high correlation between SO4 and acidity,
  4      data from summertime sampling have shown that SO4= is not always a reliable predictor of
  5      H+ for individual events at a given site (Lipfert and Wyzga, 1993).
  6            A major determinant of the lifetime of H+ in the atmosphere is the rate of
  7      neutralization by ammonia (NH3).  Ammonia reacts  with H2SO4 to form ammonium sulfate
  8      [(NH4)2SO4] and ammonium bisulfate (NH4HSO4).  The major sources of ammonia in the
  9      environment are animals and humans (Fekete and Gyenes, 1993). The then current state-of-
 10      knowledge regarding acid aerosols was reviewed by EPA  in 1989 (U.S. Environmental
 11      Protection Agency, 1989) and Spengler et al. in 1990 (Spengler et al., 1990).
 12
 13      6.7.2   Geographical Distribution
 14            In North America, ambient concentrations of H+ tend to be regional in nature with
 15      the highest concentrations found in the northeastern United States and southwestern Canada
 16      Spengler et  al. (1990) have collected  information of  maximum values of SO^ and H+ found
 17      across the U.S. and southern Canada. This information is shown in Table 6-3. Recent
 18      research has shown that regional transport is important to acid sulfate concentrations, as
 19      elevated levels of ambient H+ were measured simultaneously during a regional episode at
20      multiple sites located from Tennessee to Connecticut (Keeler et al.,  1991). It is commonly
21      believed that the source region for most of the H+ precursors (primary inorganic pollutant
22      gases — SO2 and NOX) is the Ohio River Valley (Lioy et al., 1980). The conversion of the
23      primary gases to secondary pollutants takes place as  the prevailing winds carry the precursors
24      from the source region, northeastward to the northeastern United States and southwestern
25      Canada.  This type of northeasterly wind flow occurs on the backside (western side) of mid-
26      latitude anti-cyclones (high pressure systems).
       April 1995                              6-153      DRAFT-DO NOT QUOTE OR CITE

-------
   TABLE 6-3. MAXIMUM SOj AND H+ CONCENTRATIONS MEASURED IN
 NORTH AMERICAN CITIES.  H+ CONCENTRATIONS EXPRESSED AS "H2SO4"
   EQUIVALENTS. "SC" INDICATES SEMI-CONTINUOUS MEASUREMENTS
Location
Lennox, CA
Smoky Mountains
High Point, NJ
Brookhaven, NY
Tuxedo, NY
St. Louis, MO
St. Louis, MO
Los Angeles, Ca
Harriman, TN
Watertown, MA
Fairview Lake, NJ
Warren, MI
Whiteface Mt., NY
Toronto, Ontario, Canada
Allegheny Mt., PA
Laurel Mt., PA
Harriman, TN
St. Louis, MO
Topeka, KS
Watertown, MA
Steubenville, OH
Portage, WI
Kanawha Valley, WV
Dunville, Ontario, Canada
Hendersonville, TN
Livermore, CA
Morehead, KY
Monroeville, PA
Pembroke, ON, Canada
Springdale, AR
Newtown, CT
Allegheny Mt. , PA
Uniontown, PA
State College, PA
Philadelphia, PA
Pittsburgh, PA
Sample Duration (h)
2-8
12
6
3
1-12
SC
SC
12
SC
SC
SC,4
24
24
8,16
7,10
7,10
24
24
24
24
24
24
24
24
24
24
24
24
24
24
24
12
12,24
12
24
6,24
Maximum
SC£ Oig-m-3)
18
17
37
24
41
25
43
10
47
31
27
37
59
75
45
56
28
40
14
23
56
33
46
31
23
9
23
42
29
11
26
33
52
47
39
27
Concentration
H2SO4 (pig-m"3)
0.1
10
18
10
9
7
34
3
18
14
12
9
14
19
31
42
14
6
3
9
18
4
22
15
11
2
14
18
14
2
8
20
39
25
9
15
April 1995                      6-154    DRAFT-DO NOT QUOTE OR CITE

-------
  1      6.7.3   Spatial Variation (City-scale)
  2            A study of acid aerosols and ammonia (Suh et al., 1992) found no significant spatial
  3      variation of  H+ at Uniontown, Pennsylvania, a suburb of Pittsburgh. Measurements at the
  4      central monitoring site accounted for 92% of the variability in outdoor concentrations
  5      measured at various homes throughout the town. There was no statistical difference (p >
  6      0.01) between concentrations of outdoor H+ among five sites (a central site and four satellite
  7      sites) in Newtown, Connecticut (Thompson et al.,  1991).  However,  there were differences
  8      in peak values which were probably related to the  proximity of the sampling sites to
  9      ammonia sources.  These studies suggest that long-term averages should not substantially
10      differ across a suburban community, although peak values may differ significantly.
11            In small suburban communities outdoor concentrations of H+ are fairly uniform,
12      suggesting that minor differences in population density do not significantly affect outdoor H+
13      or NH3 concentrations (Suh et al.,  1992).  In urban areas, however both  H+ and NH3 exhibit
14      significant spatial variation. Waldman et al. (1990) measured ambient concentrations of H+,
15      NH3, and SOJ at three locations in metropolitan Toronto.  The sites, located up  to 33 km
16      apart, had significant differences  in outdoor concentrations of H+.  Waldman and co-workers
17      reported that the sites with high NH3 measured low H+ concentrations.  However, the
18      limited number of sampling sites  did not allow for a conclusive determination of  the
19      relationship between population density, ammonia concentrations, and concentrations of acid
20      aerosols.
21            An intensive monitoring study has been conducted during the summers of  1992 and
22      1993 in Philadelphia (Suh et al.,  1994a). Twenty-four hour measurements of aerosol acidity
23      (H+) sulfate  and NH3 were collected simultaneously at 7 sites in metropolitan Philadelphia
24      and  at Valley Forge, 30 km northeast of the city center.  The researchers reported that SO^
25      was evenly distributed throughout the measurement area but H+ concentrations varied
26      spatially within metropolitan Philadelphia. This variation was related to local NH3
27      concentrations and the local population density (Figure 6-86).  The amount of NH3 available
28      to neutralize  H+ increased with population density, resulting in lower H+ concentrations in
29      more densely populated areas.  The extent of the spatial variation in H+  concentrations did
30      not appear to depend on the overall H+ concentration.  It did, however,  show a strong
31      inverse association with local NH3 concentrations.

        April 1995                               6-155      DRAFT-DO  NOT QUOTE OR CITE

-------
       0.6 r
       0.4
       0.3
        0.2
        0.1
        0.0
_ 3
5 r
            120
                              90
                           CO
                           09
                           O
                           E
                          .5.  60
                          o
                          CO
                              30
                                                                         SO
                                                                         NH
                                 0          5000        10000       15000      20000

                                       POPULATION DENSITY (persons/sq.mile)
      Figure 6-86. Mean air pollutant concentrations for days when winds were from the
                  southerly direction plotted vs. population density.  The solid line
                  represents H+ concentrations; the long dashed line represents SO|"
                  concentrations; the dashed and dotted line represents the ration of H+ to
                  SO|" levels; and the dotted line represents NH3 concentrations.  All data
                  collected in Philadelphia, PA, during the summers of 1992 and 1993.
                  Figure adapted from Suh et al. (1994a).
1     6.7.4   Spatial Variation (Regional-scale)

2          Recent evidence has shown that meteorology and regional transport are extremely

3     important to acid sulfate concentrations. Lamborg et al. (1992) measured H+ concentrations

4     to investigate the behavior of regional and urban plumes advecting across Lake Michigan.

5     Results suggested that aerosol acidity is maintained over long distances (up  to 100 km or

6     more) in air masses moving over large bodies of water.  Lee et al. (1993) reported that H+
      April 1995
                            6-156
                                    DRAFT-DO NOT QUOTE OR CITE

-------
  1     and SO 4 concentrations measured in Chicago over a year were similar to levels measured in
  2     St. Louis. In an analysis of acid sulfate concentrations measured at Pittsburgh, State
  3     College,  and Uniontown, Pennsylvania, Burton et al. (1995) reported high correlations for
  4     H+ between all three locations.  The three locations are separated by large distances
  5     (approximately 60 to 240 km) and have vastly different population densities.
  6
  7     6.7.5    Seasonal Variation
  8          An analysis of results  from Harvard's 24-city study (Thompson et al., 1991), which
  9     measured acid aerosols concentrations at 8 different sites across North America each year
 10     during  a  three year period,  revealed that the summer H+ mean concentrations were
 11     significantly higher than the annual means at  all sites.  The  results not only showed that at
 12     the sites with high H+ concentrations approximately two-thirds of the aerosol acidity
 13     occurred from May through September (Figure 6-87). Little or no seasonal variation was
 14     observed at sites with low acidity.  These  findings were  supported by those of Thurston et al.
 15     (1992)  in which H+ concentrations measured at Buffalo, Albany, and White Plains, New
 16     York were found to be highest during the  summertime.  Thurston and co-workers also
 17     reported that moderate concentrations of H+ could occur during non-summer months.
 18
 19     6.7.6    Diurnal Variation
 20          Evidence exists  of a distinct diurnal pattern in outdoor H+ concentrations.  Wilson
 21      et al. (1991) examined concentration data for H+, NH3,  and 804 from the Harvard 24-City
 22     Study for evidence of diurnal variability (Figure 6-88).  This investigation found a distinct
 23      diurnal pattern for H+ concentrations and  the H+/SOJ ratio, with daytime concentrations
 24      being substantially higher than nighttime levels. Both H+ and  SO4 concentrations peaked
 25      between noon and 6:00 pm.  No such diurnal variation was  found for NH3. Wilson and co-
 26      workers concluded that the diurnal variation in H+ and SO4= was probably due to
27      atmospheric mixing.  Air containing high concentrations  of H+ and SOJ mix downward
28      during daylight hours when  the atmosphere is unstable and well-mixed. During the night,
29      ammonia emitted from ground-based sources neutralize the acid in the  nocturnal boundary
30      layer, the very stable  lower  part  of the atmosphere, but a nocturnal  inversion prevents the


        April 1995                               6-157     DRAFT-DO NOT QUOTE OR CITE

-------
     I
      o
      E
      c
     1
     a
     i
     w
     o
          180
150-
          120-
 90-
 60 J
           30-
• MOREHEAD
0 HENDERSONVILLE
m PENN HILLS
0 DUNNVILLE
D NEWTOWN
• PEMBROKE
H SPRINGDALE
m LIVERMORE
               JAN   FEE  MAR  APR  MAY  JUN   JUL  AUG   SEP   OCT   NOV  DEC
                                             MONTH
      Figure 6-87. Average monthly aerosol strong acidity for Year 1 sites of the Harvard 24-
                  city study.
      Source: Thompson et al.,  1991
1     ammonia from reacting with the acid aerosols aloft.  Then in the morning as the nocturnal
2     inversion dissipates, the acid aerosols mix downward again as the process begins anew.
3     Spengler et al. (1986) also noted diurnal variations in sulfate and sulfuric acid concentrations
4     and suggested atmospheric dynamics as the cause.
      April 1995
                                  6-158
                                       DRAFT-DO NOT QUOTE OR CITE

-------
                       Harriman, TN, 8/12/84-8/19/84
                       20   40    60   80   100   120  140   160  180   200
            T3
            CO
            o
            £
            CO
            I
            J3>
            o
            c
                 Vertical lines drawn at noon.
3.2
3.0
2.8
2.6
2.4
2.2
2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
                           Harriman, TN Average
                     August 12,1984-August 19,1984
 -O- Sulfate
 -*— Hydrogen Ion
                                     10  12  14
                                       Hour
                                16  18  20   22  24
Figure 6-88. Diurnal pattern of sulfate and hydrogen ion at Harriman, TN; (a) weekly
            pattern, (b) daily average.
April 1995
                       6-159
DRAFT-DO NOT QUOTE OR CITE

-------
 1     6.7.7   Indoor and Personal
 2                   Several studies have examined indoor concentrations of acid aerosols and
 3     personal monitoring.  Brauer et al. (1989) monitored personal exposures to particles
 4     (including acidic sulfates) and gases in metropolitan Boston in the summer of 1988, and
 5     compared these to measurements collected at a centrally located ambient monitor.  They
 6     found that personal concentrations of acidic aerosols and gases differed  significantly from
 7     those measured at the centrally located site.  Summer and winter concentrations of acid
 8     aerosols and gaseous pollutants also collected in Boston (Brauer et al., 1991) showed
 9     indoor/outdoor ratios of H+  to be 40-50% of the indoor/outdoor SO4~ ratio indicating
10     neutralization of the acid by  the higher indoor NH3 levels, which were reported.
11                   Indoor, outdoor, and personal acid aerosol monitoring was performed for
12     children living in Uniontown, Pennsylvania, during the summer of 1990 (Suh et al., 1992).
13     The indoor, outdoor, and personal measurements were compared to outdoor measurements
14     collected from a centrally located ambient monitor.  Personal concentrations were lower than
15     corresponding outdoor levels but higher than indoor levels.  Air conditioning was found to be
16     an important predictor of indoor H+, while NH3 was found to influence indoor and personal
17     H+ concentrations.  Similar  results were obtained in a study of the relationships between
18     indoor/outdoor concentrations of H+ and NH3 conducted in State College, Pennsylvania in
19     1991 (Suh etal., 1994b).
20                   In a study characterizing H+ concentrations  at child and elderly care facilities,
21     Liang and Waldman (1992) measured indoor and outdoor acid aerosol concentrations.
22     Results from this study showed that indoor/outdoor H+ and SO 4 ratios  were comparable to
23     those measured inside residential buildings.  Air conditioner use and indoor NH3
24     concentrations were again identified as important determinants of indoor H+ concentrations.
25
26
27     6.8   PARTICLE NUMBER CONCENTRATION
28     6.8.1   Introduction
29                   Recent work has suggested that ultrafine particles may be responsible for some
30     of the health effects associated with exposure to particulate matter (Section XX).  The
31     hypothesis for explaining a biological effect of ultrafine particles is based on the number,

       April 1995                               6-160      DRAFT-DO NOT QUOTE OR CITE

-------
 1     composition and size of particles rather than their mass (Seaton et al., 1995).  This has led to
 2     an interest in the number concentration of ambient particles.  This section examines data on
 3     particle number concentration and the relationship between particle number and particle mass
 4     or volume.
 5
 6     6.8.2   Ultrafine Particle Number-Size Distribution
 7                  In the context of ambient particles, the term ultra fine particles refers to those
 8     particles with diameters below about 0.1 /mi.  Ultrafine aerosol size distributions from an
 9     urban site at Long Beach, California (Karch et al., 1987), and from a background site in the
10     Rocky Mountains, Colorado (Kreidenwies and Brechtel, personal communication) are shown
11     in Figures 6-89 and 6-90. Both of these sets of data were obtained by electrical mobility
12     measurements.  For the urban aerosols of Long Beach, the geometric  mean number diameter
13     can vary from 0.012 /mi  to 0.043 /xm.  Some of the ultrafine distributions, such as that
14     shown for the 1,200 to 1,400 PST time period, are bimodal.  The number concentrations
15     were higher midday, as shown in Figure 6-91. For the background aerosols from Rocky
16     Mountains the geometric  mean diameter of the ultrafine aerosols was somewhat larger than
17     for Long Beach, with geometric mean diameters ranging from 0.047 to 0.075 jum for periods
18     without urban influence.  A bimodal character for the ultrafine distribution was also observed
19     for some measurements,  as seen in Figure 6-90.
20                   The contrast between urban and background ultrafine aerosol size distribtution
21     is demonstrated in Figure 6-92, where a change in the wind direction  brought transport from
22     an urban area to the background site at Rocky Mountains.  Within a 2-h period, the number
23     concentration increased from 850 cm"3 to  19,000 cm"3, an increase of more than  a factor of
24     20.  In contrast, the volume distribution increased by less than a factor of 5.  The number
25     geometric mean diameter decreased from 0.052 />un for the background aerosol to 0.024 /jm
26     for the urban influenced aerosol.  For the urban influenced size distributions,  over 96% of
27     the particle number was measured in particles below 0.1 /zm, while 80% of the particle
28     volume was associated with particles above that size.
       April 1995                               6-161      DRAFT-DO NOT QUOTE OR CITE

-------
                              Long Beach, CA
    0.00
        0.01

Particle Diameter (|im)
                                                                       0.10
     0.00
        0.01

Particle Diameter (|im)
                                                                      0.10
Figure 6-89. Aerosol number and volume size distributions from an urban site at Long
            Beach, CA.
April 1995
        6-162     DRAFT-DO NOT QUOTE OR CITE

-------
                                  Rocky Mountains, CO
                                                                    11/23/941304


                                                                    11/23/941804


                                                                    11/24/941205
      0  ^os-
      0.01
                                           0.1
                                Particle Diameter, Dp (|im)
   0.6 T



^ 0.5
CO

E
0 0.4
CO
   0.3 -f
Q.
Q
o>
° 0.2 +
   0.1  4
      0.01
                      11/23/941304


                      11/23/941804


                      11/24/941205
                                           0.1
                                Particle Diameter, Dp (urn)
 Figure 6-90. Aerosol number and volume size distributions from a background site in
             the Rocky Mountains, CO.
 April 1995
                                       6-163     DRAFT-DO NOT QUOTE OR CITE

-------
                      90,000 -I
                      80,000 -
                 |_   70,000 -
                 °   60,000 -
                 ° co" 50,000 •
                 w E
                 o> o 40,000 -
                 o —
                 t:
                 to
                 Q.
                 I
30,000 •
20,000 •
10,000 •
    0
                            12
                                      14
                         16
    18
Time of Day
20
                                                                             22
                                                                                        24
       Figure 6-91.  Number concentrations as a function of time of day at Long Beach, CA.
 l     6.8.3   Relation of Particle Number to Particle Mass
 2          In general, the majority of airborne particle mass is associated with particles above
 3     0.1 /mi, while the highest number concentration of particles is found in particles below
 4     0.1 /*m.  This can be seen in the recent data collected in the Los Angeles, CA shown in
 5     Figure 6-93.  As with the data of Whitby and  Sverdrup, the size distributions of Figure 6-93
 6     show data collected by several instruments.  Physical size distributions were measured with
 7     an electrical aerosol analyzer for particles between 0.01 and 0.4 /zm, and with a laser optical
 8     particle counter for particles between 0.14 and 3 jum (Eldering and Cass, 1994).
 9     Additionally, Berner (John et al., 1989, 1990) and MOUDI impactors (McMurry, 1987)
10     were used to measure the size distribution of inorganic ion species and carbonacous species,
1 1     and these data have been combined to give a total mass distribution from which the number
12     distribution has been calculated assuming an effective aerosol density of 1.6 g/cm3,  and
       April 1995
                         6-164
      DRAFT-DO NOT QUOTE OR CITE

-------
                                Rocky Mountains, CO
60,000
                                                                 12/25/941452

                                                                 12/25/941550

                                                                 12/25/941648
     0.01
                              Particle Diameter, Dp ((im)
   3.0 T

   2.5 -
	 D —
	 O 	
- 12/25/9414:53
12/2b/94 lb.4b
— 12/25/9416:53
      0.01
    0.1
                                Particle Diameter, Dp (|im)
  Figure 6-92.  Number and volume size distributions at the Rocky Mountain site showing
              an intrusion of urban air.
  April 1995
6-165
DRAFT-DO NOT QUOTE OR CITE

-------
                               Downtown Los Angeles
    150,000 -,

_  125,000
co"
 E  100,000
          0.01
                           0.1                   1

                           Particle Diameter, Dp (urn)
                                                                        10
 CO
 p
 o
 CO
 E
 •o
 ^
 •o
O)
         0 -t-OMIMCMMMCiili
                               0.1                    1

                               Particle Diameter, Dp ((im)
                                                                     i
                                                                     10
                                           OPC
                                                         EAA
Figure 6-93. Number and volume size distributions from Los Angeles, CA, showing
            comparison of three measurement techniques.
April 1995
                                 6-166
DRAFT-DO NOT QUOTE OR CITE

-------
 1      assuming that the water associated with the aerosol is 15% of the measured dry particle mass
 2      (see McMurry and Stolzenburg, 1989).  The optical particle counter data were reduced using
 3      calibration with dioctyl sebacate aerosol, which has a refractive index (n = 1.45) more
 4      closely related to that of ambient aerosol than does polystyrene latex (n = 1.59) (Hering and
 5      McMurry, 1990).  No fitting has been applied to match the different size distributions in the
 6      region of overlap.
 7           Figure 6-93 shows the average of distributions collected over a six different days  in the
 8      fall of 1987 in downtown Los Angeles, as part of the Southern California Air Quality Study.
 9      Particle number distributions emphasize the ultrafine particles, or "nuclei" mode.  Volume
10      distributions place importance on 0.1 to  1 /xm particles which are associated with the
11      "accumulation" mode. For this average distribution 88% of the particle number is associated
12      with particle below 0.1 jiim, but 99% of the particle volume is from particle above that size.
13      Both the impactor and optical counter data indicate a weakly  bimodal character for the
14      accumulation mode aerosol.
15           For unimodal, log normal size distributions, the particle volume V is simply related  to
16      the particle number Nby the relation:
17
18
19      where D   is the geometric number mean diameter, and agis the geometric standard
20      deviation.  However, because of the multimodal character of ambient aerosol size
21      distributions, one does not expect this simple relationship to hold in the atmosphere.  The
22      relationship between particle number and particle volume was examined for data from the
23      Southern California Air Quality Study collected at Riverside, CA over 11  days in the
24      summer of 1987, and at downtown Los Angeles in the fall of 1987 using the  methods
25      described above.  As shown in Figure 6-94, particle number  concentrations are correlated
26      with the volume associate with particles below 0.1 /^m, but are not correlated with the total
27      fine particle volume.  Similar results are found for the data reported from Rocky Mountains,
28      CO and for the data reported by Whitby and Sverdrup (1980).
       April 1995                               6-167      DRAFT-DO NOT QUOTE OR CITE

-------
   CO
   0
160,000  T


140,000  --


120,000
   w  100,000  --
   CD

   E   80,000  --
   ^D   60,000  -•
   o

   rc   40,000  --
   O.
                            0
       0.00       2.00      4.00       6.00

                Volume <.1u,m  (iim3cm-3)
                                                   8.00
                                                    • Los Angeles


                                                    D Riverside


                                                    * Whitby Background


                                                    o Whitby Urban


                                                    A Rocky Mts.
160,000 -
140,000 -
CO
4- 120,000 -
E
0
,_ 100,000 -
CD
.a
E 80,000 -
D
z
CD 60,000 -
rt 40,000 1
a.
•
O

•
* *
• m •

•
• •
* n cj *
• • n
3 • D a U a

20.000 k LJnritin 7" tr a

• Los Angeles

D Riverside

A Whitby Background

o Whitby Urban
A Rocky Mts.

p QA D
r\ AA. 	 i- 	 	 > 	 — i 	 1
                       50        100       150       200

                     Volume < 2.5 (im  (u.m3 cm-3)
Figure 6-94. Relationship between particle number and particle volume; (a) volume

            <0.1 /on, (b) volume <2.5
April 1995
                                6-168
DRAFT-DO NOT QUOTE OR CITE

-------
  1      6.8.4   Conclusion
  2           The size distribution data for ultrafine aerosols in urban and continental backgroun
  3      regions have geometric mean diameters which vary from 0.01 to 0.08, with the larger values
  4      found in background regions. Particle number concentrations may vary from less than
  5      1,000/cm3 at clean, background sites to over  100,000/cm3 in polluted urban areas.  Particle
  6      number concentrations are dominated by the ultrafine, "nuclei"  mode aerosols.  In contrast,
  7      the volume (or mass) of fine particles is associated with particles above 0.1 yum, which are
  8      associated with the "accumulation" mode identified by Whitby and coworkers (Willeke and
  9      Whitby, 1975; Whitby and Sverdrup, 1980).  Particle number concentrations are correlated
10      with the volume of particles  below 0.1 ju,m.  The number concentration of ultrafine particles
11      results from a balance between formation and removal.  The rate of removal by coagulation
12      with accumulation mode particles will increase as the number (and mass and volume) of
13      accumulation mode particles  increases.  Therefore, a correlation between number and
14      accumulation mode volume or mass would not be anticipated. As expected no correlation is
15      found between the total number concentration and the total fine particle mass or volume.
16
17
18      6.9  AMBIENT CONCENTRATIONS OF ULTRA-FINE METALS
19      6.9.1   Introduction
20           Numerous pathways result in "ultra-fine" atmospheric aerosols, particles in the size
21      range around 0.1 m diameter.  These include both primary production processes, such as
22      combustion, and secondary processes involving gas-to-particle conversion and subsequent
23      growth by condensation and coagulation to larger particles.  There are  also  numerous
24      pathways that readily remove such particles from this mode, both by deposition and by
25      growth in size into the "accumulation mode", broadly defined as the maximum in particle
26      mass  or volume that normally occurs in the range of particle diameters between 0.2 and 0.7
11      m (Whitby et al., 1978).  The result is that in ambient conditions, the "ultra-fine" mode is
IS      generally indistinct or absent from mass or volume profiles of aerosol particles versus size.
19      However, in some situations  the ultra-fine mode can be the dominant size range for selected
50      components of the atmospheric  aerosol particles.  This is the case for metallic aerosols in


        April  1995                              6-169      DRAFT-DO NOT QUOTE OR CITE

-------
 1     which fine (Dp<2.5 m) concentrations can be dominated by the ultra-fine mode despite the
 2     strength of the processes that tend to remove particles from this mode.
 3           Removal processes are driven by the ability of ultra-fine particles to rapidly diffuse to
 4     surfaces, enhancing loss by deposition and by processes such as coagulation.  The high
 5     surface area of ultra-fine particles, which is 5 times greater than an equal mass of particles at
 6     0.5 m diameter, also enhances growth to the accumulation mode by forming nuclei for
 7     condensation of volatile species.  For these and other reasons, the mass of ultra-fine particles
 8     in the ambient atmosphere is generally much smaller than that of the accumulation mode,
 9     where removal rates of particles reach a minimum in non-cloud conditions. However, a
10     distinct ultra-fine mode below 0.1 m diameter has been observed in quasi-ambient samples
11     taken close to combustion sources, sometimes referred to as the "combustion mode" (Whitby
12     etal., 1978).
13           While there is consensus that ultra-fine metals are abundantly produced and emitted into
14     the atmosphere, there are not a lot of data on ambient concentrations of ultra-fine metals.
15     The few direct measurements available can be extended with some confidence using indirect
16     methods; i.e., from particle counting techniques that have size information but no chemical
17     information, or from filter collection methods that  have limited size information but detailed
18     compositional information.   Nevertheless, it is clear that more data on ultra-fine metals is
19     urgently needed to gain confidence in the spatial and temporal concentration profiles of this
20     key atmospheric component.
21
22     6.9.2   Formation  of ultra-fine particles
23           Combustion theory establishes that high temperature processes are generally required to
24     form ultra-fine metallic aerosols. Such processes are usually anthropogenic, although natural
25     fires, volcanic eruptions, and other such events can contribute to ultra-fine transition and
26     heavy metals in some circumstances.  Table 6-4, taken from Seeker (1990), gives the
27     temperature of formation of EPA-regulated metals  (Federal Register, 1986) as a function of
28     temperature, with and  without chlorine available in the combustion process.
        April 1995                               6-170      DRAFT-DO NOT QUOTE OR CITE

-------
          TABLE 6-4. REGULATED METALS AND THE VOLATILITY TEMPERATURE
                                          (SEEKER, 1990)
Metal
Chromium
Nickel
Beryllium
Silver
Barium
Thallium
Antimony
Lead
Selenium
Cadmium
Osmium
Arsenic
Mercury
With
Volatility
temp, (°F)
2935
2210
1930
1660
1560
1330
1220
1160
605
417
105
90
57
no chlorine
Principal
species
CrO2/CrO3
Ni(OH)2
Be(OH)2
Ag
Ba(OH)2
T1203
Sb2O3
Pb
SeO2
Cd
OsO4
As2O3
Hg
With 10%
Volatility
temp, (°F)
2930
1280
1930
1160
1660
280
1220
5
605
417
105
90
57
chlorine in waste
Principal
species
CrO2/CrO3
NiCl2
Be(OH)2
AgCl
BaCl2
T1OH
Sb203
Pb
SeO2
Cd
OsO4
As2O3
Hg
 1          Note the dramatic shift in temperature for several elements, including lead, for the
 2     chlorine-rich combustion scenario.  A similar process has been used to prevent lead from
 3     coating surfaces in internal combustion engines using leaded gasoline.  The process used
 4     chlorine and bromine-containing additives to form compounds such as PbBrCl which then
 5     leave the vehicles as ultra-fine aerosols.
 6          Numerous theoretical and laboratory studies have shown that the typical size of metals
 7     derived from combustion is ultra-fine, (Friedlander, S.K., 1977; Senior, et al.,  1982; Seeker
 8     et al.,  1990).  Analysis of particles from coal combustion by Natusch et al. (1974a, b)
 9     showed an additional aspect. There is a tendency for the condensing metal vapors to form
10     relatively uniform thickness surface coatings on more refractory particles present in the
11     combustion effluent stream.  If the particles upon which the metals coat themselves are
12     crustal, as in coal fly ash, this results in a final particle whose enrichment factor compared to
13     crustal averages depends upon the initial size of the refractory particle-minor for large
       April 1995
6-171
DRAFT-DO NOT QUOTE OR CITE

-------
 1     particles, extreme for ultra-fine particles (Davison et al., 1975).  This result also places the
 2     (potentially) toxic metals on the biologically-accessible surface.
 3          Thus, the presence of metals in a combustion process such as incineration of biological
 4     and chemical wastes or treatment of contaminated soils poses a problem. Raising the
 5     temperature of combustion high enough to completely (> 99.99%) destroy the biological and
 6     chemical species will also enhance the volatilization of metallic components in the feed stock,
 7     requiring more efficient removal methods for ultra-fine and accumulation mode metals.
 8     Figure 6-95 shows the enhanced volatilization of metals as the combustion temperature is
 9     raised from 1000 degrees F  (540 C) to 1800 degrees F (980 C) (Seeker, 1990).
10          The combustion effluent can be partitioned into three components (Seeker, 1990; Barton
11     et al., 1990); emitted (as fly ash), captured (assuming there is an attempt to capture fine
12     particles), and collected in the bottom ash.  Assuming no particle removal equipment is in
13     place on the combustion process, emitted particles will include both the "emitted" component
14     and most of the "captured" component.  In an uncontrolled incineration facility, 96% of
15     mercury, 88% of cadmium,  58% of lead, and 11% of copper might by  emitted into the
16     atmosphere.  If control is attempted, the capture efficiency is only 25%  for mercury, but is
17     better for most other metals, ranging from 86% for cadmium to  91% for copper (Barton
18     et al., 1990). In addition, the chemical state of the metals in the  ultra-fine mode can vary
19     from the more toxic phases (for example, arsenite versus arsenate) as a  function of
20     combustion conditions (Chesworth et al. 1994).  Thus, we must  expect that ultra-fine metallic
21     components will be emitted from high temperature processes  in both toxic and less toxic
22     forms.
23
24     6.9.3   Techniques for  collecting and analyzing ultra-fine metals
25          Relatively little information exists on concentrations of ultra-fine metal particles in
26     ambient air samples away from combustion sources.  There are many reasons.  The ultra-fine
27     mode falls  off rapidly away  from the combustion source,due to the rapid migration of some
28     types of ultra-fine particles into the accumulation mode,  and increased dispersion as one
29     moves away from the source.  Many sources of ultra-fine metals use tall exhaust stacks,
30     which enhances dispersion.  The largest of the ultra-fine particles can overlap the smallest
        April 1995                               6-172      DRAFT-DO NOT QUOTE OR CITE

-------
M _
E
T
A
L 4Q
E
N 30
R
I
C 20
H
M
E 10
N
T
(%)


•
-
-
m
m
" 1000 F
•
-AsSbPbCuCrZnCd

rh-i-.















Cd
••








pbiaooF
Ml






Sb
TCu

Tzn
~i
|Cr
       Figure 6-95. Impact of treatment temperature on the enrichment of metals in the fly
                    ash after the thermal treatment of soils from a superfund site.
 1     particles of the much more abundant accumulation mode, roughly 0.2 to 0.7 m aerodynamic
 2     diameter.  Particles must be size-separated using a device with a sharp cut point, ususally a
 3     multistage  physical impactor, that entails problems in particle collection and analysis.  Since
 4     ultra-fine particles may hard and dry, adhesive coatings are essential in order to avoid
 5     particle bounce in the impactors.  Particle bounce typically translates coarser particles onto
 6     finer stages, contaminating the ultra-fine particles with the enormously more abundant
 7     coarser particles.  Finally, one can collect only a few monolayers of particles (at most) on
 8     the adhesive stages before particle bounce becomes important, assuming the particles
 9     themselves are not "sticky".  A few monolayers of particles of 0.1 m diameter amounts to
10     only about 50  g/cm2 of total deposit.  If one then desires to perform minor or trace
11     elemental analysis of the deposit, one is then faced with analytical requirements that reach
12     picogram (10~12 gm) sensitivities.  This clearly limits analytical options.
13                   For these reasons, much of the data available on the "ultra-fine" mode does
14     not depend on compositional analysis.   Most information on the presence of the ultra-fine
        April 1995
6-173
DRAFT-DO NOT QUOTE OR CITE

-------
  1      mode is derived from particle counting techniques such as the Electrical Mobility Analyzer
  2      (EMA), in situations in which the source is well known (source-enriched).  This was the
  3      method pioneered in the 1972 ACHEX studies of Los Angeles (Whitby, 1978).  Particle
  4      counting devices do not normally result in collection of particles in the ultra-fine mode in a
  5      manner suitable for compositional analysis, although some of the devices ("particle
  6      classifiers") could be modified to provide samples for subsequent compositional analysis, if
  7      desired. The same can be argued for devices such as diffusion batteries, but to date little has
  8      been done along this line in ambient conditions.
  9                   Integrated samples of fine particles can be collected  on substrates  suitable for
10      analysis.  While some optical information is available as one approaches the ultra-fine mode,
11      most optical techniques do not work in the ultra-fine size range, which is well below the
12      wavelength of light.  A Scanning Electron Microscope (SEM) beam can still resolve particles
13      in the ultra-fine mode although some details are lost.  The ultra-fine mode can then be
14      derived by particle counting techniques, either manual or automated,  and metal composition
15      can be found by x-ray analysis of the single particles. The enormous gain in signal to noise
16      ratio by selecting individual particles off sets the loss of x-ray sensitivity (typically part per
17      thousand) caused by use of the electron beams to induce the x-rays.  SEM and electron
18      microprobe analyses rarely  achieve any better than one part per thousand sensitivity, but for
19      single particles, this is  often enough to classify them by  source.  Proton microprobes are,  at
20      present, not quite able to operate in the 0.1 m diameter region, but can perform  Proton
21      Induced X-ray Emission (PIXE)  analysis to one part per million by mass on single particles
22      as small as 0.3  m (Cahill,  1980).
23                   Impactors are designed to separate particles by aerodynamic size in such a way
24      as to allow compositional analysis.   Yet here, too, the ultra-fine mode poses problems.
25      First,  most impactors can not operate effectively in the ultra-fine mode.  The Stokes number
26      for separation of a 0.1  m diameter particle from an air stream requires either extremely  high
27      jet velocities, extremely low pressures in the gas stream, or both.   While such performance
28      can be achieved in a physical  impactor, most impactors used for ambient particle collection
29      in the 1970's and early 1980's did not possess this capability.  For example, the  very popular
30      cyclones and virtual impactors are ineffective below about 0.5 m diameter,  and at ambient
31      pressure, are unlikely to ever  achieve performance in the ultra-fine regime.   The Lundgren-

        April  1995                                6-174      DRAFT-DO  NOT QUOTE OR CITE

-------
  1      type impactors widely used in California studies (Lundgren et al., 1972; Flocchini et al.,
  2      1976; Barone et al.  1978) used 0.5 m as the lowest cut point. Everything smaller was
  3      collected on a filter.  The Battelle-type samplers (Mercer, 1964) favored by other groups
  4      (Van Grieken et al., 1975) used a lowest cut point of 0.25 m diameter.  Thus, while both
  5      these units generated copious information on aerosol composition, they could not  separate
  6      ultra-fine aerosols from accumulation mode aerosols.
  7                   In the mid-1980's four new impactors were developed capable  of resolving
  8      ultra-fine aerosols in ambient conditions; the Low Pressure Impactor, (LPI, Hering et al.,
  9      1978), the Berner Low Pressure Impactor (BLPI, Berner and Lurzer, 1980), the Davis
10      Rotating-drum Unit for Monitoring impactor, (DRUM. Cahill et al., 1985; Raabe et al.,
11      1988), and the Multiple Orifice Uniform Deposit Impactor (MOUDI, Marple et al.  1986).
12      Battelle-type impactors were also modified to add two size cuts below 0.25 m diameter, but
13      unlike the other four units, no certification of performance has been published to date on its
14      performance in the ultra-fine region.  The development of reliable,  clean adhesive coatings
15      such as Apiezon™-L grease was also a major advance in the  field (Wesolowski et al., 1978,
16      Cahill, 1979), allowing separation of abundant soils from ultra-fine  size ranges even in dry,
17      dusty conditions.  For nominally PM-10 soils, for example, a ratio of coarse to ultra-fine
18      soils was measured at  6,600:1 at a temperaturesabove 30 °C  and low relative humidity, RH
19      below 20 %  (Cahill et al.,  1985).  Performances and  specifications  of all these units is
20      included in a recent review paper  (Cahill and Wakabayashi,  1993)
21                   It is important to mention, however, that the motivation for  development of
22      this ultra-fine capability was not for extensive studies of ultra-fine metals,  but rather to get a
23      more complete picture of the accumulation mode behavior of sulfates, nitrates, organics, and
24      other major components of the fine aerosol mix. Thus, compositional analysis was often
25      limited to these species even when suitable samples had been collected.  For example, many
26      LPI samples were collected on stainless steel substrates, ideal for combustion analysis of
27      sulfur, but unsuitable for analysis of transition metals by x-ray techniques.
       April 1995                               6-175      DRAFT-DO NOT QUOTE OR CITE

-------
 1      6.9.4   Observations of ultra-fine metals;  Stack and source-enriched
 2              aerosols
 3                   Observations of ultra-fine metals in source or source-enriched situations
 4      lessens problems with dilution of the sample and identification of the source.  This eases both
 5      particle collection and analysis.  Figure 6-96 shows the results of such a study on a coal fired
 6      power plant (Maenhaut et al.,  1993) using the Berner Low Pressure Impactor (BLPI).  The
 7      extreme volatilization of selenium is clearly seen, which is also confirmed in aircraft
 8      sampling of power plant stacks.  Note, however, that the enrichment factor is rather constant
 9      as a function of particle size for both sulfur and its chemical analog selenium.  Other more
10      refractory elements, on the other hand, are strongly enhanced in the ultra-fine mode as
11      compared to coarser modes.
12                   The BLPI cuts are as follows: Stage number 1-0.011 m diameter, 2-0.021,
13      3-0.032, 4-0.07, 5-0.17, 6-0.30, 7-0.64, 8-1.4, 9-2.6, 10-5.5, 11-10.7 m.  All are for
14      particle density 2.45 g/cm3 and a temperature 120° C, the conditions of stack sampling in the
15      coal fired power plant.  Both these figures were normalized to Earth crustal averages.  Thus,
16      even a two order of magnitude rise in the normalized concentration may not result in a
17      visible "combustion mode"  since the mass of soil falls very rapidly as one moves towards
18      ultra-fine particles. This is exactly what is predicted by the results of Natusch et al. (1974).
19      Thus, source testing confirms combustion theory and  the laboratory studies and predicts
20      emissions of metals into the ultra-fine mode from many types of high temperature
21      combustion sources.
22
23      Observations of ultra-fine metals: Ambient aerosols
24      Direct observations
25                   Because of the difficulties in sampling and analysis, there are relatively few
26      data on concentrations of ultra-fine (Dp 0.1 m diameter) metals in ambient aerosols.  Some
27      quantitative determinations of ambient concentrations  have become available in the past
28      15 years, however, generally as  a result of a number  of short but intensive aerosol studies.
29      Examples include the extensive studies near the Grand Canyon, 1979 (Macias et  al., 1981) to
30      the Mohave Studies near the Grand Canyon NP,  1993, the Southern California Air Quality
31      Study (SCAQS), 1985-1987 (Hering et al., 1990, Cahill et al.,  1990, Cahill et al., 1992);

        April 1995                               6-176      DRAFT-DO NOT QUOTE OR CITE

-------
          lOOOr
           100
        8
        T3
            10
            0.1
S«x0.1
S
Ca
-)
Al
-H
SI
-i
K
                                                     1000T
           "S
                                                   o
                 234587(010
                     Stage number
                         Stage number
       Figure 6-96.  Average normalized concentrations as a function of stage number, for Se,
                    S, Ca, al., Si, K, Mo, W, Ni, and Cr for five BLPI samples from a coal
                    fired power plant.  The smallest size mode is to  the left, Stage number 1,
                    0.011 to Stage number 11, 10.7 m diameter.  Normalization is to average
                    crustal composition.
 1     studies at Shenandoah (1991) and Mt. Rainier (1992) National Parks (Malm et al., 1993,
 2     Malm et al.  1994 b, Cahill and Wakabayashi, 1994), and others. While almost all of these
 3     studies used several different types of impactors with ultra-fine capabilities, relatively  few
 4     were analyzed for trace metal content.
 5          An example of a persistent ambient ultra-fine mode is shown in Figure 6-97 from data
 6     collected at Grand Canyon NP 1984 (Cahill et al., 1987).  The ultra-fine mode behaves
 7     independently from the accumulation mode, in fact often showing a net anti-correlation in
 8     concentrations of sulfur as well as dramatic differences in metals (Table 6-5).  The ultra-fine
 9     mode in Table 6-6 can be attributed  to non-ferrous metal smelting  activities in the region
10     (Eldred et al. 1983, Small et al. 1981), which puts the nearest important sources a hundred
11     miles away from the sampling site.  The completely different behavior of the accumulation
       April 1995
        6-177
DRAFT-DO NOT QUOTE OR CITE

-------
                    400

                    200
                           STAGE 8 (FINE)
                    400-   STAGE?
                   200-
                           STAGE 6
                   400
                   zoo-
                                                  A..
                           STAGE 5
                       .    STAGE 4
                          i IT
                                        10
i  i  rrn r
         20
                      I'M i11  r i  i
                                  00
                                             AUGUST 1984
      Figure 6-97. Fine and ultrafine sulfur at Grand Canyon NP, summer,  1984.  The sulfur
                  peaks on August 15 and August 16 were used for the compositional
                  analysis in Table 3.  Note that the size fractions are inverted, with the
                  finest, Stage 8, at the top,  0.088-0.15 m diameter. The  succeeding stages
                  are at   0.24,   0.34, 0.56, to the coarsest,  Stage 4, 1.15 to 2.4 m. The
                  first three cut points are somewhat uncertain due to altitude and flow rate
                  corrections. Final stage configurations are given in Raabe et al.  (1989),
                  which were used for all later studies using the DRUM.
1     and ultra-fine modes in this arid site also shows that mis-sizing by particle bounce is not

2     significant.

3          Table 6-6 presents a summary of more recent  data for major EPA-regulated metals

4     (lead, nickel) and other metals, at Long Beach, CA, December, 1987 (SCAQS) and at

5     Shenandoah NP, 1991. The elements span the range from refractory metals like nickel and

6     vanadium to metals with low melting temperatures such as zinc and lead.  These data were

7     all taken with the same unit, the Davis Rotating-drum Unit for Monitoring (DRUM) using

8     greased stages and a single orifice  impactor (Cahill et al., 1985). The last two stages were
      April 1995
6-178
    DRAFT-DO NOT QUOTE OR CITE

-------
        TABLE 6-5.  COMPOSITION OF THE AEROSOLS PRESENT AT GRAND CANYON
          NP IN SUMMER 1984, FOR THE TWO SULFATE EPISODES OF AUGUST 15
          (ACCUMULATION MODE, STAGE 6) AND AUGUST 16 (ULTRA-FINE MODE,
                                          STAGE 8)
Elements
Sodium
Silicon and Aluminum
Sulfur
Chlorine
Potassium
Calcium
Titanium
Vanadium
Iron and nickel
Copper
Zinc
Arsenic
Bromine
Lead
Stage 8,
0.088-0. 15m
(ng/m3)
420
8
204
208
59
150
2
2
2
100
931
13
2
63
Stage6,
0. 15-0.60 m
(ng/m3)
10
6
392
5
3
5
4
3
2
1
2
2
2
4
 1     modified form the Gand Canyon configuration as a result of theoretical and laboratory studies
 2     (Raabe et al., 1989), yielding 0.069 to 0.24 m for Stage 8, and 0.24 to 0.34 m diameter for
 3     Stage 7.
 4         The DRUM data were used for several reasons: the DRUM'S slowly rotating greased
 5     stages  have a documented ability to handle large amounts of coarse, dry soils without
 6     contaminating the ultra-fine stages, (Cahill et al. 1985; Cahill and Wakabayashi, 1992), the
 7     elemental data are of unprecedented sensitivity for ambient ultra-fine trace metals (PIXE and
 8     synchrotron-XRF), there  is a consistency of sampler type and protcols at very different
 9     locations, and there is more trace element data from the DRUM than from any other type of
10     unit.  These advantages outweigh its disadvantages; the DRUM does not have the ultra-fine
11     sizing detail of either the LPI or BLPI impactor, or the ability to measure mass, ions and
12     organic matter of the MOUDI or BLPI.

       April 1995                            6-179      DRAFT-DO NOT QUOTE OR CITE

-------
    TABLE 6-6. MEASUREMENTS OF FINE AND ULTRA-FINE METALS
Site Name Particle Ultra-fine
Duration Aerodynamic mode
Frequency Diameters From
Dates To
(Dae, micro- 0.069
meters) 0.24
Long Beach, CA Maximum
6 days Element Values
6 samples/day ng/m3
(11,12/87) Vanadium 6.6
Nickel 3.4
(Mean min. Zinc 51
detectable Selenium MDL
limit - 0.3 ng/m3) Lead 199
Sulfur (est)
Shenandoah NP Maximum
21 days Values
6 samples/day ng/m3
(9/91) Vanadium 1.2
Nickel 1.2
(Mean min. Zinc 3.8
detectable Selenium 2.7
limit -0. 15 ng/m3) Lead 50
Sulfur
(est) - estimated from graphs
Ultra-fine Accumulation
mode Mode
From From From From From
To To To To To
0.069 0.24 0.34 0.56 1.15
0.24 0.34 0.56 1.15 2.5
Mean
Values
ng/m3 ng/m3 ng/m3 ng/m3 ng/ni3
2.5 6.1 10.5 12.2 8.6
1.3 4.4 7.7 4.5 0.5
17.6 46.3 140.4 189.4 39
MDL 0.32 3.00 1.40 0.65
71.4 47.6 59.9 69.9 25.4
200 250 350 500 250
Mean
Values
ng/m3 ng/m3 ng/m3 ng/m3 ng/m3
0.24 0.67 0.52 0.30 0.80
0.58 0.48 0.13 0.03 0.01
1.42 2.16 2.60 1.92 1.66
0.14 0.11 0.52 0.35 0.14
5.38 5.49 3.01 10.87 16.06
334 929 1235 1727 101

April 1995
6-180
DRAFT-DO NOT QUOTE OR CITE

-------
  1          The analyses were done both by PIXE and by synchrotron-XRF (Cahill et al.  1992),
  2     with most of the trace metal data from the latter source. In order to obtain sulfate, multiple
  3     sulfur by 3.0. These average values, however, obscure a great deal of structure as a function
  4     of time.
  5          The variability as a function of size and time is  shown in Figure 6-98 for nickel,
  6     selenium, and lead in Long Beach, CA as part of the SCAQS studies of 1987.  By 1987,
  7     much of the lead was no longer automotive, and there are  significant changes in the ultra-fine
  8     fraction over periods of four to twelve hours.  Note the behavior of ultra-fine metals; almost
  9     total absence for selenium, partial absence for nickel, and constant presence for lead.
 10     Almost all elements at almost every site shows similarly complex behavior.  Thus, the
 11     summary of Table 6-11 can only include the most basic types of information on fine and
 12     ultra-fine metals in the atmosphere.
 13          In addition to the limited US data, comparison data have also become available from
 14     foreign sources  such as from the Kuwaiti oil fires  (1991) and a study in Santiago, Chile,
 15     (1993).  While the former is a unique situation, the Santiago  data are especially useful since
 16     leaded gasoline  is still routinely used in Chile and other countries, generating data impossible
 17     to obtain in the  United States.  Table 6-7 summarizes some of these data for a refractory
 18     element, nickel, and a volatile  metal, lead.. However, the full data set includes 450 samples
 19     of four to six hour duration, each analyzed in five fine size fractions, generally with about
20     20 elements  found in each fraction, or approximately  40,000  individual elemental values.
21           Some general observations can be made  from the data; first, there is an enormous
22     variation in the  concentration of fine and ultra-fine metals,  sometimes spanning 4 or 5 orders
23      of magnitude in a few days.  Such behavior can be modeled by plumes of particles that
24     sweep over the site episodically, as opposed to area or regional sources.  Second, one often
25      finds a mixture  of ultra-fine and accumulation mode behaviors. However,  these may be
26      physically separated in time and size.
27           Lead in the United States follows a variety of very different patterns. In the rural
28      samples, lead tends to be bimodal, with a coarse component above 1.0 m diameter and a
29      very fine and ultra-fine mode below 0.34 m diameter.  This can be modeled by a very fresh
30      ultra-fine mode and a coarser mode associated with resuspended soil.  Urban sites, however,
31      both in the U.S. and in Santiago, show a strong ultra-fine mode and an accumulation mode.

        April 1995                               6-181       DRAFT-DO NOT QUOTE OR  CITE

-------
                            00-
                            200
                            00-
                            20.0
                            0.0-
                            10.0
                          rt
                                        LONG BEACH. CA
                                          SELENIUM
                                        LONG BEACH, CA
                                           LEAD
Figure 6-98.  Fine and ultra-fine metals, nickel, selenium, and lead,  in Long Beach,
             CA, December 10-13, 1987, in four hour increments.  Stage 8 is ultra-76-
             fme, 0.069-0.24 m, then 0.34, 0.56, 1.15, 2.5 m D(ae).
April 1995
6-182
DRAFT-DO NOT QUOTE OR CITE

-------
     TABLE 6-7. MEASUREMENTS OF FINE AND ULTRA-FINE METALS
                            LEAD AND NICKEL
    Site Name          Particle     Ultra-fine.
      Duration          Aerodynamic mode
      Frequency        Diameters
      Date      Element

                      (Dae, micrc
                      meters)

                 Lead  Mean
                       Maximum
                 Nickel  Mean
                       Maximum
                                        Accumulation
                                        mode
Long Beach
6 days
4 samples/day
(11/87)
    Shenandoah NP  Lead  Mean
    21 days             Maximum
    6 samples/day  Nickel  Mean
    (9/91)              Maximum

    Mt. Rainier NP  Lead  Mean
    28 days             Maximum
    6 samples/day  Nickel  Mean
    (7,8/92)             Maximum
From
To
0.069
0.24

71.4
199
1.3
3.4

5.4
50
0.58
1.20

2.3
6
From
To
0.24
0.34

47.6
95
4.4
11.4

5.5
20
0.48
1.60

6.5
15
From
To
0.34
0.56

59.9
129
7.7
15.0

3.0
16
0.13
0.80

2.0
21
From
To
0.56
1.15

69.9
164
4.5
13.4

10.9
70
0.03
1.00

3.4
14
From
To
1.15
2.5

25.4
58
0.5
3.7

16.1
130
0.01
0.14

6.7
29




MDL
0.45

0.22

MDL
0.2

0.09

MDL
0.5

                             Always less than MDL
                             MDL  0.4   0.8   0.4
                 0.7
    Santiago, Chile  Lead  Mean    101   53   38    108   41
    14 days
    6 samples/day
    (9/93)
    Kuwait       Lead
    14 days
    4 samples/day Nickel
    (6/91)
                   Maximum  920   340  320   640   270
                   Maximum  5
18    11    8
0.07

MDL
8

Mean
Maximum
Mean

429.9
2580
1.5

154.2
580
2.5

84.7
128
4.3

44.7
86
3.7

38.1
70
6.0
MDL
0.35

0.22
          MDL = minimum dectable limit at 95% confidence level, in ng/m^
April 1995
                               6-183
      DRAFT-DO NOT QUOTE OR CITE

-------
 1     Any resuspended soil mode is hidden under the extension of the accumulation mode above
 2     1.0 m diameter.
 3                    Other metals at Long Beach,  however, lack a distinct ultra-fine mode all the
 4     time (selenium) or part of the time (nickel),  merely possessing an accumulation mode that
 5     closely mimics sulfates and other secondary  species (Cahill et al., 1990).  It is well known
 6     that nickel and vanadium were derived from high temperature combustion sources, and since
 7     each is highly refractory, they will occur primarily in an ultra-fine mode at the source.
 8     Thus,  the similarity between the distributions of these elements and less refractory elements
 9     such as zinc can be understood through a rapid condensation and coagulation of the abundant
10     secondary species around these metals, leading to an  accumulation mode distribution as the
11     secondary acidic  species hydrate. Clearly, such processes are weaker  at dry sites such  as the
12     arid west in summer (Table 6-6). On the other hand, Shenandoah NP has a mixture of urban
13     and rural behavior, with occasional sharp peaks of ultra-fine metals (nickel) superimposed on
14     an accumulation mode behavior  (sulfur, selenium) but some coarse contribution (lead,
15     vanadium).  Only with a detailed study of meteorology and knowledge of emission sources
16     can such ambient behavior be understood.
17
18     Indirect methods
19                    Lacking a large body of direct data on ultra-fine metallic aerosols, there are
20     indirect ways to increase our knowledge of such aerosols;
21          1.  Combustion studies have established the modes of formation of ultra-fine metallic
22              aerosols, and,
23
24          2.  Considerable ambient data exist that, when combined with known combustion
25              processes,  yield estimates for the concentration of ultra-fine metallic aerosols by
26              time and locations.
27
28          3.  In conditions of low ambient concentrations  of particles and low humidity and , the
29              ultra-fine mode has been shown to persist for many hours. (Cahill et al., 1985).
30
31          Thus, the numerous observations of fine (Dp <  2.5 m) metallic aerosols  in low
32     humidity conditions can yield estimates of the presence of such metals in the ultra-fine mode
33     and set upper limits on their concentrations.   The relatively small number of actual
34     measurements can then serve as  tests or as confirmation of our level of understanding of

       April  1995                               6-184      DRAFT-DO NOT QUOTE OR CITE

-------
  1     these biologically important aerosols. As an example, Figure 6-101 shows concentration
  2     profiles of sulfur, selenium, zinc, and arsenic, all of which can have ultra-fine modes in the
  3     western United States.  Arsenic and zinc are annual averages, March,  1993 to February,
  4     1994, while the sulfur (for sulfate, times 3.0) and selenium are  for summer,  1993.  This was
  5     done to exhibit the correlation of these elements, which are chemically akin, during the
  6     eastern U.S. sulfate maximum each summer. The regional nature of the elements is very
  7     evident, as are certain strong sub-regional sources such as the copper smelter region of
  8     Arizona and New Mexico (arsenic).
  9          The non-urban values shown in Figure 6-99, which are derived from the cleanest areas
 10     of the United States, are surprisingly relevant to urban areas in  the same region for some of
 11     the species.  Table 6-8 compares major and minor fine elements at Shenandoah NP,  where
 12     there are detailed measurements of particle size, and Washington, DC, where such size
 13     information is lacking.  Summer 1993 is the comparison period. Finally, two western sites
 14     are compared,  both downwind of Los Angeles;  San Gorgonio Wilderness, and Grand Canyon
 15     NP.
 16
 17     Inhalation of ultra-fine metals
 18          An extensive literature exists on the deposition of fine metals in the human lung, much
 19     of which was derived from laboratory studies, some using radioactive tracer isotopes.  But an
 20     example of one of the  few direct measurement of lung capture by ambient ultra-fine metals is
 21      found in Dasaedeleer et al.,  1977 and shown in Figure 6-100.  The lower cut point is only
 22     0.25 m but even  so, the increased capture efficiency of the lung for very fine and ultra-fine
 23      particles is clearly shown.
 24
 25
 26      6.10  SUMMARY
27          There are few data on ambient  concentrations of ultra-fine metals. The few direct
28      measurements can be extended with some confidence using indirect methods;  i.e., particle
29      counting techniques that have size information but no chemical information, or filter
30      collection methods that have limited size information but detailed compositional information.
       April 1995                               6_!85      DRAFT-DO NOT QUOTE OR CITE

-------
                                                                I
                                                                S
                                                                I
                                                                .5
                                                                O 3
                                                                U <
                                       cw
                                       •o
                                       
-------
  TABLE 6-8. COMPARISON OF SELECTED SPECIES, SHENANDOAH NP, AND
 WASHINGTON, DC, SAN GORGONIO WILDERNESS, CA, AND GRAND CANYON
                        NP, SUMMER, 1993
Shenandoah
Concentration (ng/m^)
Mass - PM-10
Mass - PM-2.5
Composition PM-2.5 Mass
Ammonium sulfate
Ammonium nitrate
Organic matter
Soil
Trace composition (ng/m^)
Nickel
Copper
Zinc
Arsenic
Selenium
Bromine
Lead
Bio-smoke tracer
NP
31.00
22.50

11.80
0.40
2.84
1.41

0.24
1.06
7.93
0.22
1.58
2.14
2.17
8.33
Washington
DC
34.90
26.50

14.60
1.47
5.42
1.55

0.97
3.37
13.90
0.56
2.48
4.18
4.48
<2.00
San Gorgonio
Wilderness
21.70
10.30

2.55
4.44
3.88
0.86

0.18
0.76
3.72
0.16
0.44
3.67
1.36
10.00
Grand Canyon
NP
9.37
4.50

1.09
0.25
1.22
0.63

0.09
0.30
0.63
0.18
0.18
2.11
0.51
32.30
(non-soil fine potassium)
Optical Absorption
(b(abs), 10-6 m-l)
19.60

41.90

13.90

5.40

April 1995                      6_187    DRAFT-DO NOT QUOTE OR CITE

-------
    c
    0
   TJ
    QJ
    O
   .C
    x
   UJ
    O
   rr
c
g
"o
c
u
c
o
O
      05--
       02--
       01--
      005--
      0.02
                                                                      0

                                                                      •20

                                                                    -•40


                                                                    -•60
                                            i
^ Pb
D Br
  C!
                                                           c
                                                           o
                                                        80^
                                                           to
                                                           O
                                                           CL
                                                           O)
                                                           O
                                                     --90
                                                                         92  a;
                                                                            O
                                                                       -•94
4, 4-2, 2-1, 1-0.5, 0.5-
             0.25, and <0.25 /on particles of size classes 1, 2, 3, 4, 5, and 6,
             respectively. Extension of the curve to particles of diameter > 2 pm
             (classes 2 and 1) is supported by separate experiments using chalk dust
             aerosol.
Source:  Dasaedeleer et al., 1977.
April 1995
                                   6-188
                               DRAFT-DO NOT QUOTE OR CITE

-------
1      Nevertheless, it is clear that more data on ultra-fine metals is urgently needed to gain
2      confidence in the spatial and temporal concentration profiles of this key  atmospheric
3      component.
4           Ultra-fine metals are produced by a wide variety of anthropogenic activities and emitted
5      into the ambient air. Ambient concentrations of such metals have been seen not only in urban
6      settings but also at the cleanest sites in the United States. Concentrations are highly variable
7      as a function of site and time. While ultra-fine metals have been seen to persist for many
8      hours, or more, in the clean, dry environment of the arid west, they  appear to be removed
9      and/or transformed into the accumulation mode in polluted urban or humid rural sites.
      April 1995                               6489      DRAFT-DO NOT QUOTE OR CITE

-------
 1      REFERENCES
 2
 3     Adams, K. M.; Davis, L. I., Jr.; Japar, S. M.; Finley, D. R.; Cary, R. A. (1990) Measurement of atmospheric
 4            elemental carbon: real-time data for Los Angeles during  summer 1987. Atmos. Environ. Part A
 5            24: 597-604.
 6
 7     Altshuller, A. P. (1980)  Seasonal and episodic trends in sulfate concentrations (1963-1978) in the eastern United
 8            States. Environ.  Sci.  Technol. 14:  1337-1349.
 9
10     Appel, B. R.; Colodny, P.; Wesolowski, J. J. (1976) Analysis of carbonaceous materials in southern California
11            atmospheric aerosols. Environ. Sci. Technol. 10: 359-363.
12
13     Appel, B. R.; Kothny, E. L.; Hoffer, E. M.; Hidy, G. M.; Wesolowski, J. J. (1978) Sulfate and nitrate data
14            from the California Aerosol Characterization Experiment (ACHEX). Environ. Sci.  Technol. 12: 418-425.
15
16     Appel, B. R.; Hoffer, E. M.; Kothny, E. L.; Wall, S.  M.; Haik, M.; Knights, R. L. (1979) Analysis of
17            carbonaceous material in southern California atmospheric aerosols. 2. Environ. Sci. Technol. 13: 98-104.
18
19     Appel, B. R.; Tokiwa, Y.; Haik, M. (1981) Sampling of nitrates in ambient air. Atmos. Environ. 15: 283-289.
20
21     Appel, B. R.; Tokiwa, Y.; Hsu, J.; Kothny, E. L.; Hahn, E. (1985) Visibility as related to atmospheric aerosol
22            constituents. Atmos. Environ. 19: 1525-1534.
23
24     Ashbaugh et al. (1989) Estimating fluxes from California's dry deposition monitoring data. Presented at: 82nd
25            annual meeting of the Air & Waste Management Association; June; Anaheim, CA. Pittsburgh, PA: Air &
26            Waste Management Association; paper no. 89-65.3.
27
28     Barone, J. B.; Cahill, T. A.; Eldred, R. A.; Flocchini, R. G.; Shadoan, D. J.; Dietz, T. M. (1978) A
29            multivariate statistical analysis of visibility degradation at four California cities.  Atmos. Environ.
30             12: 2213-2221.
31
32     Barton, R. G.; Clark, W. D.; Seeker, W. R. (1990) In: Proceedings of the 1st international conference on toxic
33            combustion byproducts.
34
35     Bennett,  R. L.; Stockburger, L.; Barnes, H. M. (1994) Comparison of sulfur measurements from a regional fine
36            particle network with concurrent acid modes network results. Atmos. Environ. 28:  409-419.
37
38     Berner, A.; Lurzer, C.  (1980) J. Phys. Chem. 84: 2079-2083.
39
40     Bowman, H. R.; Conway, J. G.; Asaro, F. (1972) Atmospheric lead and bromine concentration in Berkeley,
41             Calif. (1963-70). Environ.  Sci. Technol. 6: 558-560.
42
43     Brauer, M.; Koutrakis, P.; Spengler, J. D.  (1989) Personal exposures to acidic aerosols and gases. Environ. Sci.
44             Technol. 23: 1408-1412.
45
46     Brauer, M.; Koutrakis, P.; Keeler, G. J.; Spengler, J.  D.  (1991) Indoor and outdoor concentrations of inorganic
47             acidic aerosols and gases. J.  Air Waste Manage. Assoc. 41: 171-181.
48
49     Burton, R. M.; Liu, L. J. S.; Wilson, W.  E.; Koutrakis, P.  (1995) Comparison of aerosol acidity in urban and
50             semi-rural  environments. Atmos. Environ.: submitted.
51
52     Cadle, S. H. (1985) Seasonal variations in nitric acid, nitrate, strong aerosol acidity, and ammonia in an urban
53             area. Atmos. Environ. 19: 181-188.
54


         April  1995                                    6-190       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Cadle, S. H.; Dasch, J. M. (1988) Wintertime concentrations and sinks of atmospheric paniculate carbon at a
  2            rural location in northern Michigan. Atmos. Environ. 22: 1373-1381.
  3
  4     Cadle, S. H.; Countess, R. J.; Kelly, N. A. (1982) Nitric acid and ammonia in urban and rural locations.
  5            Atmos. Environ. 16: 2501-2506.
  6
  7     Cahill, T. A.  (1979) Comments on surface coating for Lundgren-type impactors. In: Lundgren, D. A.;
  8            Lippmann, M.; Harris, F. S., Jr.; Clark, W. E.; Marlow, W.  H.; Durham, M. D., eds.  Aerosol
  9            measurement: [papers from a workshop]; March  1976; Gainesville, FL. Gainesville, FL:  University
 10            Presses of Florida; pp. 131-134.
 11
 12     Cahill, T. A.  (1980) Proton microbes and particle-induced X-ray analytical systems. Ann. Rev. Nucl. Particle
 13            Sci. 30: 211-252.
 14
 15     Cahill, T. A.; Wakabayashi, P. (1993) Compositional analysis of size-segregated aerosol samples. In: Newman,
 16            L.,  ed. Measurement challenges in atmospheric chemistry. Washington, DC: American Chemical Society;
 17            pp.  211-228. (Advances in chemistry series no. 232).
 18
 19     Cahill, T. A.; Wakabayashi, P. H.; James, T. A. (n.d.) Chemical states of sulfate at Shenandoah National Park
 20            during summer, 1991.
 21
 22     Cahill, T. A.; Goodart, C.; Nelson, J.  W.; Eldred, R. A.; Nasstrom, J. S.; Feeney, P. J.  (1985) In: Ariman,
 23            T.;  Veziroglu, T.  N., eds. Proceedings of the international symposium on particulate and multi-phase
 24            processes. Washington, DC: Hemisphere Publishing  Corporation;  pp. 319-325.
 25
 26     Cahill, T. A.; Eldred, R.  A.; Feeney, P. A. (1986) Particulate monitoring and data analysis for the National
 27            Park Service 1982-1985. Davis, CA: University of California.
 28
 29     Cahill, T. A.; Feeney, P.  J.; Eldred, R. A. (1987) In:  Proceedings of the 4th international PIXE  conference;
 30            June 1986;  Tallahassee, FL. Nucl. Instrum. Methods Phys. Res. B22: 344-348.
 31
 32     Cahill, T. A.; Eldred, R.  A.; Motallebi, N.; Malm, W. C. (1989) Indirect measurement of hydrocarbon aerosols
 33            across the United States by nonsulfate  hydrogen-remaining gravimetric mass correlations.  Aerosol Sci.
 34            Technol.  10: 421-429.
 35
 36     Cahill, T. A.; Surovik,  M.; Wittmeyer, I. (1990) Visibility and aerosols during the 1986 Carbonaceous Species
 37            Methods Comparison Study. Aerosol Sci. Technol. 12: 149-160.
 38
 39     Cahill, T. A.; Matsumura, R. T.; Wakabayashi, P. (1992) Presented at: 85th annual meeting and exhibition of
 40            the Air & Waste Management Association; June;  Kansas City, MO. Pittsburgh, PA: Air & Waste
 41            Management Association.
 42
 43     Cahill, T. A.;  Wilkinson,  K.; Schnell, R. (1992) J. Geophys. Res. [Atmos.] 97: 14513-14520.
 44
 45     Cass, G. R. (1979) On the relationship between sulfate air quality and visibility with examples in  Los Angeles.
46            Atmos. Environ. 13: 1069-1084.
47
48     Cass, G. R.; McRae, G. J. (1983) Source-receptor reconciliation  of routine air monitoring data for trace metals:
49            an emission inventory assisted approach. Environ. Sci. Technol. 17: 129-139
50
51      Cass, G. R.; Conklin, M.  H.; Shah,  J. J.; Huntzicker,  J.  J.;  Macias, E. S. (1984) Elemental carbon
52            concentrations: estimation of an historical data base. Atmos. Environ.  18:  153-162.
53
         April 1995                                    6_191       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Chesworth, S.; Yang, G.; Chang, D.; Jones, A. D.; Kelley, P. B.; Kennedy, I. M. (1994) Combust. Flame
  2             98: 259-266.
  3
  4      Chow, J. C. (1985) Development of a composite modeling approach to assess air pollution source/receptor
  5             relationship [dissertation]. Cambridge, MA: Harvard University.
  6
  7      Chow, J. C.; Watson, J. G. (1988) Summary of paniculate databases for receptor modeling in the United States.
  8             In: Watson, J. G., ed. Receptor models in air resources management. Pittsburgh, PA: Air and Waste
  9             Management Association.
10
11      Chow, J. C.; Watson, J. G.; Frazier, C. A.; Egami, R. T.; Goodrich, A.; Ralph, C. (1988) Spatial and
12             temporal source contributions to PM10 and PM2 5 in Reno, NV. In: Mathai, C. V.; Stonefield, D. F.,
13             eds.  PM10: implementation of standards. Pittsburgh, PA: Air Pollution Control Association; pp. 438-457.
14
15      Chow, J. C.; Watson, J. G.; Pritchett, L.; Lowenthal, D.; Frazier, C.; Neuroth, G.;  Evans, K.; Moon, D.
16             (1990) Wintertime visibility in Phoenix, Arizona. Presented at: the 83rd annual meeting & exhibition;
17             June; Pittsburgh, PA. Pittsburgh, PA: Air & Waste Management Association; paper no. 90-66.6.
18
19      Chow, J. C.; Liu, C. S.; Cassmassi, J.; Watson, J. G.; Lu, Z.; Pritchett, L. C. (1992) A neighborhood-scale
20             study of PM10 source contributions in Rubidoux, California. Atmos. Environ. Part A 26: 693-706.
21
22      Chow, J. C.; Watson, J. G.; Lowenthal, D.  H.; Frazier,  C.; Hinsvark, B.; Prichett, L.; Neuroth, G. (1992)
23             Wintertime PM10 and PM2 5 mass and chemical compositions in Tucson, Arizona.  In: Chow, J. C.; Ono,
24             D. M., eds. Transactions: PM10 standards and nontraditional paniculate source controls. Pittsburgh,  PA:
25             Air and Waste Management; pp.  231-243.
26
27      Chow, J. C.; Watson, J. G.; Lowenthal, D.  H.; Solomon, P.  A.; Magliano, K. L.; Ziman, S. D.; Richards, L.
28             W. (1992) PM10 source apportionment in California's San Joaquin Valley. Atmos.  Environ. Part A
29             26: 3335-3354.
30
31      Chow, J. C.; Watson, J. G.; Lowenthal, D.  H.; Solomon, P.  A.; Magliano, K. L.; Ziman, S. D.; Richards, L.
32             W. (1993) PM10 and PM2 5 compositions in California's San Joaquin Valley. Aerosol Sci.  Technol.
33             18: 105-128.
34
35      Chow, J. C.; Watson, J. G.; Ono, D. M.; Mathai, C.  V. (1993) PM10 standards and nontraditional paniculate
36             source controls: a summary of the A&WMA/EPA international specialty conference. Air Waste
37             43: 74-84.
38
39      Chow et al. (1994) Planning for SJVAQS/AUSPEX paniculate matter and visibility sampling and analysis.
40             In: Solomon, P., ed. Planning and managing regional air quality.  Boca Raton, FL: CRC Press, Inc.
41
42      Chow, J. C.; Watson, J. G.; Fujita, E. M.;  Lu, Z.; Lawson,  D. R.; Ashbaugh, L. L. (1994) Temporal and
43             spatial variations of PM2 5 and PM10 aerosol in the Southern California Air Quality Study. Atmos.
44             Environ.  28: 2061-2080.
45
46      Chow, J. C.; Fairley, D.; Watson, J. G.; DeMandel, R.; Fujita, E. M.; Lowenthal, D. H.; Lu, Z.;  Frazier, C.
47             A.; Long, G.; Cordova, J. (1995) Source apportionment of wintertime PM10 at San Jose, CA. J.  Am.
48             Soc.  Civ. Eng.: in print.
49
50      Chow, J. C.; Watson, J. G.; Lu, Z.; Lowenthal, D. H.; Frazier, C. A.;  Solomon, P. A.;  Thuiller, R. H.;
51             Magliano, K. (1995) Descriptive  analysis of PM2 5 and PM10 regionally representative locations during
52             SJVAQS/AUSPEX. Atmos. Environ.: in press.
53
         April 1995                                   6-192       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Chu, L.-C.; Macias, E. S. (1981) Carbonaceous urban aerosol—primary or secondary? In: Macias, E. S.;
  2            Hopke, P. K., eds. Atmospheric aerosol: source/air quality relationships; based on a symposium jointly
  3            sponsored by the Divisions of Nuclear Chemistry and Technology and Environmental Chemistry at the
  4            180th national meeting of the American Chemical Society; August  1980; Las Vegas, NV. Washington,
  5            DC: American Chemical Society; pp. 251-268. (Comstock, M. J.,  ed. ACS symposium series: 167).
  6
  7     Coboura, W.  G.; Husar, R.  B.  (1982) Diurnal and seasonal patterns of paniculate sulfur and sulfuric acid in
  8            St. Louis, July 1977-June 1978. Atmos. Environ. 16: 1441-1450.
  9
 10     Conklin, M. H.; Cass, G. R.; Chu, L.-C.; Macias, E. S. (1981) Wintertime carbonaceous aerosols in Los
 11            Angeles: an exploration of the role of elemental carbon. In: Macias, E. S.; Hopke, P. K., eds.
 12            Atmospheric aerosols: source/air quality relationships; based on a symposium jointly sponsored by the
 13            Divisions of Nuclear Chemistry and Technology  and Environmental Chemistry at the 180th national
 14            meeting of the American Chemical Society; August 1980; Las Vegas, NV. Washington, DC:  American
 15            Chemical Society; pp. 235-250. (Comstock, M. J., ed. ACS symposium series: 167).
 16
 17     Conner, T. L.; Miller, J. L.; Willis, R. D.; Kellogg, R. D.; Dann, T. F. (1993) Source apportionment of fine
 18            and coarse particles in southern Ontario, Canada.  Presented at: 86th annual meeting and exhibition of the
 19            Air & Waste Management Association; June;  Denver, CO. Pittsburgh, PA: Air & Waste Management
 20            Association;  paper no. 93-TP-58.05.
 21
 22     Countess, R. J.; Wolff, G. T.; Cadle, S. H. (1980) The  Denver winter aerosol: a comprehensive chemical
 23            characterization. J. Air Pollut. Control Assoc. 30:  1194-1200.
 24
 25     Countess, R. J.; Cadle, S. H.; Groblicki, P. J.; Wolff, G. T. (1981)  Chemical analysis of size-segregated
 26            samples of Denver's ambient particulate. J. Air Pollut. Control Assoc. 31: 247-252.
 27
 28     Cronn, D. R.; Charlson, R. J.;  Knights, R. L.; Crittenden, A. L.; Appel,  B. R. (1977) A survey of the
 29            molecular nature of primary and secondary components of particles in urban air by high-resolution mass
 30            spectrometry. Atmos. Environ.  11: 929-937.
 31
 32     Davidson, C.  L; Osborn, J. F. (1984) The  sizes of airborne trace metal-containing particles.  In: Nriagu, J.  O.;
 33            Davidson, C. I., eds. Toxic metals in the atmosphere. New York, NY: John Wiley & Sons. (Advances in
 34            environmental science and technology: v. 17).
 35
 36     Davis, B. L.;  Johnson, L. R.; Stevens, R. K.; Courtney, W. J.; Safriet, D. W. (1984) The quartz content and
 37            elemental composition of aerosols from selected sites of the EPA inhalable particulate network. Atmos.
 38            Environ. 18: 771-782.
 39
 40     Davison, R. L.; Natusch, D. F. S.; Wallace, J.  R.; Evans, C. A., Jr.  (1975) Environ. Sci. Technol. 9: 862.
 41
42     Desaedeleer, G. G.;  Winchester, J. W.; Akselsson, K. R. (1977) Nucl. Instrum. Methods 142: 97-99.
43
44     Desert Research Institute, (n.d.) [Unpublished data].
45
46     Dockery, D. W.;  Schwartz, J.; Spengler, J. D. (1992) Air pollution and daily mortality: associations with
47            particulates and acid aerosols. Environ. Res. 59: 362-373.
48
49     Dockery, D. W.;  Pope, C. A., Ill; Xu,  X.; Spengler, J.  D.; Ware, J. H.; Fay, M. E.; Ferris, B. G., Jr.;
50            Speizer, F. E. (1993) An association between  air pollution and mortality in six U.S. cities. N. Engl. J.
51            Med. 329: 1753-1759.
52
53     Dresser, A. L. (1988) A dispersion and  receptor model analysis of western community's PM-10 problem.
54            JAPCA38: 1419-1421.


         April 1995                                    6-193      DRAFT-DO NOT QUOTE OR CITE

-------
  1     Dronamraju, M.; Peters, L. K.; Carmichael, G. R.; Kasibhatla, P.; Cho, S.-Y. (1988) An Eulerian
  2            transport/transformation/removal model for SO2 and sulfate—III. comparison with the July 1974 SURE
  3            database. Atmos. Environ. 22: 2003-2011.
  4
  5     Dzubay, T. G. (1980) Chemical element balance method applied to dichotomous sampler data. Ann.  N. Y. Acad.
  6            Sci. 338: 126-144.
  7
  8     Dzubay, T. G.; Stevens, R. K.; Lewis, C. W.; Hern, D. H.; Courtney, W. J.; Tesch, J. W.; Mason, M. A.
  9            (1982) Visibility and aerosol composition in Houston, Texas. Environ. Sci. Technol. 16: 514-525
 10
 11     Dzubay, T. G.; Stevens, R. K.; Haagenson, P. L. (1984) Composition and origins of aerosol at a forested
 12            mountain in Soviet Georgia.  Environ. Sci. Technol.  18: 873-883.
 13
 14     Dzubay, T. G.; Stevens, R. K.; Gordon, G. E.; Olmez, I.; Sheffield,  A. E.; Courtney, W. J. (1988) A
 15            composite receptor method applied to Philadelphia aerosol. Environ. Sci. Technol. 22: 46-52
 16
 17     Einfeld, W.; Dattner, S. (1988) The Dallas winter visibility study. Austin, TX: Texas Air Control Board.
 18
 19     Eldering,  A.;  Cass, G.  R. (1994) A  source-oriented model for air pollutant effects on visibility. J. Geophys. Res.
20            [Atmos.]: submitted.
21
22     Eldred, R. A. (1994) Comparison of selenium  and sulfur in the IMPROVE network.  J. Air Waste Manage.
23            Assoc.: submitted.
24
25     Eldred, R. A.; Cahill, T. A. (1994)  Trends in  elemental concentrations of fine particles at remote sites in the
26            United States of America. Atmos. Environ. 28: 1009-1019.
27
28     Eldred, R. A.; Cahill, T. A. (1994)  J. Air Waste Manage. Assoc.: submitted.
29
30     Eldred, R. A.; Ashbaugh, L. L.; Cahill, T. A.; Flocchini, R. G.; Pitchford, M. L. (1983) Sulfate levels in the
31            southwest during the 1980 copper smelter strike. J. Air Pollut. Control Assoc. 33: 110-113.
32
33     Eldred, R. A.; Cahill, T. A.; Feeney, P. J.; Malm, W. C. (1987) Regional patterns in paniculate matter from
34            the National Park Service Network, June  1982 to May 1986. In: Bhardwaja, P. S., ed. Visibility
35            protection research and policy aspects,  pp. 386-396.
36
37     Eldred, R. A.; Cahill, T. A.; Pitchford, M.; Malm, W. C. (1988) IMPROVE—a new remote area paniculate
38            monitoring system for visibility studies. Presented at: the 81st annual meeting of the Air Pollution
39            Control Association; June; Dallas, TX. Pittsburgh, PA: Air Pollution Control Association; paper no. 88-
40            54.3.
41
42     Eldred, R. A.; Cahill, T. A.; Wilkinson, L. K.; Feeney, P. J.; Chow, J. C.; Malm,  W. C. (1989) Measurement
43            of fine particles and their chemical components in the IMPROVE/NPS networks. In:  Mathai, C. V., ed.
44            Visibility and fine particles; pp. 187-196.
45
46     Eldred, R. A.; Cahill, T. A.; Feeney, P. J. (1994) Regional patterns of selenium and other trace elements in the
47            IMPROVE network. In: Proceedings of the international specialty conference  on aerosol  and  atmospheric
48            optics: radiative balance and  visual air  quality: v. A. Pittsburgh, PA: Air and Waste Management
49            Association.
50
51     Eldred, R. A.; Cahill, T. A.; Flocchini, R.  G.  (1994) Comparison of PM10 and PM2 5 aerosols in the IMPROVE
52            network. Proceedings of the international specialty conference on aerosol and atmospheric optics:
53            radiative balance and visual air quality, volume A. Air Waste:  submitted.
54


         April 1995                                   6-194       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Ellestad, T. G.; Knapp, K. T. (1988) The sampling of reactive atmospheric species by transition-flow reactor:
  2            application to nitrogen species. Atmos. Environ. 22: 1595-1600.
  3
  4     Fairley, D.  (1990) The relationship of daily mortality to suspended particulates in Santa Clara county, 1980-86.
  5            Environ. Health Perspect. 89: 159-168.
  6
  7     Federal Register. (1986) Hazardous waste management system; land disposal restrictions. F. R. (November 7)
  8            51:  40572-40654.
  9
 10     Federal Register. (1989) F. R. (February 6) 54: 5746.
 11
 12     Fekete, K. E.; Gyenes, L. (1993) Regional scale transport model for ammonia and ammonium. Atmos. Environ.
 13            Part A 27: 1099-1104.
 14
 15     Ferman, M. A.; Wolff, G. T.; Kelly, N. A. (1981) The nature and sources of haze in the Shenandoah
 16            Valley/Blue Ridge Mountains area. J. Air Pollut. Control Assoc.  31: 1074-1082.
 17
 18     Flocchini, R. G.; Cahill, T. A.; Shadoan, D. J.; Lange, S. J.; Eldred, R. A.;  Feeney, P. J.; Wolfe, G. W.;
 19            Simmeroth, D. C.; Suder, J. K. (1976) Monitoring California's aerosols by size and elemental
 20            composition. Environ. Sci. Technol. 10: 76-82.
 21
 22     Flocchini, R. G.; Cahill, T. A.; Eldred, R. A.; Feeney, P. J. (1989) Paniculate sampling in the northeast, a
 23            description of the Northeast States for Coordinated Air Use Management  (NESCAUM) network. In:
 24            Mathai, C. V., ed. Visibility and fine particles, pp. 197-206.
 25
 26     Forrest, J.;  Tanner, R. L.; Spandau, D.; D'Ottavio, T.; Newman, L. (1980) Determination of total inorganic
 27            nitrate utilizing collection of nitric acid on NaCl-impregnated filters. Atmos. Environ. 14: 137-144.
 28
 29     Frazier, C.  A.  (1989) Letter report on mass and chemical concentrations  and babs measurements for Phoenix
 30            Pilot Study samples. Phoenix, AZ:  Arizona  Department of Environmental Quality.
 31
 32     Friedlander, S. K. (1977) Smoke, dust and haze: fundamentals of aerosol behavior. New York, NY: John Wiley
 33            & Sons, Inc.
 34
 35     Gebhart, K. A.; Malm, W. C. (1987) Source apportionment of paniculate sulfate concentrations at three National
 36            Parks in the Eastern United States.  In: Bhardwaja, P. S., ed. Visibility protection research and policy.
 37            pp.  898-913.
 38
 39     Gebhart, K. A.; Malm, W. C.; Day, D. (1994) Examination of the effects of sulfate acidity and relative humidity
40            on light scattering  at Shenandoah National Park. Atmos.  Environ. 28:  841-849.
41
42     Gillani, N. V.; Kohli, S.; Wilson, W. E. (1981) Gas-to-particle conversion of sulfur in power plant plumes—I.
43            parametrization of the conversion rate for dry, moderately  polluted ambient conditions. Atmos. Environ.
44            15:  2293-2313.
45
46     Gillette, D.  A.;  Sinclair, P. C. (1990) Estimation of suspension of alkaline material by dust devils in the United
47            States. Atmos. Environ. Part A 24: 1135-1142.
48
49     Gregory, J.  M.; Peterson,  R. E.; Lee, J. A.; Wilson, G. R. (1994) Modeling wind and relative humidity effects
50            on air quality. In: Proceedings of the international specialty conference on aerosol and atmospheric
51            optics: radiative balance and  visual air quality: v. A. Pittsburgh, PA: Air and Waste Management
52            Association.
53
         April 1995                                    6-195       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Groblicki, P. J.; Wolff, G. T.; Countess, R. J. (1981) Visibility-reducing species in the Denver "brown
  2            cloud"—I. relationships between extinction and chemical composition. In: White, W. H.; Moore, D. J.;
  3            Lodge, J. P., Jr., eds. Plumes and visibility: measurements and model components, proceedings of the
  4            symposium; November 1980; Grand Canyon National Park, AZ. Atmos. Environ.  15: 2473-2484.
  5
  6      Grosjean, D. (1983) Polycyclic aromatic hydrocarbons in Los Angeles air from samples collected on Teflon,
  7            glass and quartz filters. Atmos. Environ. 17: 2565-2573.
  8
  9      Grosjean, D.; Bytnerowicz, A. (1993) Nitrogeneous air pollutants at a southern California mountain forest smog
10            receptor site. Atmos.  Environ. Part A 27: 483-492.
11
12      Gundel, L. (1979) Discussion. In: Novakov, T., ed. Proceedings: carbonaceous particles in the atmosphere;
13            March 1978; Berkeley, CA. Berkeley, CA: University of California, Lawrence Berkeley Laboratory; p.
14            91; report no. LBL-9037.
15
16      Harrison, P. R.; Draftz,  R. G.; Murphy, W. H. (1976) Identification and impact of Chicago's ambient suspended
17            dust. In: Atmosphere-surface exchange of particulate and gaseous pollutants (1974): proceedings of a
18            symposium; September 1974; Richland, WA. Oak Ridge, TN: Energy Research and Development
19            Administration; pp. 540-556. (ERDA symposium series 38). Available from:  NTIS, Springfield, VA;
20            CONF 740921.
21
22      Hering, S. V.; Friedlander, S. K. (1982) Origins of aerosol sulfur size distributions in the Los Angeles basin.
23            Atmos. Environ. 16: 2647-2656.
24
25      Hering, S. V.; McMurry, P.  H.  (1991) Optical counter response to monodisperse atmospheric aerosols. Atmos.
26            Environ. Part A 25: 463-468.
27
28      Hering, S. V.; Flagan, R. C.; Friedlander, S.  K. (1978) Design and evaluation of new low-pressure impactor. I.
29            Environ. Sci. Technol. 12: 667-673.
30
31      Hering, S. V.; Appel, B. R.; Cheng, W.; Salaymeh, F.; Cadle, S. H.; Mulawa, P. A.; Cahill, T.  A.; Eldred,
32            R. A.; Surovik, M.; Fitz, D.; Howes,  J. E.; Knapp, K. T.; Stockburger, L.; Turpin, B. J.; Huntzicker,
33            J. J.; Zhang, X.-Q.; McMurry, P. H.  (1990) Comparison of sampling methods for carbonaceous aerosols
34            in ambient air. Aerosol Sci. Technol. 12: 200-213.
35
36      Hidy, G. M.; Mueller, P. K.; Grosjean, D.; Appel, B. R.; Wesolowski, J. J., eds. (1980) The character and
37            origins of smog aerosols: a digest of results from the California Aerosol Characterization Experiment
38            (ACHEX).  New York, NY: John Wiley & Sons.
39
40      Hildemann, L. M.; Markowski, G. R.; Cass, G. R. (1991) Chemical composition of emissions from urban
41            sources  of fine organic aerosol. Environ. Sci. Technol. 25: 744-759.
42
43      Hildemann, L. M.; Cass,  G. R.; Mazurek, M. A.; Simoneit, B. R. T. (1993) Mathematical modeling of urban
44            organic  aerosol: properties measured by high-resolution gas chromatography.  Environ. Sci. Technol.
45            27: 2045-2055.
46
47      Hildemann, L. M.; Mazurek, M. A.; Cass, G. R.; Simoneit, B. R. T. (1994) Seasonal trends in Los Angeles
48            ambient organic aerosol observed by high-resolution gas chromatography. Aerosol Sci. Technol.
49            20: 303-317.
50
51      Hopke, P. K. (1985) Receptor modeling in environmental chemistry. New York, NY: John Wiley and Sons.
52
         April 1995                                   6-196       DRAFT-DO NOT QUOTE OR CITE

-------
  1
  2
  3
  4
  5
  6
  7
  8
  9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
Houck et al. (1992) Source apportionment—Pocatello, Idaho PM10 nonattainment area.  In: Chow; Ono, eds.
       PM10 standards and nontraditional paniculate source controls. Pittsburgh, PA: Air & Waste Management
       Association; p.  219.

Husar, R. B.  (1991) Historical visibility trends. In: Irving, P. M., ed. Acid deposition: state of science and
       technology, volume III: terrestrial, materials, health and visibility effects. Washington, DC: The National
       Acid Precipitation Assessment Program; pp. 24-68—24-76. (Trijonis, J.  C.; Malm, W. C.; Pitchford,
       M.; White, W.  H.  Visibility: existing and historical conditions—causes and effects: state of science and
       technology report no. 24).

Husar, R. B.; Frank, N. H. (1991) Interactive exploration and analysis of EPA's Aerometric  Information and
       Retrieval System (AIRS) data sets. Presented at: the 84th annual meeting of the Air and Waste
       Management Association; June; Vancouver,  BC, Canada.  Pittsburgh, PA: Air & Waste Management.

Husar, R. B.; Holloway, J. M. (1984) The properties and climate of atmospheric haze. In: Ruhnke, L. H.;
       Deepak, A., eds. Hygroscopic aerosols. Hampton, VA: A. Deepak Publishing.

Husar, R. B.; Poirot, R. (1992) Exploration of the AIRS database: weekly cycle of ozone and PM10 aerosols.
       Presented at: 85th annual meeting of the Air & Waste Management Association; June; Kansas  City, MO.
       Pittsburgh, PA: Air & Waste Management Association; paper no. 92-77.06.

Husar, R. B.; Stowe, L. L. (1994) Tropospheric aerosols  over the oceans, submitted.

Husar, R. B.; Wilson, W.  E., Jr. (1986) Climatic trends over North America: potential role of aerosols. In: Lee,
       S. D.; Schneider, T.; Grant, L. D.; Verkerk, P. J., eds. Aerosols: research, risk assessment and control
       strategies, proceedings of the second U.S.-Dutch international symposium; May 1985; Williamsburg,
       VA. Chelsea, MI: Lewis Publishers, Inc.; pp. 821-833.

Husar, R. B.; Wilson, W.  E. (1993) Haze and sulfur emission trends in the eastern United States. Environ. Sci.
       Technol. 27: 12-16.

Husar, R. B.; Patterson, D. E.; Holloway, J. M.; Wilson, W. E.; Ellestad, T. G. (1979)  Trends of eastern U.S.
       haziness since 1948. In: Preprints for the fourth symposium on  turbulence, diffusion,  and air pollution;
       January;  Reno,  NV. Boston, MA: American Meteorological Society; pp. 249-256.

Husar, R. B.; Holloway, J. M.; Patterson,  D. E.; Wilson, W. E. (1981)  Spatial and temporal pattern of eastern
       U.S. haziness: a summary. Atmos.  Environ. 15: 1919-1928.

Husar, R. B.; Oberman, T.; Hutchins, E. A.  (1990) Environmental Informatics: implementation through the
       Voyager data exploration software.  Presented at: the 83rd annual  meeting of the Air and Waste
       Management Association; June; Pittsburgh, PA. Pittsburgh,  PA: Air & Waste Management Association.

Husar, R. B.; Elkins, J. B.; Wilson, W.  E. (1994) US visibility trends, 1960-1992. Presented at: the 87th annual
       meeting of the Air  and Waste Management Association; June; Cincinnati, OH. Pittsburgh, PA: Air &
       Waste Management Association.

Irving, P. M., ed. (1991) Acid deposition: state of science and technology, volume III: terrestrial, materials,
       health and visibility effects. Washington, DC: The National  Acid  Precipitation Assessment Program.

Jaenicke, R. (1980) Washington, DC: National Academy of Science; pp.  97.

Javitz, H. S. (1988) Feasibility study of receptor modeling for apportioning utility contributions to air
       constituents, deposition quality and  light extinction [draft report].  Palo Alto,  CA: Electric Power
       Research Institute.
         April  1995
                                                 6-197
DRAFT-DO NOT QUOTE OR CITE

-------
  1     John, W.; Kaifer, R.; Rahn, K.; Wesolowski, J. J. (1973) Trace element concentrations in aerosols from the San
  2            Francisco Bay area. Atmos. Environ. 7:  107-118.
  3
  4     John, W.; Wall, S.  M.; Ondo, J. L. (1988) A new method for nitric acid and nitrate aerosol measurement using
  5            the dichotomous sampler. Atmos. Environ. 22:  1627-1635.
  6
  7     John, W.; Wall, S.  M.; Ondo, J. L.;  Wmklmayr, W. (1989) Acidic aerosol size distributions during SCAQS.
  8            Sacramento, CA: California Air Resources Board.
  9
10     John, W.; Wall, S.  M.; Ondo, J. L.;  Winklmayr, W. (1990) Modes in the size distributions of atmospheric
11            inorganic aerosol. Atmos. Environ. Part A 24: 2349-2359.
12
13     Johnson, R. L.; Shah, J. J.; Gary, R. A.; Huntzicker, J. J. (1981) An automated thermal-optical method for the
14            analysis of carbonaceous aerosol. In:  Macias, E. S.; Hopke, P. K., eds. Atmospheric aerosol:  source/air
15            quality relationships; based on a symposium jointly sponsored by the Divisions of Nuclear Chemistry and
16            Technology at the 180th national meeting of the American Chemical  Society; August 1980; Las Vegas,
17            NV. Washington, DC: American Chemical Society; pp. 223-233. (Comstock, M. J., ed.  ACS
18            symposium  series:  167).
19
20     Johnson, B. J.; Huang, S. C.; LeCave, M.; Porterfield, M. (1994) Seasonal  trends of nitric acid, paniculate
21            nitrate, and  paniculate sulfate concentrations at  a southwestern U.S. mountain site.  Atmos.  Environ.
22            28: 1175-1179.
23
24     Joseph, J. H.; Metza, J.; Malm, W. C.; Pitchford, M.  L. (1987) Plans for IMPROVE: a federal program to
25            monitor visibility in Class I areas. In: Bhardwaja, P. J., ed. Visibility protection: research and policy
26            aspects. Pittsburgh, PA: Air Pollution Control Association.
27
28     Kadowaki, S.  (1976) Size distribution of atmospheric total aerosols, sulfate, ammonium and nitrate particulates in
29            the Nagoya  area. Atmos. Environ.  10: 39-43.
30
31     Kao, A. S.; Friedlander, S. K. (1994) Chemical  signatures of the Los Angeles aerosol (d,2<3.5 /zm). Aerosol
32            Sci. Technol. 21: 283-293.
33
34     Kao, A. S.; Friedlander, S. K. (1995) Frequency distributions of PM10 chemical components and their sources.
35            Environ. Sci. Technol. 29: 19-28.
36
37     Karch, R.; Kruisz, C.; Reischl, G. P.; Winklmayr, W.  (1987) Continuous measurement of ultrafine particle size
38            distributions in the range from 2 to 200 nm from the Southern California Air Quality Study, June-July
39            1987 [preliminary report]. Vienna,  Austria: Institut fur Experimentalphysik der Universitat Wein.
40
41     Keeler, G. J.; Spengler, J. D.; Koutrakis, P.; Allen, G. A.; Raizenne, M.; Stern,  B. (1990) Transported acid
42            aerosols measured in southern Ontario. Atmos.  Environ. Part A 24: 2935-2950.
43
44     Keeler, G. J.; Spengler, J. D.; Castillo, R. A. (1991) Acid aerosol measurements  at  a suburban Connecticut site.
45            Atmos. Environ. Part  A 25: 681-690.
46
47     Kim et al. (1992) Source apportionment study for state implementation plant development in the Coachella
48            Valley. In: Chow; Ono, eds. PM10 standards  and nontraditional paniculate source controls. Pittsburgh,
49            PA: Air & Waste Management Association; p. 979.
50
51     Kinney, P. L.; Ozkaynak, H.  (1991) Associations of daily mortality and air pollution in Los Angeles County.
52            Environ. Res. 54: 99-120.
53
         April 1995                                    6-198       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Kleinman, M. T.; Tomczyk, C.; Leaderer, B. P.; Tanner, R. L. (1979) Inorganic nitrogen compounds in New
  2            York City air. Ann. N.  Y. Acad.  Sci. 322: 114-123.
  3
  4     Knapp, K. T.; Ellestad, T. G. (1990) Nitric acid and paniculate measurements by transition-flow reactor during
  5            the carbonaceous species methods comparison study. Aerosol Sci. Technol. 12: 39-43.
  6
  7     Koutrakis, P.; Kelly, B. P. (1993) Equilibrium size of atmospheric aerosol sulfates as a function of particle
  8            acidity and ambient relative humidity. J. Geophys. Res. [Atmos.] 98: 7141-7147.
  9
10     Koutrakis, P.; Spengler, J. D. (1987) Source apportionment of ambient particles in Steubenville, OH using
11            specific rotation factor analysis. Atmos. Environ.  21: 1511-1519.
12
13     Koutrakis, P.; Wolfson, J. M.;  Spengler, J. D.; Stern, B.; Franklin, C.  A. (1989) Equilibrium size of
14            atmospheric aerosol sulfates as a function of the relative humidity. J. Geophys. Res. [Atmos.]
15            94: 6442-6448.
16
17     Lamborg, C.; Keeler, G. J.; Evans, G. (1992) Atmospheric acidity  measurements during the Lake Michigan
18            Urban Air Toxics Study. Research Triangle Park, NC: U.S. Environmental Protection Agency,
19            Atmospheric Research and Exposure Assessment Laboratory; EPA report no.  EPA/600/A-92/250.
20            Available from: NTIS,  Springfield, VA; PB93-121069.
21
22     Larson, S. M.; Cass, G. R. (1989) Characteristics of summer midday low-visibility events in the Los Angeles
23            area. Environ. Sci. Technol. 23: 281-289.
24
25     Larson, S. M.; Cass, G. R.; Gray, H. A. (1989) Atmospheric carbon particles and the Los Angeles visibility
26            problem. Aerosol Sci. Technol. 10: 118-130.
27
28     Leaderer, B. P.; Bernstein, D.  M.; Daisey, J. M.; Kleinman, M. T.; Kneip, T. J.; Knutson, E. O.; Lippmann,
29            M.; Lioy, P. J.; Rahn, K. A.; Sinclair, D.; Tanner, R. L.;  Wolff, G. T. (1978) Summary of the New
30            York Summer Aerosol Study (NYSAS). J. Air Pollut. Control Assoc.  28: 321-327.
31
32     Leaderer, B. P.; Tanner, R. L.; Holford,  T. R. (1982) Diurnal variations, chemical composition and relation to
33            meteorological variables of the summer aerosol in the New York subregion. Atmos. Environ.
34            16: 2075-2087.
35
36     Lee, H. S.; Wadden, R. A.; Scheff,  P. A. (1993) Measurement and evaluation of acid air pollutants in Chicago
37            using an annular denuder system. Atmos. Environ. Part A 27: 543-553.
38
39     Lewis, C. W.; Dzubay, T. G. (1986) Measurement of light absorption extinction in Denver. Aerosol Sci.
40            Technol. 5:  325-336.
41
42     Lewis, C. W.; Macias, E. S. (1980) Composition of size-fractionated aerosol in Charleston, West Virginia.
43            Atmos. Environ. 14: 185-194.
44
45     Lewis, C. W.; Baumgardner, R. E.;  Stevens, R. K.;  Russwurm, G. M.  (1986) Receptor modeling study of
46            Denver winter haze. Environ. Sci. Technol. 20: 1126-1136.
47
48     Lewis, C. W. et al.  (1989) Comparison of three measures of visibility extinction in Denver, CO. Presented at:
49            the 82nd annual meeting of the Air Pollution Control Association; June; Anaheim, CA. Pittsburgh, PA:
50            Air Pollution Control Association; paper no. 89-151.4.
51
52     Liang, C. S. K.; Waldman, J. M. (1992) Indoor exposures to acidic aerosols at child and elderly care facilities.
53            Indoor Air 2: 196-207.
54


         April 1995                                    6-199       DRAFT-DO NOT QUOTE OR CITE

-------
 1      Lioy, P. J.; Waldman, J. M. (1989) Acidic sulfate aerosols: characterization and exposure. In: Symposium on
 2             the health effects of acid aerosols; October 1987; Research Triangle Park, NC. Environ. Health Perspect.
 3             79:  15-34.
 4
 5      Lioy, P. J.; Samson, P. J.; Tanner, R. L.; Leaderer, B. P.; Minnich, T.; Lyons, W. (1980) The distribution and
 6             transport of sulfate "species" in the New York metropolitan area during the 1977 Summer Aerosol Study.
 7             Atmos. Environ. 14: 1391-1407.
 8
 9      Lioy, P. J.; Daisey, J. M.; Greenberg, A.; Harkov, R. (1985) A major wintertime (1983) pollution episode in
10             northern New Jersey: analyses of the accumulation and spatial distribution of inhalable paniculate matter,
11             extractable organic matter and other species. Atmos. Environ. 19: 429-436.
12
13      Lipfert, F. W.; Wyzga, R. E. (1993) On the spatial and temporal variability of aerosol acidity and sulfate
14             concentration. Air Waste 43: 489-491.
15
16      Lundgren, D.  A. (1967) An aerosol sampler for determination of particle concentration as a function of size and
17             time. J. Air Pollut. Control Assoc. 17: 225-229.
18
19      Macias, E. S.; Zwicker, J. O.; Ouimette, J. R.; Hering, S. V.; Friedlander, S. K.; Cahill, T. A.;  Kuhlmey, G.
20             A.;  Richards, L. W. (1981) Regional haze case studies in the southwestern U.S.—I. aerosol chemical
21             composition. In: White, W. H.; Moore, D. J.; Lodge, J.  P., Jr.,  eds. Plumes and visibility:
22             measurements and model components, proceedings of the symposium; November 1980; Grand Canyon
23             National Park, AZ. Atmos. Environ. 15: 1971-1986.
24
25      Macias, E. S.; Zwicker, J. O.; White, W. H. (1981) Regional haze case  studies in the southwestern U.S.—II.
26             source contributions. In: White, W. H.; Moore, D. J.; Lodge, J.  P., Jr., eds. Plumes and visibility:
27             measurements and model components: proceedings of the symposium; November 1980; Grand Canyon
28             National Park, AZ. Atmos. Environ. 15: 1987-1997.
29
30      Macias, E. S.; Vossler, T. L.;  White, W. H. (1987) Carbon and sulfate particles in the western  U.S. In:
31             Bhardwaja, P. J., ed. Visibility protection: research and policy aspects. Pittsburgh, PA: Air Pollution
32             Control Association.
33
34      Maenhaut, W.; Royset, 0.; Vadset, M.;  Kauppinen, E. I.; Lind,  T.  M. (1993) Nucl. Instrum. Methods
35             B75: 266-272.
36
37      Malissa, H. (1979) Some analytical approaches to the chemical characterization of carbonaceous  particulates. In:
38             Novakov, T., ed. Proceedings: carbonaceous particles in the atmosphere; March 1978; Berkeley, CA.
39             Berkeley, CA: University of California, Lawrence  Berkeley Laboratory; pp. 3-9; report no. LBL-9037.
40
41      Malm, W. C.  et al. (1987) Comparison of atmospheric extinction measurements made by a transmissometer,
42             integrating nephelometer, and teleradiometer with natural  and artificial black targets. In:  Bhardwaja, P.
43             J., ed. Visibility protection: research and policy aspects. Pittsburgh,  PA: Air Pollution Control
44             Association.
45
46      Malm, W. C.  et al. (1987) An eigenvector analysis of paniculate data in  the western United States. In:
47             Bhardwaja, P. J.,  ed. Visibility protection: research and policy aspects. Pittsburgh, PA.: Air Pollution
48             Control Association.
49
50      Malm, W.;  Gebhard,  K.; Cahill, T.; Eldred, R.; Pielke, R.; Watson, J.;  Latimer, D. (1989) Winter haze
51             intensive experiment: draft final report. Fort Collins, CO: National Park Service.
52
53      Malm, W. C.; Gebhart, K. A.; Henry, R. C. (1990) An investigation of  the dominant source regions of fine
54             sulfur in the western United States and their areas of influence. Atmos. Environ. Part A  24: 3047-3060.


         April 1995                                   6-200       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Malm, W. C.; Gebhart, K. A.; Molenar, J.; Cahill, T.; Eldred, R.; Huffman, D. (1994) Examining the
  2            relationship between atmospheric aerosols and light extinction at Mount Rainier and North Cascades
  3            National Parks. Atmos. Environ. 28: 347-360.
  4
  5     Malm, W. C.; Sisler, J. F.; Huffman, D.; Eldred, R. A.;  Cahill, T. A. (1994) Spatial and seasonal trends in
  6            particles concentration and optical extinction in the United States. J. Geophys. Res. 99: 1347-1370.
  7
  8     Mamane, Y.; Noll, K. E. (1985) Characterization of large particles at a rural site in the eastern United States:
  9            mass distribution and individual particle analysis. Atmos. Environ. 19: 611-622.
 10
 11     ManTech. (1994) LRGVES report. Research Triangle Park, NC: ManTech.
 12
 13     Mangelson, N. F.; Rees, L. B.; Lewis, L.; Joseph, J. M.; Cui, W.; Machir, J.;  Eatough, D. J.; Wilkerson, T.
 14            D.; Jensen, D. T. (1994) The formation of sulfate  and nitrate during winter inversion fogs in Cache
 15            Valley, Utah. In: Proceedings of an international specialty conference on aerosol and atmospheric optics:
 16            radiative balance  and visual air quality: v. A.  Pittsburgh, PA: Air and Waste Management Association.
 17
 18     Marple, V.; Rubow, K.;  Anath, G.; Fissan, H. J. (1986) J. Aerosol Sci. 17: 489-494.
 19
 20     Marshall, B. T.; Patterson, E. M.; Grams, G. W. (1986) Characterization of the Atlanta area aerosol elemental
 21            composition and possible sources. Atmos. Environ. 20: 1291-1300.
 22
 23     McDade, C. E.; Tombach, I. H. (1987) Goals and initial findings from SCENES. In: Bhardwaja, P. J., ed.
 24            Visibility protection: research and policy aspects. Pittsburgh, PA: Air Pollution Control Association.
 25
 26     McLaughlin, S. B.; Schorn, V. J.; Jones, H. C. (1976) A programmable exposure system for kinetic dose-
 27            response studies with air pollutants. J. Air Pollut. Control Assoc. 26: 132-135.
 28
 29     McMurry, P.  (1989) Organic and  elemental carbon size distributions of Los Angeles aerosols measured during
 30            SCAQS [final report]. Atlanta, GA: Coordinating Research Council; CRC project no. SCAQS-6-1;
 31            Particle Technology Laboratory report no. 713.
 32
 33     McMurry, P.  H.; Stolzenburg, M. R. (1989) On the sensitivity of particle size to relative humidity for Los
 34            Angeles aerosols. Atmos. Environ. 23: 497-507.
 35
 36     McMurry, P.  H.; Zhang, X. Q. (1989) Size distributions of ambient organic and elemental carbon. Aerosol Sci.
 37            Technol. 10: 430-437.
 38
 39     Mercer, T. T. (1964) The stage constants of cascade impactors. Ann. Occup. Hyg. 7: 115-124.
40
41     Milford, J. B.; Davidson, C.  I. (1985) The sizes of particulate trace elements in the atmosphere—a review. J.
42            Air Pollut. Control Assoc. 35: 1249-1260.
43
44     Miller, D. F.; Schorran, D. E.; Hoffer, T. E.; Rogers, D. P.; White, W. H.;  Macias, E. S. (1990) An analysis
45            of regional haze using tracers of opportunity.  J. Air Waste Manage. Assoc. 40:  757-761.
46
47     Molenar, J. V.; Dietrich, D. L.; Cismoski, D. S.; Cahill, T. A.; Wakabayashi, P. H. (1994) In: Proceedings of
48            an international specialty conference on aerosol and atmospheric optics: radiative balance and visual air
49            quality, volume A. Pittsburgh, PA: Air & Waste Management Association.
50
51     Morandi, M. T.; Kneip, T. J.; Cobourn,  W. G.; Husar, R. B.; Lioy, P. J. (1983) The measurement of H2SO4
52            and other sulfate species at Tuxedo, New York with a thermal analysis flame photometric detector and
53            simultaneously collected quartz filter samples. Atmos.  Environ. 17: 843-848.
54


         April  1995                                    6-201        DRAFT-DO NOT QUOTE OR CITE

-------
 1     Moskowitz, A. H. (1977) Particle size distribution of nitrate aerosols in the Los Angeles Air Basin. Research
 2             Triangle Park, NC: U.S.  Environmental Protection Agency, Environmental Sciences Research
 3             Laboratory; EPA report no. EPA-600/3-77-053. Available from: NTIS, Springfield, VA; PB-269 349.
 4
 5     Mosteller, P.; Tukey, J. W. (1977) Data analysis and regression. Reading, PA: Addison-Wesely Publishing
 6             Company.
 7
 8     Moyers, J. L. (1981) 1981-1982 aerosol studies for nine southwest national parks and monuments [final report].
 9             Washington, DC: U.S. Department of the Interior, National Park Service.
10
11     Mueller, P. K.; Hansen, D. A.; Watson, J. G., Jr.  (1986) The Subregional Cooperative Electric Utility,
12             Department of Defense, National Park Service, and EPA Study (SCENES) on visibility: an overview.
13             Palo Alto, CA: Electric Power Research Institute; report no. EA-4664-SR.
14
15     Mukerjee et al.  (1993) A methodology to apportion ambient air measurements to investigate potential effects on
16             air quality near waste  incinerators. In: Proceedings of the 1993 incineration conference;  Knoxville, TN.
17             p. 527.
18
19     Mylonas, D.  T.; Allen, D. T.; Ehrman, S. H.; Pratsinis,  S. E. (1991) The sources and size distributions of
20             organonitrates in Los Angeles aerosol. Atmos. Environ. Part A 25: 2855-2861.
21
22     National Research Council. (1993) Protecting visibility in national parks and wilderness areas. Washington, DC:
23             National Academy of Science.
24
25     Natusch, D. F.  S.; Wallace, J. R. (1974) Urban aerosol toxicity: the influence of particle size. Particle size,
26             adsorption, and respiratory deposition profiles combine to determine aerosol toxicity. Science
27             (Washington, DC) 186: 695-699.
28
29     Natusch, D. F.  S.; Wallace, J. R.; Evans, C. A., Jr. (1974) Toxic trace elements: preferential concentration in
30             respirable particles. Science (Washington, DC) 183: 202-204.
31
32     Newman, J. E.;  Abel. M. D.; Harrison, P. R.; Yost, K. J. (1976) Wind as related to critical flushing speed
33             versus reflotation speed by high-volume sampler paniculate loading. In: Atmospheric-surface exchange of
34             particles and gaseous pollutants. Washington, DC: U.S. Department of Energy; pp. 466-496. Available
35             from: NTIS, Springfield,  VA; CONF 740-921.
36
37     Nifong, G. D. (1970) [Ph.D. thesis]. Ann Arbor, MI: Univeristy of Michigan.
38
39     Noll, K. E.; Draftz, R.; Fang, K. Y.  P. (1987) The composition of atmospheric coarse particles at an urban and
40             non-urban site.  Atmos. Environ. 21: 2717-2721.
41
42     Noll, K. E.; Yuen, P.-F.; Fang, K. Y.-P. (1990) Atmospheric coarse paniculate concentrations  and dry
43             deposition fluxes for ten metals in two urban environments. Atmos. Environ. Part A 24: 903-908.
44
45     Ouimette, J. R.; Flagan, R. C.; Kelso, A. R. (1981) Chemical species contributions to light scattering by
46             aerosols at a remote arid site:  comparison of statistical and theoretical results. In: Macias, E. S.; Hopke,
47             P.  K., eds. Atmospheric aerosols: source/air quality relationships;  based on a symposium jointly
48             sponsored by the Divisions of Nuclear Chemistry  and Technology  and Environmental Chemistry at the
49             180th national meeting of the  American Chemical Society; August  1980; Las Vegas, NV. Washington,
50             DC: American  Chemical  Society; pp. 125-156. (Comstock, M. J.,  ed. ACS symposium series: 167).
51
52     Patterson, R. K.; Wagman, J. (1977)  Mass and composition of an urban aerosol as a function of particle size for
53             several  visibility levels. J. Aerosol Sci. 8: 269-279.
54


         April 1995                                    6-202       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Patterson, D. E.; Holloway, J. ML; Husar, R. B. (1980) Historical visibility over the eastern U.S.: daily and
  2            quarterly extinction coefficient contour maps. Research Triangle Park, NC: U.S. Environmental
  3            Protection Agency, Environmental Sciences Research Laboratory; EPA report no. EPA-600/3-80-043a.
  4            Available from: NTIS, Springfield, VA; PB81-196974.
  5
  6     Patterson, D. E.; Husar, R.  B.; Wilson, W. E.; Smith, L. F.  (1981) Monte Carlo simulation of daily regional
  7            sulfur distribution: comparison with SURE sulfate data and visual range observations during August
  8            1977. J. Appl. Meteorol. 20: 404-420.
  9
 10     Patterson, R. E.; Bernier,  S. A.; Gregory, J. (1994) Blowing  dust across the Texas high plains: spatial and
 11            temporal variations.  In: Proceedings of the international specialty conference on aerosol and atmospheric
 12            optics: radiative balance and visual air quality, volume A. Pittsburgh, PA; Air & Waste Management
 13            Association.
 14
 15     Pickle, T.; Allen,  D. T.; Pratsinis, S. E. (1990) The sources and size distributions of aliphatic and carbonyl
 16            carbon in Los Angeles aerosol. Atmos. Environ. Part A 24: 2221-2228.
 17
 18     Pierson, W. R.; Brachaczek, W. W. (1988) Coarse- and fine-particle atmospheric nitrate and HNO3(g) in
 19            Claremont, California, during the  1985 Nitrogen Species Methods Comparison  Study. Atmos. Environ.
 20            22:  1665-1668.
 21
 22     Pierson, W. R.; Russell, P. A. (1979) Aerosol carbon in the Denver area in November 1973.  Atmos. Environ.
 23            13:  1623-1628.
 24
 25     Pierson, W. R.; Brachaczek, W. W.; Korniski, T. J.; Truex, T.  J.; Butler, J. W. (1980) Artifact formation of
 26            sulfate, nitrate, and hydrogen ion on backup filters: Allegheny Mountain experiment. J. Air Pollut.
 27            Control Assoc. 30: 30-34.
 28
 29     Pinto, J. (n.d.) [Unpublished data].
 30
 31     Poirot, R. L.; Flocchini, R.  G.; Husar, R. B. (1990) Winter fine particle composition in the northeast:
 32            preliminary results from the NESCAUM monitoring network. Presented at: 83rd annual meeting of the
 33            Air & Waste Management Association; June; Pittsburgh, PA. Pittsburgh, PA: Air & Waste Management
 34            Association; paper no.  90-84.5.
 35
 36     Pope, C. A., Ill; Schwartz, J.; Ransom, M. R. (1992) Daily mortality and PM10 pollution in Utah valley. Arch.
 37            Environ. Health 47: 211-217.
 38
 39     Raabe, O. G.; Braaten, D. A.; Axelbaum, R. L.; Teague, S. V.; Cahill, T. A. (1988) J. Aerosol Sci.
 40            192: 183-195.
 41
42     Rahn, K. A.; Harrison, P. R. (1976)  The chemical composition of Chicago street dust.  In: Atmosphere-surface
43            exchange of paniculate and gaseous pollutants (1974):  proceedings of a  symposium;  September 1974;
44            Richland, WA.. Oak Ridge, TN: Energy Research and Development Administration; pp. 557-570.
45            (ERDA symposium series 38). Available from: NTIS,  Springfield, VA;  CONF 740921.
46
47     Rahn, K. A.; McCaffrey, R. J. (1980) On the origin and transport of the winter Arctic  aerosol. Ann. N. Y.
48            Acad.  Sci. 388: 503.
49
50     Rahn, K. A.; Wesolowski, J. J.; John, W.; Ralston, H. R. (1971) Diurnal variation of aerosol trace element
51            concentrations in Livermore, California. J. Air Pollut.  Control Assoc. 21: 406-409
52
53     Richards, L. W. (1983) Comments on the  oxidation of NO2 to  nitrate—day and  night. Atmos. Environ
54            17: 397-402.


         April 1995                                   6_203       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Richards, L. W.; Markowski, G. R.; Waters,  N. (1981) Comparison of nephelometer, telephotometer, and
  2            aerosol data in the southwest. Presented at: the 74th annual meeting of the Air Pollution Control
  3            Association; June; Philadelphia, PA. Pittsburgh, PA: Air Pollution Control Association; paper
  4            no. 81-54.5.
  5
  6     Rodes, C. E.; Evans, E. G. (1985) Preliminary assessment of 10 /urn paniculate sampling at eight locations in the
  7            United States. Atmos. Environ. 19: 293-303.
  8
  9     Rolph, G. D.; Draxler,  R. R.; de Pena, R. G. (1992) Modeling sulfur concentrations and depositions in the
 10            United States during ANATEX. Atmos. Environ. Part A 26: 73-93.
 11
 12     Russell, A. G.; Cass, G. R. (1984) Acquisition of regional air quality model validation data for nitrate, sulfate,
 13            ammonium ion and their precursors. Atmos. Environ. 18:  1815-1827.
 14
 15     Saldiva, P. H. N.; Lichtenfels, A. J. F. C.; Paiva, P. S. O.; Barone, I.  A.; Martins, M. A.; Massad, E.;
 16            Pereira, J. C. R.; Xavier, V. P.; Singer, J. M.; Bohm, G. M. (1994) Association between air pollution
 17            and mortality due to respiratory diseases in children in Sao Paulo,  Brazil: a preliminary report. Environ.
 18            Res. 65: 218-225.
 19
 20     Savoie, D. L.;  Prospero, J. M.  (1982) Particle size distribution of nitrate and sulfate in the marine atmosphere.
 21            Geophys. Res.  Lett. 9: 1207-1210.
 22
 23     Saxena, P. () [Personal communication]. Palo Alto, CA: Electric Power  Research Institute.
 24
 25     Schichtel, B. A.;  Husar, R. B. (1991) Apportionment of light extinction by aerosol types. Air Force Systems
 26            Command, Phillips Laboratory.
 27
 28     Schroeder, W. H.; Dobson, M.; Kane, D. M.; Johnson, N. D. (1987) Toxic trace elements associated with
 29            airborne paniculate matter: a review. JAPCA 37:  1267-1285.
 30
 31     Schwartz, J. (1994) Air pollution and hospital admissions for the elderly in Birmingham,  Alabama. Am. J.
 32            Epidemiol. 139: 589-598.
 33
 34     Schwartz, J.; Dockery, D. W. (1992) Increased mortality in Philadelphia associated with  daily air pollution
 35            concentrations. Am.  Rev. Respir. Dis. 145: 600-604.
 36
 37     Seaton, A.; MacNee, W.; Donaldson, K.; Godden, D. (1995) Paniculate air pollution and acute health effects.
 38            Lancet (8943):  176-178.
 39
40     Seeker, W. R. (1990) In: Waste combustion: proceedings of the 23rd international symposium on combustion;
41            pp. 867-885.
42
43     Senior, C. L.; Flagan, R. C. (1982) Ash vaporization and condensation during combustion of a suspended coal
44            particle. Aerosol Sci. Technol.  1: 371-383.
45
46     Shah, J. J.; Watson, J. G., Jr.; Cooper, J. A.; Huntzicker, J. J. (1984) Aerosol chemical composition and light
47            scattering in Portland, Oregon: the role of carbon. Atmos.  Environ. 18: 235-240.
48
49     Shaw, R. W., Jr.; Paur, R. J. (1983) Composition of aerosol particles collected at rural sites in the Ohio River
50            Valley. Atmos. Environ. 17: 2031-2044.
51
52     Simoneit, B. R. T.; Mazurek, M. A. (1982) Organic  matter of the troposphere—II. natural background of
53            biogenic lipid matter in aerosols over the rural western United States. Atmos.  Environ. 16: 2139-2159.
54


         April 1995                                    6-204       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Sinclair, J. D.; Psota-Kelty, L. A.; Weschler, C. J.; Shields, H. C. (1990) Measurement and modeling of
  2            airborne concentrations and indoor surface accumulation rates of ionic substances at Neenah, Wisconsin.
  3            Atmos. Environ. Part A 24: 627-638.
  4
  5     Sirois, A.; Fricke, W. (1992) Regionally representative daily air concentrations of acid-related substances in
  6            Canada; 1983-1987. Atmos. Environ. Part A 26:  593-607.
  7
  8     Sisler, J. F.; Malm, W. C. (1994) The relative importance of soluble aerosols to spatial and seasonal trends of
  9            impaired visibility in the United States. Atmos. Environ. 28: 851-862.
10
11     Sisler, J. F.; Huffman, D.; Lattimer, D. A.; Malm, W. C.;  Pitchford, M. (1993) Spatial and temporal patterns
12            and the chemical composition of the haze in the United States: an anlysis of data from the IMPROVE
13            network, 1988-1991. Fort Collins,  CO: Colorado State University, CIRA.
14
15     Skidmore,  L. W.;  Chow, J. C.; Tucker, T. T.  (1992) PM10 air quality assessment for the Jefferson County,
16            Ohio air quality control region. In: Chow, J. C.;  Ono, D. M., eds. Transactions: PM10 standards and
17            nontraditional paniculate source controls. Pittsburgh, PA: Air & Waste  Management Association;
18            pp. 1016-1031.
19
20     Sloane, C. S. (1982) Visibility trends—II. mideastern United States 1948-1978.  Atmos. Environ.  16: 2309-2321.
21
22     Sloane, C. S. (1982) Visibility trends—I. methods of analysis. Atmos. Environ. 16:  41-51.
23
24     Sloane, C. S. (1986) Effect of composition on aerosol light scattering efficiencies. Atmos. Environ.
25            20: 1025-1037.
26
27     Sloane, C. S.; Wolff, G. T. (1985) Prediction of ambient light scattering using a physical model responsive to
28            relative humidity: validation with measurements from Detroit. Atmos. Environ. 19:  669-680.
29
30     Sloane, C. S.; Watson, J.; Chow, J.; Pritchett, L.; Richards, L. W. (1991) Size-segregated fine particle
31            measurements by chemical species  and their impact on visibility impairment in Denver. Atmos. Environ.
32            Part A 25: 1013-1024.
33
34     Small, M.; Germani, M. S.; Small, A.  M.; Zoller, W. H.; Moyers, J. L. (1981) Airborne plume study of
35            emissions from the processing of copper ores in southeastern Arizona. Environ. Sci. Technol.
36            15: 293-299.
37
38     Solomon, P. A.; Moyers, J. L. (1986) A chemical characterization  of wintertime haze in Phoenix, Arizona.
39            Atmos. Environ. 20: 207-213.
40
41     Solomon, P. A.; Larson, S.  M.; Fall, T.; Cass, G. R. (1988) Basinwide nitric acid and related species
42            concentrations observed during the  Claremont Nitrogen Species Comparison Study. Atmos. Environ.
43            22: 1587-1594.
44
45     Spengler, J. D.; Thurston, G. D. (1983) Mass and elemental composition of fine and coarse particles in six U.S.
46            cities. J. Air Pollut.  Control Assoc. 33: 1162-1171.
47
48     Spengler, J. D.; Allen, G. A.; Foster,  S.; Severance, P.;  Ferris, B., Jr. (1986)  Sulfuric acid and sulfate aerosol
49            events in two U.S. cities. In: Lee, S. D.; Schneider,  T.; Grant, L. D.; Verkerk, P.  J., eds. Aerosols:
50            research, risk assessment and control strategies—proceedings of the second U.S.-Dutch international
51            symposium; May 1985; Williamsburg, VA. Chelsea,  MI: Lewis Publishers,  Inc.; pp. 107-120.
52
         April 1995                                    6-205       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Spengler, J. D.; Keeler, G. J.; Koutrakis, P.; Ryan, P. B.; Raizenne, M.; Franklin, C. A. (1989) Exposures to
  2             acidic aerosols. In: Symposium on the health effects of acid aerosols; October 1987; Research Triangle
  3             Park, NC. Environ. Health Perspect.  79: 43-51.
  4
  5      Spengler, J. D.; Brauer, M.; Koutrakis, P. (1990) Acid air and health. Environ. Sci. Technol. 24:  946-956.
  6
  7      Spicer, C. W. (1977) The fate of nitrogen oxides in the atmosphere. In: Pitts, J. N., Jr.; Metcalf, R. L.; Lloyd,
  8             A. C., eds. Advances in environmental science and technology: v. 4. New York, NY: John Wiley  and
  9             Sons, Inc.; pp. 163-261.
10
11      Spicer, C. W.; Schumacher, P. M. (1979) Paniculate nitrate: laboratory and field studies of major sampling
12             interferences. Atmos. Environ. 13: 543-552.
13
14      Spix, C.; Heinrich, J.;  Dockery, D.; Schwartz, J.; Volksch, G.; Schwinkowski, K.; Collen, C.; Wichmann,  H.
15             E. (1993) Air pollution and daily mortality in Erfurt, East Germany, 1980-1989. Environ. Health
16             Perspect. 101: 518-526.
17
18      Stevens, R.  K., ed. (1979) Current methods to measure atmospheric nitric acid and nitrate artifacts: proceedings
19             of a workshop on measurement of atmospheric nitrates; October 1978; Southern Pines, NC. Research
20             Triangle Park, NC: U.S. Environmental Protection Agency, Environmental Sciences Research
21             Laboratory; EPA report no. EPA-600/2-79-051. Available from: NTIS, Springfield, VA; PB-294 098.
22
23      Stevens. (1985) Sampling and analysis  methods for use in source apportionment studies to determine impact of
24             wood burning on fine particle mass. Environ. Int. 11: 271.
25
26      Stevens et al. (n.d.) A comparison of air quality measurements in Roanoke, VA, and other integrated air cancer
27             project monitoring locations.
28
29      Stevens, R.  K.; Dzubay, T. G.; Shaw, R. W., Jr.; McClenny, W. A.; Lewis, C. W.; Wilson, W.  E. (1980)
30             Characterization of the aerosol in the  Great Smoky Mountains. Environ. Sci. Technol. 14:  1491-1498.
31
32      Stevens, R.  K.; Dzubay, T. G.; Lewis, C.  W.; Shaw, R. W., Jr. (1984) Source apportionment methods applied
33             to the determination of the origin of ambient aerosols that affect visibility in forested areas. Atmos.
34             Environ. 18: 261-272.
35
36      Struempler,  A. W. (1975) Trace element composition in atmospheric particulates during 1973 and the summer of
37             1974 at Chadron, Neb. Environ. Sci.  Technol. 9: 1164-1168.
38
39      Suh, H. H.; Spengler, J. D.; Koutrakis, P. (1992) Personal exposures to acid aerosols and ammonia. Environ.
40             Sci.  Technol. 26: 2507-2517.
41
42      Suh, H. H.; Koutrakis, P.; Spengler, J. D. (1993) Validation of personal exposure models for sulfate and aerosol
43             strong acidity. J. Air Waste Manage.  Assoc. 43: 845-850.
44
45      Suh, H. H.; Allen, G. A.; Koutrakis, P.; Burton, R. M. (1994a) Spatial variation in acidic sulfate and ammonia
46             concentrations within metropolitan Philadelphia. J. Air Waste Manage. Assoc.: submitted.
47
48      Suh, H. H.; Koutrakis, P.; Spengler, J. D. (1994b) The relationship between airborne acidity and ammonia in
49             indoor environments. J. Exposure Anal. Environ. Epidemiol. 4:  1-23.
50
51      Sunyer, J.; Anto, J. M.; Murillo, C.; Saez, M. (1991) Effects of urban air pollution on emergency room
52             admissions for chronic obstructive pulmonary disease. Am. J.  Epidemiol. 134:  277-286.
53
         April 1995                                    6-206       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Thanukos, L. C.; Miller, T.; Mathai, C. V.; Reinholt, D.; Bennett, J. (1992) Intercomparison of PM10 samplers
  2            and source apportionment of ambient PM10 concentrations at Rillito, Arizona. In: Chow, J. C.; Ono, D.
  3            M., eds. PM]0 standards and non-traditional paniculate source controls: proceedings of an AWMA/EPA
  4            international specialty conference. Pittsburgh, PA: Air & Waste Management Association;  pp. 244-261.
  5
  6     Thompson, K. M.; Koutrakis, P.; Brauer, M.; Spengler, J. D.; Wilson, W. E.; Burton, R. M. (1991)
  7            Measurements of aerosol acidity: sampling frquency,  seasonal variability, and spatial variation. Presented
  8            at: the 84th annual meeting of the Air & Waste Management Association; June; Vancouver, BC, Canada.
  9            Pittsburgh, PA: Air & Waste Management Association; paper no. 91-89.5.
 10
 11     Thurston, G. D.; Gorczynski, J.  E.; Jaques, P.; Currie, J.; He, D.  (1992) An automated sequential sampling
 12            system for paniculate aerosols: description, characterization, and field sampling results. J.  Exposure
 13            Anal. Environ. Epidemiol. 2: 415-428.
 14
 15     Tombach,  I. H. et al.  (1987) SCENES light extinction methods study, September -October 1985. In: Bhardwaja,
 16            P.  J., eds. Visibility protection: research and policy aspects. Pittsburgh, PA: Air  Pollution Control
 17            Association.
 18
 19     Trijonis, J. (1979) Visibility in the southwest—an exploration of the historical data base. Atmos. Environ.
 20            13: 833-843.
 21
 22     Trijonis, J.; McGown, M.;  Pitchford, M.; Blumenthal, D.; Roberts, P.; White, W.; Macias, E.; Weiss, R.;
 23            Waggoner, A.; Watson, J.; Chow, J.; Flocchini, R. (1988)  Visibility conditions and causes of visibility
 24            degradation in the Mojave Desert of California: executive summary, RESOLVE project final report.
 25            China Lake, CA: Department of the Navy, Naval Weapons  Center; document no. NWC TP 6869.
 26            Available from: NTIS,  Springfield, VA; AD-A206 322.
 27
 28     Turpin, B. J.; Huntzicker, J. J. (1991) Secondary formation of organic aerosol in the Los Angeles basin: a
 29            descriptive analysis  of organic and elemental carbon concentrations. Atmos. Environ. Part  A
 30            25: 207-215.
 31
 32     Turpin, B. J.; Gary, R. A.; Huntzicker, I. J. (1990) An in situ, time-resolved analyzer for aerosol organic and
 33            elemental carbon. Aerosol Sci.  Technol. 12: 161-171.
 34
 35     Turpin, B. J.; Huntzicker, J. J.; Larson, S. M.; Cass, G. R.  (1991) Los Angeles summer midday  particulate
 36            carbon: primary and secondary aerosol.  Environ. Sci. Technol. 25: 1788-1793.
 37
 38     U.S. Environmental Protection Agency. (1982) Air quality criteria for particulate matter and sulfur oxides.
 39            Research Triangle Park, NC: Office of Health and Environmental Assessment, Environmental Criteria
 40            and Assessment Office; EPA report no.  EPA-600/8-82-029aF-cF. 3v. Available from: NTIS, Springfield,
 41            VA; PB84-156777.
 42
 43     U.S. Environmental Protection Agency. (1989) An acid aerosols issue paper:  health effects and aerometrics.
44            Research Triangle Park, NC: Office of Health and Environmental Assessment, Environmental Criteria
45            and Assessment Office; EPA report no.  EPA-600/8-88-005F. Available from:  NTIS, Springfield, VA;
46            PB91-125864.
47
48     U.S. Environmental Protection Agency. (1989) Washington, DC: Office of Solid Waste and Emergency
49            Response.
50
51     U.S. Environmental Protection  Agency. (1991) MOHAVE.
52
53     U.S. Environmental Protection Agency. (1995) Visibility research related to source assessment {draft report].
54


         April 1995                                   6-207        DRAFT-DO NOT QUOTE OR CITE

-------
  1      Van Grieken, R. G.; Johansson, J.; Winchester, J.; Odom, A. L. (1975) Anal. Chem. 275: 343.
  2
  3      Vasconcelos, L. A.; Macias, E.  S.; White, W.  H.  (1993) Aerosol composition as a function of haze and
  4            humidity levels in the southwestern U.S. [personal communication].
  5
  6      Vermette et al. (1992) PMJO source apportionment  using local surface dust profiles: examples from Chicago. In:
  7            Chow; Ono, eds. PM10 standards and nontraditional particulate source controls. Pittsburgh, PA: Air &
  8            Waste Management  Association; p. 262.
  9
10      Vossler, T. L.;  Lewis,  C. W.; Stevens, R. K.;  Dzubay, T. G.; Gordon, G. E.; Tuncel, S. G.; Russwurm, G.
11            M.; Keeler, G, J. (1989) Composition and  origin of summertime air pollutants at Deep Creek Lake,
12            Maryland. Atmos. Environ. 23: 1535-1547.
13
14      Waldman, J. M.; Lioy, P. J.; Thurston, G. D.; Lippmann, M. (1990) Spatial and temporal patterns in
15            summertime sulfate  aerosol acidity and  neutralization within a metropolitan area. Atmos. Environ. Part B
16            24: 115-126.
17
18      Waldman, J. M.; Liang, S.-K. C.; Lioy, P. J.;  Thurston, G. D.; Lippmann,  M. (1991) Measurements of sulfate
19            aerosol and its acidity in the SO2 source region of Chestnut Ridge, PA. Atmos. Environ. Part A
20            25: 1327-1333.
21
22      Waldman, J. M.; Koutrakis, P.;  Allen, G. A.;  Thurston, G. D.; Burton, R. M.; Wilson, W. E. (1995) Human
23            exposures to particle strong acidity, submitted.
24
25      Wall, S. M.; John, W.; Ondo, J. L. (1988) Measurement of aerosol size distributions for nitrate and major ionic
26            species. Atmos. Environ. 22: 1649-1656.
27
28      Wang, H.-C.; John, W. (1988) Characteristics of the Berner impactor for sampling inorganic ions. Aerosol Sci.
29            Technol. 8:  157-172.
30
31      Watson, J.  G.; Cooper, J. A.; Huntzicker, J. J. (1984) The effective variance weighting for least squares
32            calculations applied  to the mass balance receptor model.  Atmos. Environ. 18: 1347-1355.
33
34      Watson, J.  G.; Chow, J. C.; Richards, L. W.;  Neff, W. D.; Andersen, S. R.; Dietrich, D. L.;  Houck, J. E.;
35            Olnez, I. (1988) The 1987-88 metro Denver brown cloud study: v. I, II, III. Reno, NV: Desert Research
36            Institute; final report no. 8810.1F(l-3).
37
38      Watson, J.  G.; Chow, J. C.; Mathai, C. V. (1989) Receptor models in air resources management:  a summary of
39            the APCA international specialty conference. JAPCA 39: 419-426.
40
41      Watson, J.  G.; Chow, J. C.; Lu, Z.; Fujita, E. M.; Lowenthal,  D. H.; Lawson, D. R.; Ashbaugh, L. L. (1994)
42            Chemical mass  balance source apportionment of PM10 during the Southern California Air Quality Study.
43            Aerosol Sci. Technol. 21: 1-36.
44
45      Webber, J. S.; Dutkiewicz,  V. A.; Husain, L.  (1985) Identification of submicrometer coal fly ash  in a high-
46            sulfate episode at Whiteface Mountain,  New York. Atmos. Environ. 19: 285-292.
47
48      Weiss, R. E.; Larson, T. V.; Waggoner, A. P. (1982) In situ rapid-response measurement of HjSO^NH^jSC^
49            aerosols in rural Virginia. Environ. Sci. Technol. 16: 525-532.
50
51      Wesolowski, J. J.; John, W.; Devor, W.; Cahill, T. A.; Feeney, P. J.; Wolfe, G.; Flocchini, R. (1978) In:
52            Dzubay, T. G., ed.  X-ray fluorescence  analysis of environmental samples.  Ann Arbor,  MI: Ann Arbor
53            Science Publishers; pp.  121-130.
54


         April 1995                                    6-208       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Whitby, K. T.  (1977) Physical characterization of aerosols, methods and standards for environmental
  2            measurement. In: Kirchoff, W. H., ed. Washington, DC: National Bureau of Standards; pp. 165-173;
  3            NBS special publication no. 464.
  4
  5     Whitby, K. T.  (1978) The physical characteristics of sulfur aerosols. Atmos. Environ. 12:  135-159.
  6
  7     Whitby, K. T.; Sverdrup, G. M. (1980) California aerosols: their physical and chemical characteristics. In: Hidy,
  8            G. M.; Mueller, P. K.; Grosjean, D.; Appel, B. R.; Wesolowski, J. J., eds. The character and origins
  9            of smog aerosols: a digest of results from the California Aerosol Characterization Experiment (ACHEX).
 10            New York, NY: John Wiley & Sons, Inc.; pp. 477-517. (Advances in environmental science and
 11            technology: v.  9).
 12
 13     Whitby, K. T.; Husar, R. B.; Liu, B. Y.  H. (1972) The aerosol size distribution of Los Angeles smog. J.
 14            Colloid Interface Sci. 39:  177-204.
 15
 16     White, W. H. (1986) On the theoretical and empirical basis for apportioning extinction by aerosols: a critical
 17            review. Atmos. Environ. 20:  1659-1672.
 18
 19     White, W. H. (1990) The components of atmospheric light extinction: a survey of ground-level budgets. Atmos.
 20            Environ. Part A 24:  2673-2679.
 21
 22     White, W. H.; Husar,  R. B. (1976) A Lagrangian model of the Los Angeles smog aerosol.
 23
 24     White, W. H.; Macias, E. S. (1987) Paniculate nitrate measurements in rural areas of the western United States.
 25            Atmos. Environ. 21: 2563-2571.
 26
 27     White, W. H.; Macias, E. S. (1987) On measurements error and the empirical relationship of atmospheric
 28            extinction to aerosol  composition in the non-urban  west. In: Bhardwaja, P. J., ed. Visibility protection:
 29            research and policy aspects. Pittsburgh, PA: Air Pollution Control Association.
 30
 31     White, W. H.; Macias, E. S. (1989) Carbonaceous particles and regional haze in the western United States.
 32            Aerosol Sci. Technol. 10: 111-117.
 33
 34     White, W. H.; Macias, E. S.; Miller, D.  P.; Schorran, D. E.; Hoffer, T. E.; Rodgers, D. P. (1990)  Regional
 35            transport of the urban work week: methylchloroform cycles in the Nevada-Arizona desert.  Geophys. Res.
 36            Lett. 17: 1081-1084.
 37
 38     White, W. H.; Macias, E. S.; Nininger, R.  C.; Schorran,  D. (1994) Size-resolved measurements of light
 39            scattering by ambient particles in the southwestern U.S.A. Atmos. Environ.  28: 909-921.
 40
 41     Willeke, K.; Whitby, K. T. (1975) Atmosperic aerosols: size distribution interpretation. J.  Air Pollut. Control
 42            Assoc.  25: 529-534.
 43
44     Wilson, W. E.; Koutrakis, P.; Spengler, J. D. (1991) Diurnal variations of aerosol acidity, sulfate, and ammonia
45            in the atmosphere. Presented at: 84th annual meeting and exhibition; June; Vancouver, BC, Canada.
46            Pittsburgh, PA: Air and Waste Management Association; paper no. 91-89.9.
47
48     Wolff, G. T. (1984) On the nature of nitrate in coarse continental aerosols.  Atmos. Environ.  18: 977-981.
49
50     Wolff, G. T.; Korsog,  P. E. (1985) Estimates of the contributions of sources to inhalable paniculate
51            concentrations in Detroit. Atmos. Environ. 19: 1399-1409.
52
         April 1995                                    6-209       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Wolff, G. T.; Countess, R. J.; Groblicki, P. J.; Ferman, M. A.; Cadle, S. H.; Muhlbaier, J. L. (1981)
  2             Visibility-reducing species in the Denver "brown cloud"—II. sources and temporal patterns. Atmos.
  3             Environ. 15: 2485-2502.
  4
  5      Wolff, G. T.; Ferman,  M. A.; Kelly, N. A.; Stroup, D. P.; Ruthkosky, M. S. (1982) The relationships between
  6             the chemical composition of fine particles and visibility in the Detroit metropolitan area. J. Air Pollut.
  7             Control Assoc.  32: 1216-1220.
  8
  9      Wolff, G. T.; Korsog, P. E.;  Stroup, D. P.; Ruthkosky, M. S.; Morrissey, M. L. (1985) The influence of local
10             and regional sources on the concentration of inhalable paniculate matter in southeastern Michigan.
11             Atmos.  Environ. 19: 305-313.
12
13      Wolff, G. T.; Ruthkosky, M.  S.; Stroup, D. P.; Korsog, P. E. (1991) A characterization of the principal PM-10
14             species  in Claremont (summer) and Long Beach (fall) during SCAQS. Atmos. Environ. Part A
15             25: 2173-2186.
16
17      Zak, B. D.; Einfeld, W.; Church, H. W.; Gay, G.  T.; Jensen, A. L.; Trijonis, J.; Ivey, M. D.; Homann, P. S.;
18             Tipton,  C. (1984) The Albuquerque winter visibility study. Volume 1. Overview and data analysis.
19             Albuquerque, NM: U.S. Department of Energy, Sandia National Laboratories; Sandia report SAND84-
20             0173/1. Available from: NTIS, Springfield, VA; DE84014356.
21
22      Zemba, S.  G.; Golomb, D.; Fay, J. A. (1988)  Wet sulfate and nitrate deposition patterns in eastern North
23             America. Atmos. Environ. 22: 2751-2761.
24
25      Zhang,  X.; Turpin,  B. J.; McMurry, P. J.; Herring, S. V.; Stolzenburg, M. R. (1994) Mie theory evaluation of
26             species  contributions to the 1990 wintertime visibility reduction in the Grand Canyon. J. Air Waste
27             Manage. Assoc.  44: 153-162.
28
29      Zoller, W.  H.;  Gordon,  G. E. (1970) Instrumental neutron activation analysis of atmospheric pollutants utilizing
30             Ge(Li) 7ray detectors. Anal.  Chem. 42: 257-265.
         April 1995                                   6-210       DRAFT-DO NOT QUOTE OR CITE

-------
                 APPENDIX 6A:
   TABLES OF CHEMICAL COMPOSITION OF PM
April 1995                 6A-1   DRAFT-DO NOT QUOTE OR CITE

-------
                Table 1. Summary of PM2.5 Studies
EAST
Smoky Mtns
Shenandoah
Camden
Philadelphia
Deep Creek
Roanoke
Raleigh
Watertown
Hartford
Boston
Res.Tr. Pk.
Charlotte
















Ref
1
1
2
3
4
5
5
6,7
8
8
8
20
















Note WEST
Boise
Tarrant CA
b Five Points CA
Riverside CA
c San Jose
d Honolulu
d Winnemucca NV
Portland
a Seattle
a Southern California
a San Joaquin Valley
e Phoenix
Nevada















Ref
5
8
8
8
8
8
8
8
8
9,31
10
11
12















Note
d
a
a
a
a
a
a
a
a
g-n
i
j
f















Table 1 . Summary of Coarse Fraction
EAST
Smoky Mtns
Shenandoah
Camden
Philadelphia
Watertown
Hartford
Boston
Res.Tr. Pk.





Ref
1
1
2
3
6,7
8
8
8





Note WEST
o Tarrant CA
o Five Points CA
b Riverside CA
ab San Jose
o,p Honolulu
a,o Winnemucca NV
a,o Portland
a,o Seattle
Southern California
San Joaquin Valley
Phoenix


Ref
8
8
8
8
8
8
8
8
9,31
10
11


Note
a,o
a,o
a,o
a,o
a,o
a,o
a,o
a,o
g
i
j


CENTRAL
Albuquerque
St. Louis
Steubenville
Harriman
Portage
Topeka
Inglenook AL
Braidwood IL
Kansas City KS
Minneapolis
St. Louis
Kansas City MO
Akron
Cincinnati
Buffalo
Dallas
El Paso
Denver
Urban Denver
Non-urban Denver
Chicago
Houston
St. Louis
Harriman
St. Louis
Steubenville
Brownsville
Ontario
Studies
CENTRAL
St. Louis
Steubenville
Harriman
Portage
Topeka
Inglenook AL
Braidwood IL
Kansas City KS
Minneapolis
St. Louis
Kansas City MO
Cincinnati
Buffalo
Ref
5
6,7
6,7
6,7
6,7
6.7
8
8
8
8
8
8
8
8
8
8
8
13
14
14
15
16
17
17
18
21
24
37

Ref
6,7
6,7
6,7
6,7
6,7
8
8
8
8
8
8
8
8
Note
d





a
a
a
a
a
a
a
a
a
a
a

m
aa




k

n
I

Note
o-P
o,P
o.P
o,P
o,P
a,o
a,o
a,o
a,o
a,o
a,o
a,o
a,o
April 1995
6A-2
DRAFT-DO NOT QUOTE OR CITE

-------
               Table 1.  Summary of Coarse Fraction Studies (Cont.)
EAST











EAST
Smoky Mtns
Shenandoah
Camden
Philadelphia
Kingston
Watertown
Hartford
Boston
Res.Tr. Pk.


















Ref Note










Table
Ref Note
1 o,q
1 o,q
2 b
3 ab
6,7 p,q
6,7 p,q
8 a,q
8 a,q
8 a,q


















WEST










1. Summary of PM10
WEST
Tan-ant CA
Five Points CA
Riverside CA
San Jose CA
Honolulu HI
Winnemucca NV
Portland OR
Seattle
Southern California
San Joaquin Valley
Phoenix
San Fran. Bay
San Jose
Palm Springs
Pocatello, ID
Tuscon
Rillito, AZ










Ref










Note










CENTRAL
Dallas
El Paso
Denver
Chicago
Houston
St Louis
Harriman
St. Louis
Brownsville
Ontario
Ref
8
8
13
15
16
17
17
18
24
37
Note
a,o
a,o
o
s
o


k,r
n
I
Studies
Ref
8
8
8
8
8
8
8
8
9,31
10
11
29
29
38
39
40
42










Note
a.q
a,q
a,q
a,q
a,q
a,q
a,q
a,q
g,h
i
i
V
w
t

u











CENTRAL
St. Louis
Harriman
Steubenville
Portage
Topeka
Inglenook AL
Braidwood IL
Kansas City KS
Minneapolis
St. Louis
Kansas City MO
Akron
Cincinnati
Buffalo
Dallas
El Paso
Denver
Chicago
Houston
St. Louis
Harriman
St. Louis
Brownsville
Utah Valley
Ontario
SE Chicago, IL
Ohio
Ref
6,7
6,7
6,7
6,7
6,7
8
8
8
8
8
8
8
8
8
8
8
13
15
16
17
17
18
24
26
37
41
43
Note
p,q
p,q
p-q
p,q
p-q
a,q
a,q
a,q
a,q
a-q
a,q
a,q
a,q
a,q
a,q
a,q
q
s
q


X


I

y
April 1995
6A-3     DRAFT-DO NOT QUOTE OR CITE

-------
  Table 1 Footnotes

   a.  Inhalable Particle Network (IPN) Data. Only represents days of elevated
       concentrations, i.e., dichot filter loadings >50 ug/cm2.
   b.  Data from Site 28 only.
   c.  Average of all 6-hr samples.
   d.  Avg over all day/nite samples.
   e.  Average of all 12-hr samples at 2 incin. sites and 2 background sites.
       Only XRF values which exceeded associated uncertainties more than half the time
       at all four sites were included.
   f.  Average from Sparks site and Reno site.
   g.  Sampling only during intensive episodes.
   h.  Averages based on 12-hr day/nite samples. There were 59 sampling days at
       Claremont and 23 sampling days at Long Beach.
   i.  Avg over all sites: Stockton, Crow's Landing, Fresno, Kern, Fellows, and Bakersfield.
   j.  Average of Central Phoenix, West Phoenix, and Scottsdale sites.
   k.  Average of RAPS site 106.
   I.  Average from Walpole, Windsor 1,  and Windsor 2 sites.
   m.  Average of 3 urban sites: Auraria, Federal, and Welby.
   n.  Median VAPS values from Central site.
   o.  2.5-1 Sum.
   p.  Coarse concentrations may be 30% or more underestimated
       due to losses from handling filters.
   q.  PM15.
   r.  2.4-20 um.
   s.  No upper size cutoff on VAPS inlet.
   t.  Average of Palm Springs and Indio, CA.
   u.  Avg. of Downtown Tuscon, Orange Grove, Craycroft, and Corona de Tuscon sites.
   v.  Mean of annual avgs (1988-1992) from ~ 9 sites in Alameda, San Francisco,
       and Santa Clara counties.
   w.  24-hr average of day/nite concentrations at two sites in San Jose.
   x.  PM20. Average from RAPS site 106.
   y.  Avg. of Follansbee,  Mingo, Sewage Plant, Steubenville, and WTOV Tower sites.
   z.  Average of urban sites: Fresno, Bakersfield and Stockton.
  aa.  Average of non-urban sites: Brighton and Tower.
  ab.  Castor Avenue site only.
April 1995                             6A-4      DRAFT-DO NOT QUOTE OR CITE

-------

mparison
osition fo
Co
1)
co
                            *"*
                            8
                              §
                              EJ
               u
               8
                                              }
                              
        Pi n _ 7
        * 73 5
         Q   5 CO
/•^V /^N ^**\
—i tS CO
                            to
                            *— <
                            oo
                            4
                                              CO
                                              oo
                                              ON

                     in so oo


                     IT) OO OO
                  ^*- OO ^** "**-
                  tS ^ CS O

                  /-•v X*v /-N /-*\
                  « tS CO it
                                         c
                                         4
                                         00
           00
                 *
                 •i
                2 S
                fl J3
               00 00
             co

             .2
                                              i
                            fi-S
C4
*r*
•a
*s
1
I
                                            *
                                         "1 tS
                            ts
April 1995
                       6A-5      DRAFT-DO NOT QUOTE OR CITE

-------
                                                   -
                                                 H >
     o
50
|o
Ji
8
S
                         4)


            u
          t
       t-i  ,
                                                     — •  o
FP & IP(2.5-15). 24-hr
(midnite-midnite), every
other No Carbon
day.
24-hr
6th day
F+C(2.5-15)
sample every
ere included. No
       §
     OO .i
     »  i3

     Fl
                       3
                        \
                n   -^ «N - «
              »-< CU (^ ,_, U i ;*3
              X-S«« 5-vJ§ ^ x^ S»
              ^ H tS Pi OH CO 00
                               5'll'lS
                       3
                       •a
                       1
                       a
                            1
                            09
                                                 j§   a

     VO
                                     OO
                                                 Ov
April 1995
                       6A-6
                      DRAFT-DO NOT QUOTE OR CITE

-------
>
-La
 H
 6
 o

 •z
 o
 H
O
 CJ
n
h— <
a
10
11
12
13
14
San Joaquin Valley
6 sites
Aerosol
Composition
Phoenix PM Study
Phoenix
4 sites
Also comparison
aerosol data from
Denver, Reno, and
Sparks
Denver
Denver (SCENIC)
6/88-6/89
10/89 - 1/90
Sept'89 -
Jan'90
1/11-1/30/82
Nov'87-Jan'88
24-hr PM10 & PM2.5
every 6 days.
Mass, elements, ions
(K+,SO4=,NH4+,Na+),
EC, OC
F&C mass, elements,
uncertainties from 6 sites
6-hr samples, 2x/day,
(0600-1200, 1300-1900)
PM10 & PM2.5: mass,
elements, HNO3, SO2,
NH3, FP N03 and SO4,
ionic species, OC, EC.
Dichotomous sampler,
OC, EC, nitrate, sulfate
2x daily (0900-1600,
1600-0900). PM2.5 mass,
comp, sulfate, nitrate,
OC, EC, ionic species,
gases
1) Summary of annual geometric
avg, arith. avg, max 24-h PM10
and PM2.5 mass by site.
2) Ann. Avg Mass and comp. for
PM10 and PM2.5 by site.

1) temporal variation of PM2.5
mass at 4 sites.
2) Mean, SD, & Max: PM2.5,
EC, OC, NO3, SO4-, NH4+ and
elements for 3 Phoenix sites
3) Same for Denver (11/87-1/88)
4) Same for Reno (11/86-1/87)
5) Same for Sparks (11/86-1-87)
1) Measured PM2.5 and Coarse,
elements, OC, EC, nitrate,
day /night samples; light
extinction.
1) Avg, SD, Min, Max PM2.5
mass for 6 sites.
2) Avg, SD, Min, Max, for
PM2.5 mass, ionic species, EC,
OC, elements for 3 sites.
3) Source profiles
4) SCE for 4 sites by day and
night
PM10 highest hi
winter and
dominated by F
mass; C >50% of
PM10 in summer
and fall. Data show
spatial and temporal
variations of PM10
and PM2.5

Moudi size-resolved
(0-5.6 fim in 9 bins)
mass, NO3, SO4,
OC, EC.
Source
apportionment for
F&C particles and
extinction.
Source
Apportionment study

-------
ON

>

00
 O
 O
 z

 s
O
 C
 s
 w
O
I— I
H
tfl
15
16
17
18
19
20
Chicago
Houston
St. Louis &
Harriman
St. Louis
1) Albuquerque
2) Denver
Charlotte (2 incin
sites and 2 control
sites).
July, 1994
10-19
September
1980
Sept'85-
Aug'86
7/76-8/76
(St.Louis)
RAPS data for
St. Louis exist
for 5/75-3/77
but were not
in this article
1) Jan 3-4/83
2) Jan 19-20,
'82
4/30-6/4/92 &
9/21-9/28/92.
VAPS & Dichot. FM,
CM, OC, EC, elements,
SO2, MONO, HNO3.
Dichotomous sampler:
0.1-2.5, 2.5-15. 4 sites.
Consecutive 12h samples.
Daily F & C (2.5-10/mi).
Also SO2, NO2, and O3.
F(< 2.4) &C (2.4-20) 6-
12 hr. No Carbon.
F & C (2.5-10) +
Carbon, Nitrate & Sulfate
(1C) 12-h samples,
Day/Night:
0700-1900,1900-0700.
VAPS F&C + Acid
gases.
no carbon. 12-hr samples
1) Avg VAPS mass, SD, uncert.
for F&C, OC, EC.
1) Average F&C mass, elements,
Carbon, NH4, NO3, Sulfate
1) Mean, SD, range for PM10,
PM2.5, SO4, H+, SO2, NO2, O3
for both sites.
1) 2-mo avg of F-f-C mass,
metals, sulfate, for one site.
2) F+C composition of selected
samples (different sites) during
events.
3) CMB apportionment of F+C
fractions to 6 components (crustal
shale, crustal limestone,
ammonium sulfate, motor
vehicles, steel, paint).
4) Plots ofintercity variations in
source component concentrations
1) Mean daytime and nightime
comp. of F&C, EC, OC, nitrate,
sulfate, for each site.
2) Source app. of Denver winter
FP composition.
l)'Mean ambient FP cone. +-
XRF unc. @ 4 sites
2) CMB results for FP.

Source
apportionment.

1) Crude CMB
source apportionment
of FP with 6
sources.
More complete
source app results in
Lewis & Enfield
paper.


-------







$
1) avg F mass 4- comp.
2) avg source contributions by
SRFA
3) SRFA-derived source profil
|
u
o
^
o"
uT
S
oo
0
Q
S3
p— .


«
J3
>
 ^
^ 53 .^3
"g w? «J
^ ?i ^5
1 1 g
i-H
1
We
CO
3
O
S
ex,
CO
W CO
u .a




0
OH

O\
i

o
o.
<4-l
o
E $
••H *^^
> "2
£ ^
PC CQ
rs
cs






&
Q
6
1
+ 60 °
C^U rT
.
00S'gH
«N 1
A w
•*—S W
O *
is
o?oo
^ o
^-^ oo
UH -M
OO
r-
r~-"
i— i





X
1
o
fi
CO
CN
•^
O
8
1
> .
M G
^j
o "8
Z 6

S
£
«c
^
i?l



lH
FPMES
door/outdoc
'~> C
*"^ .1^
1/3
+
.I* OS
W< •


1 1
0> ct
H3 "5
> •-
11
ll







•k
C ^
8|
u -5
+ =3
8S|
rT 6 8

•& 'S
•a "a
H tl
!§
§1
> b
rTco'






-*— »
OS
•a
1
u
rs

1
*^
s
«J
Ul
® .
0 ja'S
1) Seasonal avg SCE for PMl
3 sites, (geological, motor ve
construction, vegetative, sulfa1
nitrate, OC, EC)
I.S ff
9> O..S
« wl S
^ Ul
£j m "O
2 "^ o
PH JC **
^VO^
"r-
00
oo


1
(V ^
t"H ^^ r3^
|
It
SO
(S

SI
§5 ffi
.s S o
c3 55 ^i
pSll
.^
^
1
•"a
-o
ea
II
o c^
O2 U4
•gSri
|o*
"3 ^ T3
tfc rt S
fefl
1 * o
OO >^ j-

OS C ^ co
-H 0 Q 3
C3
3
O
U
c3
^™3
u
2
C/3
r-
ts
April 1995
6A-9     DRAFT-DO NOT QUOTE OR CITE

-------
us
45%
mb
co
tes
1. Highest PMIO
mass during Nov,
Dec, Jan.
2. Wood
contribu
PM
of
                                      _
                                   •a .8
                                    Sw
                                   .2 a
                                   •° «> oi
                                    *.S 3

                                   U&2
                                   •B ^ S
                                   Q-SE
         5
S
      00
 - 8   2
Is   |g
                         G
              S
            8 a
            II

      ~H eu ja
                   ^h »^H '"^ . » gj   . **iH
                   °rtob^g8o».

                   *7? **^ S *" "^ "M 3 W
  3 0 ^
  55«ts
      A.
         >>
                   0)
            .
            G
            — <
             ^  r-
            I o
            i*§
            |u
            «>
         re
          * 1
                IT) ft-
  il
•    M »   *
                   •
ass, elements, ions,
lfate, nitrate, Carbc
                                           .a   s s
      £
      oo
      oo
      vo
                                      (O
                                         o
                 fa .52 -o .S2 oo
                 3 CL, C P-ON
4)
'a
>
a
I
11
a «
b? \O
            el
            § 8:
               s
                                    B  S3
                                    •u ^ O
                        $
PQ §
  '
•< o
§ I.
Si
      oo
      n
ON
«S
April 1995
             6A-10     DRAFT-DO NOT QUOTE OR CITE

-------
 VO
 Ui
               31
 1) Claremont
 (SCAQS)

 2) Long Beach
 (SCAQS)
 1) Summer '87
 (59d)
 2) Fall '87
 (23d)
 Continuous 12-hr
 PMIO and PM2.5.

 Mass, elements, ionic
 species, EC, OC
 1) Mean, SD, & Max: PMIO,
 FPM, CPM, EC, OC, NO3,
 SO4=, NH4+.
 2) Mean values of above species
 during intensive and non-intensive
 periods.
 3) Day/nite values of above
 4) PMIO and PM2.5 mass
 balances
 5) Summary of EC, OC data.
               32
CADMP ~ 8 sites:
Gasquet, Fremont,
Bakersfield,
Yosemite,  Sequoia,
Long Beach, Los
Angeles, Azusa
Summer '88
2 samples every 6th day.
0600-01800,1800-0600.
PM2.5, PMIO.  Mass,
ionic species,
                                                                                 1) Graph of avg PMIO & PM2.5
                                                                                 mass and ratio @ 8 sites
                                                                                 2)Graphs of PMIO & PM2.5 ionic
                                                                                 concentrations.
                                Ask Chow/Watson
                                for raw data.
•n
H
6
o
s
o
H-
w
o
90
n
t— i
3
               33
Central California -
53 sites in SF Bay
area, Sacramento
Valley, San Joaquin
Valley, North and
South Central
Coast, Mountain
Counties
1) 1989
2) July &
August, 1988
PMIO every 6th day.
Sulfate and nitrate
measured on a subset of
these samples.
1) 1989 Max and Avg PMIO
mass, Sulfate, and Nitrate for
-53 sites.
2) Summertime '88 Avg, SD, and
Max PMIO and PM2.5 Mass,
comp, OC,EC, Ionic species, for
3 SJVAQS sites. [Annual data
summary is in ref 20].
               34
Birmingham
1986-1989
Daily 24-hr PMIO mass.
Also Ozone data.

No composition data.
1) Table of percentile points of the
distribution of PMIO, 03, T,
DewPoint, Pneumonia, Chronic
obstructive pulmonary disease.
2) Avg PMIO and O3 by season
                                                                        Aside:
                                                                        Indoor/Outdoor
                                                                        ratios of 0.63 for
                                                                        PMIO were reported
                                                                        in Tuscon.

-------
35
36
37
38-43
Philadelphia
State College, PA
Southern Ontario
3 sites
Miscellaneous sites
14 sites
1973-1980
summer 1990
Jan-Nov, 1991
1984-1990
24-hr (midnite-midnite)
TSP.
No composition data.
Indoor, outdoor, personal
SO4=,H+, andNH3
24-hr, midnite-midnite,
every 6th day. PMIO
dichot sampler.
PMIO concentrations.
1) Table of percentile points of the
distribution of TSP, SO2, T,
DewPoint, Mortality.

l)Avg mass, elements, for F&C
fractions, for 3 sites. NO OC,
EC.
1) Measured PMIO mass and avg
source contributions (up to 10
source categories).

Validation of
personal exposure
models

Primary reference is
Ref 10.
0\
 T1
 H

 6
 o
 z
 o
 H
o
H
w
o
h-H
H
tn

-------
List of References

1.     Stevens, R.K., Dzubay, T.G., Lewis, C.W., and Shaw, R.W. (1984).  Source apportionment
       methods applied to the determination of the origin of ambient aerosols that affect visibility in
       forested areas. Atmos. Environ. 18 261.

2.     Dzubay, T.G. and Stevens, R.K.  (1988).  A composite receptor model applied to Philadelphia
       aerosol. Environ. Sci. Techn. 22 46.

3.     Unpublished data from J. Pinto, U.S. EPA, Research Triangle Park, NC (1995).

4.     Vossler,  T.L., Lewis,  C.W.,   Stevens, R.K.,  Dzubay,  T.G.,  Gordon, G.E.,  Tuncel,
       S.G.,Russworm, G.M., and Keeler,  G.J. (1989). Composition and origin of summertime air
       pollutants at Deep Creek Lake, Maryland. Atmos. Environ. 23_ 1535.

5.     Stevens, R.K., Hoffman, A.J., Baugh, J.D., Lewis, C.W.,  Zweidinger, R.B., Cupitt, L.T.,
       Kellogg, R.B. and Simonson, J.H.  (1995).   A comparison of air quality measurements in
       Roanoke, Va, and other integrated air cancer project monitoring locations. In: Measurement of
       Toxic and Related Air Pollutants. Proceedings of the 1993 U.S. EPA/A&WMA International
       Symposium. A&WMA, Pittsburgh, PA, p!85.

6.     Spengler, J.D. and Thurston, G.D. (1983). Mass and elemental composition of fine and coarse
       particles in six US cities. JAPCA  33 1162.

7.     Dockery, D.W., Pope, C.A., Xu,  X., Spengler, J.D., Ware, J.H., Fay, M.E., Ferris, B.G. and
       Speizer, F.E. (1993). An association between air pollution and mortality in six U.S. cities. New
       Eng. J. Medicine 329 1753.

8.     Davis, B.L. Johnson, L.R., Stevens, R.K., Courtney, W.J. and Safriet, D.W. (1984). The quartz
       content and elemental composition of aerosols from selected sites of the EPA inhalable particulate
       network. Atmos. Environ. 18 771.

9.     Chow, J.C., Watson, J.G., Fujita, E.M., Lu, Z., Lawson, D.R. and Ashbaugh, L.L. (1994).
       Temporal  and spatial variations of PM2.5 and PM10 aerosol in the Southern California Air
       Quality Study. Aerosol Sci. and Tech. 21  2061.

10.    Chow, J.C., Watson, J.G., Lowenthal, D.H., Solomon, P.A., Magliano, K.L. Ziman, S.D. and
       Richards,  L.W. (1993). PM10 and PM2.5 compositions hi  California's San Joaquin Valley.
       Aerosol Sci & Tech. 18 105.

11.    Unpublished data from the Desert Research Institute (1995).

12.    Chow, J.C., Watson, J.G., Pritchett, L., Lowenthal, D.H., Frazier, C., Neuroth, G. and Evans
       K. (1990). Wintertime visibility in Phoenix, Arizona. Paper 90-66.6 in Proceedings of the 83rd
       National Meeting  of the Air  & Waste Management Association, Pittsburgh, PA,  24-29 June,
       1990.

13.    Lewis, C.W., Baumgardner, R.E., Stevens, R.K.,  and Russworm,  G.M.  (1986).  Receptor
       modeling study of Denver whiter haze. Environ. Sci. Techn. 20 1126.
  April 1995                                 6A.13      DRAFT-DO NOT QUOTE OR CITE

-------
 14.     Watson, J.G., Chow, J.C., Richards, L.W., Neff, W.D., Andersen, S.R., Dietrich, D.L. Houck,
        J.E.  and  Olmez,  I. (1988).  The  1987-88  metro Denver brown cloud study, vol 3: data
        interpretation. Desert Research Institute. Document No. 8810  1F3.

 15.     Unpublished data (1995).

 16.     Johnson, D.L.,  Davis,  B.L.,  Dzubay, T.G., Hasan, H., Crutcher,  E.R.,  Courtney,  W.J.,
        Jaklevic, J.M.,  Thompson, A.C., and Hopke, P.K. (1984). Chemical and physical analyses of
        Houston aerosol for interlaboratory comparison of source apportionment procedures. Atmos.
        Environ. 18, 1539.

 17.     Dockery,  D.W.,  Schwartz, J.  and  Spengler, J.D. (1992). Air pollution and daily mortality:
        associations with particulates and acid aerosols. Environ. Research 5.9 362.

 18.     Dzubay, T.G. (1980). Chemical elements balance method applied to dichotomous sampler data.
        In: Annals of the New York Academy of Sciences  338 126.

 19.     Stevens, R.K. (1985). Sampling and analysis methods for use in source apportionment studies
        to determine impact of wood burning on fine particle mass. Environment International H 271.

 20.     Mukerjee, S., Stevens, R.K., Vescio, N., Lumpkin, T.A., Fox, D.L., Shy, C. andKellogg, R.B.
        (1993). A methodology to apportion ambient air measurements to investigate potential effects on
        air quality near waste  incinerators. In: Proceedings of the 1993 Incineration Conference,
        Knoxville, TN.  527.

 21.     Koutrakis, P. and Spengler,  J.D.  (1987). Source  apportionment of  ambient particles  in
        Steubenville, OH using specific rotation factor analysis. Atmos. Environ. 2_1 1511.

 22.     Chow,  J.C.,  Watson,  J.G.,  Ono, D.M. and Mathai, C.V.  (1993). PM10 standards and
        nontraditional paniculate source controls: a summary of the A&WMA/EPA international specialty
        conference. Air & Waste 43 74.

 23.     Solomon,  P.A.  and Moyers, J.L. (1986). A chemical characterization of wintertime haze hi
        Phoenix, Arizona.  Atmos. Environ.  20 207.

 24.     Ellenson,  W.D.,  Schwab, M., Egler, K.A., Shadwick, D.  and Willis,  R.D.  (1994).  Draft
        technical report for the pilot project of Lower Rio Grande Valley environmental study. ManTech
        Environmental Technology, Inc. Submitted to the U.S. EPA.

 25.     Chow, J.C., Watson, J.G., Frazier, C.A., Egami, R.T., Goodrich, A. and Ralph, C. (1988).
        Spatial and temporal source contributions to PM10 and PM2.5 hi Reno, NV. In: Transactions:
        PM10 Implementation of Standards, an APCA/EPA  international specialty conference.  Air
        Pollution Control Association, 439.

 26.     Pope, C.A., Schwartz, J. and Ransom, M.R. (1992). Daily mortality and PM10 Pollution in Utah
        Valley. Archives of Environ. Health 47 211.

 27.     Fairley, D. (1990). The  relationship of daily mortality to suspended particulates hi Santa Clara
        county 1980-1986. Environ. Health  Perspectives £2 159.
April 1995                                 6A-14       DRAFT-DO NOT QUOTE OR CITE

-------
28.    Chow, J.C., Watson, J.G., Lowenthal, D.H., Solomon, P.A., Magliano, K.L. Ziman, S.D. and
       Richards, L.W. (1992). PM10 source apportionment in California's San Joaqin Valley. Atmos.
       Environ. 26A 3335.

29.    Chow, J.C., Fairley, D., Watson, J.G., De Mandel, R., Fujita, E., Lowenthal, D.H., Lu, Z.,
       Frazier, C.A., Long, G. and Cordova, J. (1994). Source apportionment of wintertime PM10 at
       San Jose, CA. J. Environ Engineers, in press.

30.    Watson, J.G., Chow, J.C., Lu, Z., Fujita,  E.M.,  Lowenthal, D.H., Lawson, D.R.  and
       Ashbaugh, L.L. (1994).  Chemical mass balance source apportionment of PM10 during the
       Southern California Air Quality Study. Atmos.  Environ. 12 2061.

31.    Wolff, G.T., Ruthkosky, M.S., Stroup, D. adn Korsog, P.E. (1991). A characterization of the
       principal PM10 species in Claremont (summer) and Long Beach (fall) during SCAQS. Atmos.
       Environ. 25A 2173.

32.    Ashbaugh, L.L., Watson, J.G. and Chow, J. (1989). Estimating fluxes from California's dry
       deposition monitoring data. Paper 89-65.3 in: Proceedings of the 82nd Annual Meeting of the
       Air & Waste Management Association, Anaheim, CA.

33.    Chow, J.C., Watson, J.G.,  Solomon,  P.A., Thuillier, R.H., Magliano, K.L. Ziman, S.D.,
       Blumenthal, D.L. and Richards, L.W. (1994). Planning for SJVAQS/AUSPEXParticulate matter
       and visibility  sampling and  analysis. In: Planning and Managing Regional Air Quality, ed. by
       Paul Solomon. CRC Press, Inc.

34.    Schwartz, J.  (1994).  Air pollution  and hospital admissions for the elderly  in Birmingham,
       Alabama. Am. J. Epidemiology.  139 589.

35.    Schwartz, J. and Dockery, D.W. (1992). Increased mortality in Philadelphia associated with daily
       air pollution concentrations. Am. Rev. Respir.  Dis.  145 600.

36.    Suh, H.H., Koutrakis, P. and Spengler, J.D. (1993). Validation of personal exposure models for
       sulfate and aerosol strong acidity. J.  Air Waste Manage. Assoc. 43 845.

37.    Conner, T.L., Miller, J.M.,  Willis, R.D,, Kellogg,  R.B.  and Dann, T.F.  (1993).  Source
       apportionment of fine and coarse particles in Southern Ontario, Canada. In: Proceedings of the
       86th Annual Meeting of the Air & Waste Management Association, Denver, CO. Paper 93-TP-
       58.05.

38.    Kim, B.M., Zeldin, M. and Liu, C. (1992). Source apportionment study for state implementation
       plan development in the  Coachella Valley.  In: PM10 Standards and nontraditional paniculate
       source controls.  Chow and Ono, Eds., Air & Waste Management Association, Pittsburgh, PA,
       979.

39.    Houck, J.E.,  Rau, J.A.,  Body, S. and Chow,  J.C.  (1992). Source apportionment - Pocatello,
       Idaho PM10 nonattainment area.  Ibid, 219.

40.    Chow, J.C., Watson, J.G., Lowenthal, D.H., Frazier, C.A., Hinsvark, B.A., Pritchett, L.C. and
       Neuroth,  G.R.  (1992).  Wintertime  PM10  and PM2.5  chemical  compositions  and  source
       contributions in Tuscon, Arizona. Ibid, 231.
 April 1995                                6A-15      DRAFT-DO NOT QUOTE OR CITE

-------
41.    Vermette, S.J., Williams, A.L. and Landsberger, S. (1992). PMIO source apportionment using
       local surface dust profiles: examples from Chicago. Ibid, 262.

42.    Thanukos, L.C., Miller, T., Mathai, C.V., Reinholt, D. and Bennett, J. (1992). Intercomparison
       of PMIO samplers and source apportionment of ambient PMIO concentrations in Rillito, Arizona.
       Ibid, 244.

43.    Skidmore, L.W., Chow, J.C. and Tucker, T.T. (1992). PMIO air quality assessment for the
       Jefferson County, Ohio air quality control region. Ibid, 1016.
     April 1995                                6A-16      DRAFT-DO NOT QUOTE OR CITE

-------
               PM2.5 COMPOSITION  (24-h AVG)
                                             EASTERN U.S.
Units = ug/m3
>
 I
I—*
^1
 O
 o
 2
 o
 H
O
 a
 o
o
H- (
H
m
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
1
Smoky Mtns
9/20-26/78
12
12
24.00
2.22
1.10
0.30
12.00

<0.054
<0.003

0.018
0.016

<0.010

0.003
0.028
0.040






0.097

3.744

0.001
0.038


<0.006
<0.004
0.009
1
Shenandoah
7/23-5/08/80
12
28
27.00
0.44
1.12

13.60

<0.105
<0.003

0.008
0.035

0.010

0.005
0.054
0.061






0.052

4.539

0.001
0.116


<0.010
<0.010
0.011
2(b)
Camden
7/14-8/13 '82
6am-6pm-6am
12
50
28.70
2.05
1.87
<0.48
11.20

0.053
0.001

0.029
0.040
0.002
0.003
0.002

0.091
0.101

0.006
0.001
0.146
0.011

0.249

4.200
0.079
0.002
0.103
<0.012
< 0.002
<0.027
0.013
0.082
3
Philadelphia
7/25-8/14/94
24
21
32.18
4.51
0.76



0.114


0.009
0.058

0.026

0.007
0.127
0.060
0.023
0.003

0.070
0.007
0.015
0.019

3.251

<0.002
0.165


<0.042
<0.013
0.041
4(c)
Deep Creek
8/83
4x daily
6
98
40.00
1.45
0.18
0.57



0.001

0.005
0.048




0.058
0.044

0.003

0.034


0.048

6.700
0.001
0.003
0.150



0.001
0.013
5(d)
Raleigh
1/85-3/85
7am-7pm-7am
12
30.30
10.00
0.50



0.009
0.001

0.028
0.018

0.007
0.000
0.020
0.044
0.159

0.003
0.001



0.096

1.729

0.002
0.076



0.003
0.015
5(d)
Roanoke
10/88-2/89
7am-7pm-7am
12
19.90
7.30
1.50



0.176
0.002

0.005
0.047

0.053
0.001
0.007
0.114
0.177

0.012
0.001



0.027

1.177

0.002
0.077



0.004
0.083
6,7
Watertown
5/79-6/81
0000-0000
24
354
14.90



6.50
20.3



0.088
0.041

0.084


0.074


0.004


0.009

0.329

1.800

0.001
0.100



0.022

8(a)
Hartford
1980
24
2
26.75





0.035


0.036
0.070


0.003
0.043
0.125
0.171

0.007


0.010

0.510

2.219

0.001
0.177


0.002
0.017
0.079
8(a)
Boston
1980
24
1
34.80






0.002

0.020
0.070


0.004
0.035
0.121
0.096

0.001


0.012
0.009
0.285

3.869

0.001
0.144



0.020
0.046
  References are listed in Table 1 Appendix. Associated notes are explained in Table 1.

* Values for this size fraction are calculated from the average measured values reported for the other two size fractions.

-------
 oo





 o



 T1
 o
 H

O
 C
 o
 H
 W
n
HH
H
W
PM2.5 COMPOSITION (24-h
Ret
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
8(a)
Res.Tr.Pk
1980
24
3
28.77





0.073
0.002

0.007
0.035



0.016
0.120
0.148

0.003


0.001
0.042
0.106

2.835

0.002
0.350




0.018
9(9)
Los Angeles
Summer'87
4,5 and 7
11 days
41.10
8.27
2.37
4.34
9.41

0.035
0.022
0.015
0.013
0.022

0.093
0.022
0.063
0.099
0.041
0.024
0.016

0.202
0.005
0.060
0.038

2.832

0.013
0.052

0.019
0.005
0.006
0.090
AVG)
9(9)
Los Angeles
Fall'87
4 and 6
6 days
90.20
18.46
7.28
22.64
4.38

0.250
0.015
0.043
0.065
0.335

0.453
0.025
0.273
0.557
0.217
0.075
0.043

0.466
0.007
0.046
0.185

1.998

0.011
0.520

0.028
0.060
0.007
0.298
WESTERN U.S.
10(i)
San Joaquin Valley
6'88-6'89
24
~35
29.89
4.87
3.24
8.17
3.00

0.152

0.012
0.010
0.096
<0.007
0.094
0.003
0.096
0.180
0.188

0.006


0.016
0.007
0.029
0.001
1.242
<0.002
0.001
0.460
<0.015
0.002
0.017
0.015
0.078

11(j)
Phoenix
10/13/89-1/17/90
6 h, 2x/day
~ 100 days
29.37
10.10
7.47
3.60
1.33

0.130
< 0.020
<0.106
0.011
0.170
<0.018
0.365
0.003
0.015
0.216
0.207

0.023
<0.006

0.003
<0.051
0.039
<0.0025
0.437
<0.033
< 0.002
0.430
< 0.028

<0.030
<0.016
0.056

5(d)
Boise
12/86-3/87
7am-7pm-7a
12
35.70
12.70
1.70



0.102
0.002

0.014
0.026

0.122
0.001
0.011
0.022
0.145

0.002
0.002



0.045

0.603

0.001
0.069



0.001
0.019
Units =
12(f)
Nevada
ug/m3
8(a)
Tarrant CA
11/86-1/87 1980
00-2400
24 24
24 6
56.92
19.97
15.17
2.43
1.67

0.275
0.001
0.013
0.033
0.215

0.145
0.002
0.010
0.310
0.280

0.015


0.006
0.041
0.115
0.001
0.765

0.000
0.860

0.004
0.043
0.009
0.033
57.05





0.177


0.102
0.455


0.002
0.047
0.316
0.186

0.032


0.003

0.619

2.578


0.583


0.010

0.095

8(a)
Five Points C
1980
24
3
31.80





0.239


0.015
0.150

0.004
0.001
0.024
0.216
0.244

0.005


0.025
0.007
0.087

1.129

0.001
0.656


0.005
0.006
0.016

8(a)
Riverside CA
1980
24
4
35.18





0.036


0.037
0.301

0.009

0.040
0.127
0.120

0.007


0.007

0.376

1.653

0.001
0.234



0.003
0.029
References are listed in Table 1 Appendix. Associated notes are explained in Table 1.

Values for this size fraction are calculated from the average measured values reported for the other two size fractions.

-------
           PM2.5 COMPOSITION (24-h AVG)
                                                            CENTRAL U.S.
Units = ug/m3
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
8(a)
San Jose CA
1980
24
6
36.28





0.123
0.001

0.188
0.089

0.050
0.003
0.043
0.148
0.248

0.006


0.006
0.013
0.891

0.852


0.292



0.002
0.061
8(a)
Honolulu
1980
24
1
21.10





1.127


0.017
1.024

0.518
0.004
0.018
0.726
0.371

0.020


0.002
0.002
0.071

0.313


2.363


0.063
0.001
0.011
8(a)
Winnemucca
1980
24
5
9.68





0,361


0.006
0.243



0.026
0.231
0.149

0.003


0.001

0.042

0.358


0.914


0.009

0.011
8(a)
Portland
1980
24
4
37.18





0.581
0.012

0.093
0.154

0.021
0.009
0.072
0.270
0.218

0.052


0.027
0.017
0.422

1.944

0.001
0.377


0.005
0.014
0.081
8(a)
Seattle
1980
24
1 .
10.70





0.002
0.006

0.019
0.037


0.002
0.024
0.098
0.080

0.004


0.006
0.006
0.215

0.831

0.001
0.092




0.059
5(d)
Albuquerque
12/84-3/85
7am-7pm-7am
12
20.60
13.20
2.10



0.077


0.085
0.059

0.036


0.045
0.074


0.000



0.237

0.507


0.076




0.007
13
Denver
1/11-30/82
6am-6pm-6a
12
~26
20.73
7.11
2.15
2.22
2.06

0.394
<0.002
0.031
0.103
0.047
0.006
0.052
<0.009
0.010
0.079
0.079

0.011


0.003
0.043
0.326
<0.003
0.709


0.277

<0.003
<0.027

0.046
14(m)
Urban Denver
11/87-1/88
9am-4pm-9am
7&17
~136
19.67
7.25
4.41
3.96
1.55

0.037


0.018
0.058
0.005
0.141
0.003
0.017
0.111
0.077

0.012


0.002

0.075

0.642
0.004
0.001
0.272
0.006
0.001
0.009

0.031
14(aa) 15
Non-urban Denver Chicago
11/87-1/88 7/94
9am-4pm-9am 0800-0800
7&17 24
~150 16
10.35 13.57
5.39
1.31



0.046
< 0.003
< 0.091
0.004
0.045
<0.029
0.011
<0.005
0.011
0.089
0.061
0.012
0.005
<0.002
0.022
<0.001
0.008
0.027

1.321
<0.042
<0.001
0.074
<0.049

<0.029
<0.009
0.052
D
O
O

2
O
H

O
G
O
H
W

O
*J

n
H-i
H
W
  References are listed in Table 1 Appendix.  Associated notes are explained in Table 1.

* Values for this size fraction are calculated from the average measured values reported for the other two size fractions.

-------
                  PM2.5 COMPOSITION (24-h AVG)
                                                           CENTRAL U.S.
                                                  Units = ug/m3
 vo
 vo
ON
to
o
 "fl
 H

 6
 O

 z
 o
 H

O
 d
 o
 H
 m
o
h-H
H
W
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
8
Sb
Se
Si
Sn
Sr
Ti
V
Zn
16
Houston
9/10-19/80
12
20
38.60
5.68
1.42
0.59
14.61

0.123
<0.005
0.048
0.055
0.155
<0.003
0.032
<0.005
0.028
0.162
0.119

0.014

<0.38
0.004
0.028
0.465
<0.002
4.834
0.006
<0.002
0.210
<0.005
<0.002
<0.014
<0.008
0.084
6,7
Harriman
5/80-5/81
0000-0000
24
256
20.80



8.10
36.1



0.038
0.150

0.021


0.120


0.017


BQL

0.180

2.500

0.002
0.120



BQL

17 6,7
Harriman Kingston
9/85-8/86 5/80-6/81
0000-0000
24 24
330 169
21 .00 24.60



8.70
36.1



0.044
0.120

BQL


0.097


0.010


BQL

0.194

2.400

0.002
0.200



BQL

6,7
Portage
3/79-5/81
0000-0000
24
271
11.00



6.81
10.5



0.011
0.045

0.027


0.049


0.003


BQL

0.061

1.400

0.001
0.075



BQL

6,7
Topeka
8/79-5/81
0000-0000
24
286
12.50



6.05
11.6



0.045
0.250

0.031


0.090


0.004


BQL

0.163

1.100

0.000
0.190



BQL

8(a)
El Paso
1980
24
10
27.16





0.155
0.025

0.070
0.332


0.001
0.036
0.134
0.127

0.004


0.001

0.481

0.823

0.002
0.436


0.003

0.055
8(a)
Inglenook
1980
24
8
32.03





0.082
0.001

0.040
0.326

0.003
0.002
0.032
0.281
0.408

0.037


0.001
0.008
0.309

2.655

0.001
0.685




0.133
8(a)
Braidwood
1980
24
1
28.20





0.089


0.003
0.084



0.024
0.071
0.052

0.001


0.001

0.041

2.060

0.001
0.220




0.011
8(a)
Kansas City KS
1980
24
8
25.66





0.091
0.003

0.027
0.519


0.004
0.032
0.189
0.311

0.006


0.002
0.013
0.180

1.816

0.001
0.434


0.004

0.034
  References are listed in Table 1 Appendix,

* Values for this size fraction are calculated
 Associated notes are explained in Table 1.

from the average measured values reported for the other two size fractions.

-------
               PM2.5 COMPOSITION  (24-h AVG)
CENTRAL U.S.
Units = ug/m3
>
o
o

2
o
H

O
G
O
H
W

O
»

O
HH

m
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
8(a)
Minneapolis
1980
24
6
15.50






0.004

0.047
0.103


0.001
0.035
0.087
0.092

0.005


0.001

0.308

0.907

0.001
0.169




0.045
8(a)
Kansas City MO
1980
24
3
16.77






0.007

0.064
0.213


0.002
0.021
0.140
0.142

0.006


0.001

0.369

0.763


0.177




0.046
8(a)
Akron
1980
24
7
36.09





0.046
0.012

0.039
0.110


0.010
0.037
0.609
0.268

0.085


0.006
0.059
0.412

3.419

0.008
0.522


0.009

0.150
8(a)
Cincinnati
1980
24
2
29.80





0.062
0.013

0.024
0.062


0.003
0.024
0.174
0.136

0.011


0.004
0.043
0.343

2.876

0.005
0.328


0.003

0.053
8(a)
Buffalo
1980
24
14
38.75





0.192
0.009

0.003
0.218


0.002
0.026
0.671
0.310

0.033


0.008
0.060
0.359

3.706

0.005
0.241



0.001
0.078
8(a)
Dallas
1980
24
4
28.93





0.111
0.033

0.223
0.691


0.005
0.043
0.248
0.125

0.015


0.002
0.018
1.066

1.514


0.442


0.007
0.002
0.054
8(a)
St. Louis
1980
24
5
23.06





0.119
0.003

0.025
0.090



0.018
0.076
0.126

0.002


0.002
0.020
0.277

2.333

0.002
0.170




0.023
18(k)
St. Louis
8-9/76
6-12
34.00





0.203
0.002
0.020
0.132
0.132
0.004
0.087
0.006
0.029
0.275
0.261

0.036


0.004
0.001
0.688
0.000
4.655
0.006
0.004
0.458
0.009
0.002
0.112
0.002
0.101
6,7 17
St. Louis St. Louis
9/79-6/81 9/85-8/86
0000-0000
24 24
306 311
19.00 17.70



8.10 8.00
10.3 9.7



0.078
0.101

0.052


0.190


0.021


0.003

0.327

2.100

0.002
0.160



BQL

6,7
Steubenville
4/79-4/81
0000-0000
24
499
29.60



12.80
25.2



0.042
0.097

0.092


0.590


0.029


0.005

0.216

4.700

0.005
0.290



0.011

                References are listed in Table 1 Appendix. Associated notes are explained in Table 1.

              * Values for this size fraction are calculated from the average measured values reported for the other two size fractions.

-------
               PM10 COMPOSITION  (24-hr AVG)
EASTERN U.S.
>
to
to
O
O

Z
O
H

O
c:
o
H
w

o
&

o
t—t
H
Units = ug/m3
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
1(o,q)*
Smoky Mtns
9/20-26/78
12
12
29.60
2.22
1.10
0.30
12.00

BQL
BQL

0.023
0.338

BQL

0.003
0.146
0.148




BQL

0.111

3.744

0.001
0.618


0.018
BQL
0.009
1(o,q)*
Shenandoah
7/23-5/08/80
12
28
34.40
0.44
1.12

14.38

0.311


0.011
0.339

0.189

0.011
0.212
0.190

BQL


BQL

0.061

4.539

0.001
0.929


0.017
BQL
0.017
2(b)*
Camden
7/14-8/13 '82
6am-6pm-6am
12
50
40.10
2.05
2.29
0.57
11.20

0.603
0.001

0.044
0.400
0.002
0.072
0.002

0.581
0.252

0.017
0.001
0.146
0.015

0.303

4.430
0.260
0.002
1.713
BQL
0.002
0.065
0.020
0.112
3(ab)* 4(c) 5(d)
Philadelphia Deep Creek Raleigh
7/25-8/14/94 8/83 1/85-3/85
4x daily 7am-7pm-7a
24 6 12
21 98
40.60
4.51
0.76



0.439


0.012
0.479

0.073

0.021
0.479
0.160
0.126
0.010

0.206
0.009
0.042
0.032

3.251


1.098


0.030

0.092
5(d) 6,7(p,q)
Roanoke Watertown
10/88-2/89 5/79-6/81
7am-7pm-7a 0000-0000
12 24
354
24.20



8.94




0.110
0.250

0.389


0.350


0.009


0.011

0.405

2.000

0.001
1.100



0.022

8(a.q)
Hartford
1980
24
2
54.60





1.910


0.082
0.934

0.302
0.011
0.069
1.195
0.481

0.028


0.015
0.033
0.681

2.647

0.001
4.694


0.096
0.025
0.133
8(a,q)
Boston
1980
24
1
140.40





3.458
0.003

0.045
1.139

0.301
0.008
0.058
1.733
0.629

0.030


0.034
0.025
0.462

4.371

0.001
6.904


0.154
0.028
0.100
               References are listed in Table 1 Appendix. Associated notes are explained in Table 1.

               * Values for this size fraction are calculated from the average measured values reported for the other two size fractions.

-------
          PM10 COMPOSITION  (24-hr AVG)
                                        WESTERN U.S.
Units = ug/m3
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Suifate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
8(a,q)
Res.Tr.Pk
1980
24
3
36.93





0.679
0.002

0.010
0.121


0.002
0.026
0.302
0.216

0.006


0.001
0.042
0.119

3.058

0.002
1.737


0.021

0.025
9(g)
Los Angeles
Summer'87
4,5 and 7
1 1 days
67.40
11.61
3.19
9.47
11.28

0.758
0.007
0.070
0.016
0.585

1.119
0.023
0.022
0.836
0.237
0.335
0.033

1.632
0.005
0.187
0.084

3.353

0.008
2.040

0.018
0.077
0.005
0.114
9(9)
Los Angeles
Fall'87
4 and 6
6 days
98.70
23.35
8.49
27.50
5.39

0.847
0.019
0.127
0.072
1.190

0.880
0.042
0.178
2.192
0.460
0.287
0.063

0.518
0.005
0.099
0.251

2.262

0.010
2.162

0.024
0.165
0.009
0.293
10(i)
San Joaquin Valley
6'88-6'89
24
~35
74.05
10.59
5.62
10.55
3.62

3.570

0.051
0.015
1.057

0.487
0.010
0.087
1.633
0.820

0.037


0.010
0.059
0.061
0.004
1.463

0.001
8.037

0.014
0.147
0.014
0.094
11(D* 5(d)
Phoenix Boise
10/13/89-1/17/90 12/86-3/87
7am-7pm-7a
6h,2x/day 12
~ 100 days
62.45
14.56
8.30
4.46
2.34

2.67
BQL
0.01
0.01
2.10
BQL
0.56
0.01
0.04
1.47
0.88
BQL
0.05
BQL
BQL
0.01
0.05
0.06
BQL
0.62
BQL
BQL
7.44
BQL
0.01
0.14
BQL
0.09
12(f) 8(a,q)*
Nevada Tarrant CA
11/86-1/87 1980
00-2400
24 24
24 6
100.90





2.407


0.149
4.543


0.007
0.077
1.257
0.441

0.067


0.006
0.002
0.786

2.888


5.791


0.093

0.147
8(a,q)*
Five Points CA
1980
24
3
124.37





7.317


0.019
1.786

0.026
0.007
0.037
3.275
1.437

0.055


0.037
0.155
0.105

1.422

0.001
16.657


0.277
0.013
0.032
8(a,q)*
Riverside CA
1980
24
4
106.20





3.549


0.065
5.082

0.173
0.005
0.061
2.015
1.081

0.049


0.013
0.144
0.489

2.373

0.001
7.778


0.182
0.003
0.059
 ON
 >

 to
 O
 O

 Z
 O
 H

O
 C
 O

 S
o
          References are listed

          * Values for this size
in Table 1 Appendix. Associated notes are explained in Table 1.

fraction are calculated from the average measured values reported for the other two size fractions.

-------
            PM10 COMPOSITION  (24-hr AVG)
                                                           CENTRAL U.S.
                                                                                  Units = ug/m3
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
8(a,q)*
San Jose CA
1980
24
6
66.68





2.053
0.001

0.250
0.771

0.480
0.009
0.071
1.214
0.508

0.027


0.014
0.045
1.119

1.109


5.506


0.086
0.002
0.105
8(a,q)*
Honolulu
1980
24
1
46.90





2.992


0.023
1.981

1.456
0.009
0.025
1.384
0.665

0.034


0.005
0.002
0.093

0.571


6.129


0.130
0.001
0.019
8(a,q)*
Winnemucca
1980
24
5
65.42





6.925


0.010
2.177

0.176
0.006
0.043
1.995
1.200

0.044


0.003

0.063

0.573


12.817


0.173

0.026
8(a,q)*
Portland
1980
24
4
117.55





6.932
0.014

0.121
1.459

0.197
0.019
0.109
2.059
0.805

0.108


0.036
0.028
0.537

2.371

0.001
12.505


0.191
0.018
0.119
8(a,q)*
Seattle
1980
24
1
36.00





2.296
0.008

0.033
0.585

0.228
0.005
0.041
1.001
0.231

0.022


0.007
0.006
0.292

0.952

0.001
4.424


0.091

0.093
8(a,q)* 13(q)*
Albuquerque Denver
12/84-3/85 1/11-30/82
7am-7pm-7am 6am-6pm-6am
12 12
~26
56.46
7.11
2.15
2.22
2.45

3.294
<0.004
0.089
0.127
0.705
0.018
1.287
<0.018
0.018
1.033
0.727

0.031


0.008
0.155
0.424
0.005
0.709
< 0.004
<0.004
7.737
<0.004
0.009
0.09
<0.004
0.085
14(m) 14(aa) 15(s)*
Urban Denver Non-urban Denver Chicago
1 1 /87-1 /88 1 1 /87-1 /88 7/94
9am-4pm-9am 9am-4pm-9am 0800-0800
7&17 7&17 24
~136 ~150 16
28.54
5.39
1.31

5.46

0.269
<0.0043
<0.130
0.011
0.761
<0.041
0.047
<0.0073
0.017
0.432
0.161
0.118
0.013
<0.0041
0.022
<0.0018
0.035
0.032

1.363
<0.059
<0.0017
0.813
<0.070

0.019
<0.013
0.090
to
H
6
O
2
O
H
O
c:
O
H
W
O
HH
H
W
References are listed
* Values for this size
in Table 1 Appendix. Associated notes are explained in Table 1.
fraction are calculated from the average measured values reported for the other two size fractions.

-------
                    PM10 COMPOSITION (24-hr AVG1
                                       CENTRAL U.S.
Units = ug/m3
o\
>
to
o
o
*
o
H
o
3
o
h-H
H
W
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
16(q)*
Houston
9/10-19/80
12
20
63.40
8.78
1.42
2.22
15.52

1.216
<0.015
0.139
0.091
2.935
<0.012
0.398
0.007
0.046
0.766
0.289

0.035

<1.49
0.008
0.128
0.589
<0.006
4.83
0.006
<0.003
3.200


0.036
< 0.045
0.142
6,7(p,q)
Harriman
5/80-5/81
0000-0000
24
256
32.50



11.14




0.052
1.800

0.050


0.690


0.038


0.001

0.237

2.500

0.002
2.000



ND

17* 6,7(p,q)
Harriman Kingston
9/85-8/86 5/80-6/81
0000-0000
24 24
330 169
30.00 35.40



8.70 13.63
36.1



0.056
0.960

0.018


0.360


0.027


ND

0.234

2.400

0.002
1.900



ND

6,7(P,q)
Portage
3/79-5/81
0000-0000
24
271
18.20



7.29




0.014
0.380

0.083


0.230


0.009


0.001

0.074

1.500

0.001
0.980



ND

6,7(p,q)
Topeka
8/79-5/81
0000-0000
24
286
26.40



6.60




0.055
2.400

0.031


0.580


0.020


0.001

0.203

1.200

0.000
2.500



ND

8(a,q)*
El Paso
1980
24
10
76.21





2.903
0.037

0.103
3.964

0.043
0.004
0.083
0.946
0.623

0.027


0.002

0.672

1.072

0.003
5.813


0.080

0.112
8(a,q)*
Inglenook
1980
24
8
72.45





2.508
0.001

0.061
2.924

0.003
0.006
0.059
1.474
0.717

0.078


0.003
0.030
0.388

2.969

0.001
6.997


0.116

0.188
8(a,q)*
Braidwood
1980
24
1
56.90





2.020
0.002

0.006
1.490


0.002
0.044
0.727
0.355

0.018


0.002
0.014
0.054

2.632

0.002
5.987


0.083

0.023
8(a,q)*
Kansas City KS
1980
24
8
70.33





2.144
0.003

0.036
4.371


0.010
0.048
0.989
0.660

0.026


0.005
0.013
0.237

2.031

0.001
4.976


0.076

0.060
                    References are listed

                      Values for this size
in Table 1 Appendix. Associated notes are explained in Table 1.

fraction are calculated from the average measured values reported for the other two size fractions.

-------
             PM10 COMPOSITION (24-hr AVG)
                                            CENTRAL U.S.
Units = ug/m3
 Ui
to
ON
H
6
o
z
o
H
O
d
o
H
ffl
0
H-t
H
W
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
8(a,q)*
Minneapolis
1980
24
6
46.35





2.191
0.005

0.069
1.674

0.293
0.003
0.057
0.831
0.402

0.031


0.002

0.406

1.131

0.001
4.848


0.062

0.072
8(a,q)*
Kansas City MO
1980
24
3
58.43





2.284
0.010

0.093
3.967

0.530
0.006
0.036
1.119
0.503

0.031


0.003

0.478

1.043


4.986


0.074

0.086
8(a,q)*
Akron
1980
24
7
70.90





2.555
0.015

0.064
1.541

0.572
0.024
0.055
2.249
0.592

0.129


0.011
0.059
0.509

3.870

0.008
5.531


0.116

0.219
8(a,q)*
Cincinnati
1980
24
2
62.95





2.972
0.013

0.041
1.374

0.103
0.005
0.038
1.057
0.499

0.032


0.007
0.080
0.442

3.265

0.005
6.961


0.099

0.201
8(a,q)*
Buffalo
1980
24
14
83.32





3,000
0,009

0.015
2.768

0.728
0.017
0.048
2.711
0.516

0.111


0.017
0.060
0.467

4.471

0.005
2.916


0.051
0.001
0.121
8(a,q)*
Dallas
1980
24
4
61.55





1.405
0.039

0.274
4.127

0.029
0.010
0.066
0.968
0.335

0.035


0.004
0.018
1.318

1.754


3.652


0.058
0.002
0.084
8(a,q)*
St. Louis
1980
24
5
56.82





3.956
0.004

0.046
1.874

0.053
0.001
0.032
0.663
0.417

0.019


0.004
0.020
0.372

2.612

0.002
4.638


0.058

0.044
18(x)*
St. Louis
8-9/76
6-12
62.00





1.412
0.003
0.054
0.179
2.949
0.005
0.344
0,015
0.043
1.493
0.653

0.071


0.009
0.099
0.877
0.002
5.188
0.007
0.005
4.928
0.010
0.009
0.587
0.006
0.175
6,7(p,q) 17*
St. Louis St. Louis
9/79-6/81 9/85-8/86
0000-0000
24 24
306 311
31 .40 27.60



11.14 8.00
9.7



0.099
1.600

0.145


0.770


0.040


0.005

0.415

2.300

0.002
2.100



ND

6,7(p,q)
Steubenville
4/79-4/81
0000-0000
24
499
46.50



17.60




0.052
1.120

0.303


2.200


0.068


0.008

0.259

5.500

0.005
2.300



0.013

             References are listed
             * Values for this size
in Table 1 Appendix. Associated notes are explained in Table 1.
fraction are calculated from the average measured values reported for the other two size fractions.

-------
                                                             EASTERN U.S.
>
K)
H

6
o

2
o
H

O
G
O
H
W

O
»

n
H-H
H
W
                                                                                  Units =  ug/m3
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
1(0)
Smoky Mtns
9/20-26/78
12
12
5.60





< 0.300
<0.001

0.005
0.322

<0.012

<0.005
0.118
0.108




<0.002

0.014

<0.560

<0.0006
0.580


0.018

<0.004
1(0)
Shenandoah
7/23-5/08/80
12
28
7.40



0.78

0.311
<0.002

0.003
0.304

0.179

0.006
0.158
0.129

<0.006


<0.003

0.009

<0.711

<0.001
0.813


0.017

0.006
2(b)
Camden
7/1 4-8/1 3 '82
6am-6pm-6am
12
50
H.4u
<3.00
0.42
0.57
<0.90

0.550


0.015
0.360
<0.006
0.069
<0.009

0.490
0.151

0.011


0.004

0.054

0.230
0.181
<0.0015
1.610
<0.009
0.002
0.065
0.007
0.030
3(ab) 4(c) 5(d) 5(d)
Philadelphia Deep Creek Raleigh Roanoke
7/25-8/14/94 8/83 1/85-3/85 10/88-2/89
4x daily 7am-7pm-7a 7am-7pm-7am
24 6 12 12
21 98
8.42





0.325


0.003
0.421

0.047

0.014
0.352
0.100
0.104
0.006

0.136
0.002
0.027
0.013

BQL

BQL
0.933


0.030
BQL
0.052
6,7(o,p)*
Watertown
5/79-6/81
0000-0000
24
354
9.30



2.44




0.022
0.209

0.305


0.276


0.006




0.076

0.200


1.000





8(a,o)
Hartford
1980
24
2
27.85





1.875


0.046
0.864

0.302
0.008
0.026
1.070
0.310

0.021


0.005
0.033
0.171

0.428


4.517


0.094
0.008
0.054
8(a,o)
Boston
1980
24
1
105.60





3.458
0.001

0.025
1.069

0.301
0.004
0.023
1.612
0.533

0.029


0.022
0.016
0.177

0.502


6.760


0.154
0.008
0.054
References are listed in Table 1 Appendix. Associated notes are explained in Table 1.

* Values for this size fraction are calculated from the average measured values reported for the other two size fractions.

-------
 OS
 N>
 oo
 H

 6
 O
O
H
w

O
fo
O
H
W
COARSE COMPOSITION (24-hr AVG)
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
8(a,o)
Res.Tr.Pk
1980
24
3
8.17





0.606


0.003
0.086


0.002
0.010
0.182
0.068

0.003




0.013

0.223


1.387


0.021

0.007
9(9)*
Los Angeles
Summer'87
4,5 and 7
1 1 days
26.30
3.34
0.82
5.13
1.87

0.723
BQL
0.055
0.003
0.563

1.026
0.002
BQL
0.737
0.196
0.311
0.017

1.431
BQL
0.127
0.046

0.520

BQL
1.988

BQL
0.072
BQL
0.024
9(9)*
Los Angeles
Fall'87
4 and 6
6 days
8.50
4.89
1.21
4.86
1.01

0.597
0.004
0.084
0.006
0.854

0.426
0.017
BQL
1.635
0.243
0.212
0.021

0.052
BQL
0.053
0.066

0.264

BQL
1.642

BQL
0.106
0.003
BQL
10(i)*
San Joaquin Valley
6' 88-6' 89
24
~35
44.17
5.71
2.38
2.38
0.62

3.418
0.000
0.040
0.006
0.961

0.393
0.007
BQL
1.453
0.632
0.000
0.031

0.000
BQL
0.052
0.032

0.222

0.000
7.577

0.012
0.130
BQL
0.016
WESTERN U.S.
11 (i) 5(d)
Phoenix Boise
10/13/89-1/17/90 12/86-3/87
7am-7pm-7a
6h,2x/day 12
~ 100 days
33.09
4.46
0.84
0.86
0.37

2.539
<0.002
<0.077
0.002
1.929
<0.016
0.194
0.008
0.021
1.259
0.669

0.032
< 0.005

0.003
0.038
0.022
0.003
0.178
<0.030
< 0.002
7.013
< 0.026
0.014
0.121
<0.014
0.034
Units =
ug/m3
12(f) 8(a,o) 8(a,o)
Nevada TarrantCA Five Points CA
11/86-1/87 1980
00-2400
24 24
24 6
43.85





2.230


0.047
4.088


0.005
0.030
0.941
0.255

0.035


0.003
0.002
0.167

0.310


5.208


0.083

0.052
1980
24
3
92.57





7.078


0.004
1.636

0.022
0.006
0.013
3.059
1.193

0.050


0.012
0.148
0.018

0.293


16.001


0.272
0.007
0.016

8(a,o)
Riverside CA
1980
24
4
71.63





3.513


0.028
4.781

0.164
0.005
0.021
1.888
0.961

0.042


0.006
0.144
0.113

0.720


7.544


0.182

0.030
            References are listed in Table 1 Appendix.  Associated notes are explained in Table 1.

            * Values for this size fraction are calculated from the average measured values reported for the other two size fractions.

-------
 VO
 VO
 (Ji
0\
 O
 O

 1
 O
 H

O
 c
 O
 H
n
i— i
H
W
COARSE COMPOSITION (24-hr AVG)
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
8(a,o)
San Jose CA
1980
24
6
30.40





1.930


0.062
0.682

0.430
0.006
0.028
1.066
0.260

0.021


0.008
0.032
0.228

0.257


5.214


0.086

0.044
8(a,o)
Honolulu
1980
24
1
25.80





1.865


0.006
0.957

0.938
0.005
0.007
0.658
0.294

0.014


0.003

0.022

0.258


3.766


0.067

0.008
8(a,o)
Winnemucca
1980
24
5
55.74





6.564


0.004
1.934

0.176
0.006
0.017
1.764
1.051

0.041


0.002

0.021

0.215


11.903


0.164

0.015
8(8,0)
Portland
1980
24
4
80.38





6.351
0.002

0.028
1.305

0.176
0.010
0.037
1.789
0.587

0.056


0.009
0.011
0.115

0.427


12.128


0.186
0.004
0.038
CENTRAL U.S.
8(a,o)
Seattle
1980
24
1
25.30





2.294
0.002

0.014
0.548

0.228
0.003
0.017
0.903
0.151

0.018


0.001

0.077

0.121


4.332


0.091

0.034
5(d) 13(o)
Albuquerque Denver
12/84-3/85 1/11-30/82
7am-7pm-7am 6am-6pm-6a
12 12
~26
35.73



0.39

2.900

0.058
0.024
0.658
0.012
1.235
<0.009
0.008
0.954
0.648

0.021


0.005
0.113
0.099
0.005
<0.48


7.460

0.009
0.090

0.039
Units = ug/m3
14(m) 14(ab) 15(s)
Urban Denver Non-urban Denver Chicago
1 1 /87-1 /88 1 1 /87-1 /88 7/94
9am-4pm-9am 9am-4pm-9am 0800-0800
7&17 7&17 24
~136 ~150 16
14.97





0.223
< 0.001 3
<0.038
0.007
0.716
<0.012
0.036
< 0.0024
0.006
0.344
0.101
0.106
0.008
<0.0017
<0.017
< 0.0007
0.027
0.005

0.043
<0.017
< 0.0006
0.739
<0.021

0.019
<0.004
0.038
References are listed in Table 1 Appendix. Associated notes are explained in Table 1.

* Values for this size fraction are calculated from the average measured values reported for the other two size fractions.

-------
COARSE COMPOSITION (24-hr AVG)
CENTRAL U.S.
Units = ug/m3
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
8
Sb
Se
Si
Sn
Sr
Ti
V
Zn
16(o)
Houston
9/10-19/80
12
20
24.80
3.10

1.63
0.91

1.093
<0.006
0.091
0.036
2.780
<0.006
0.366
0.007
0.018
0.604
0.170

0.021

<0.74
0.004
<0.1
0.124
<0.003
<1.29
<0.009

2.990
<0.009
<0.008
0.036
<0.03
0.058
6,7(o,p)* 17
Harriman Harriman
5/80-5/81 9/85-8/86
0000-0000
24 24
256 330
1 1 .70 9.00



3.04




0.014
1.650

0.029


0.570


0.021


0.001

0.057

BQL


1.880





6,7(o,p)*
Kingston
5/80-6/81
0000-0000
24
169
10.80








0.012
0.840

0.018


0.263


0.018


BQL

0.040

BQL


1.700





6,7(o,p)*
Portage
3/79-5/81
0000-0000
24
271
7.20



0.48




0.003
0.335

0.056


0.181


0.006


0.001

0.013

BQL


0.905





6,7(o,p)*
Topeka
8/79-5/81
0000-0000
24
286
13.90



0.55




0.010
2.150

0.000


0.490


0.016


0.001

0.040

BQL


2.310





8(a,o)
El Paso
1980
24
10
49.05





2.748
0.012

0.033
3.632

0.043
0.003
0.047
0.812
0.496

0.023


0.001

0.191

0.249

0.001
5.377


0.077

0.057
8(a,o)
Inglenook
1980
24
8
40.43





2.426


0.021
2.598


0.004
0.027
1.193
0.309

0.041


0.002
0.022
0.079

0.314


6.312


0.116

0.055
8(a,o)
Braidwood
1980
24
1
28.70





1.931
0.002

0.003
1.406


0.002
0.020
0.656
0.303

0.017


0.001
0.014
0.013

0.572

0.001
5.767


0.083

0.012
8(a,o)
Kansas City KS
1980
24
8
41.67





2.284
0.003

0.029
3.754

0.530
0.004
0.015
0.979
0.361

0.025


0.002

0.109

0.280


4.809


0.074

0.040
References are listed in Table 1 Appendix. Associated notes are explained in Table 1.
* Values for this size fraction are calculated from the average measured values reported for the other two size fractions.

-------
              COARSE COMPOSITION  (24-hr AVG)
                                   CENTRAL U.S.
Os
U)
 H

 a
 o

 2
 o


o
 a
 o
 H
Q


W
Units = ug/m3
Ref
Site
Dates
Hours
Dur
Number
Mass
OC
EC
Nitrate
Sulfate
Acidity
Al
As
Ba
Br
Ca
Cd
Cl
Cr
Cu
Fe
K
Mg
Mn
Mo
Na
Ni
P
Pb
Rb
S
Sb
Se
Si
Sn
Sr
Ti
V
Zn
8(a,o)
Minneapolis
1980
24
6
30.85





2.191
0.001

0.022
1.571

0.293
0.002
0.022
0.744
0.310

0.026


0.001

0.098

0.224


4.679


0.062

0.027
8(a,o)
Kansas City MO
1980
24
3
41.67





2.284
0.003

0.029
3.754

0.530
0.004
0.015
0.979
0.361

0.025


0.002

0.109

0.280


4.809


0.074

0.040
8(a,o)
Akron
1980
24
7
34.81





2.509
0.003

0.025
1.431

0.572
0.014
0.018
1.640
0.324

0.044


0.005

0.097

0.451


5.009


0.107

0.069
8(a,o)
Cincinnati
1980
24
2
33.15





2.910


0.017
1.312

0.103
0.002
0.014
0.883
0.363

0.021


0.003
0.037
0.099

0.389


6.633


0.096

0.148
8(a,o)
Buffalo
1980
24
14
44.57





2.808


0.012
2.550

0.728
0.015
0.022
2.040
0.206

0.078


0.009

0.108

0.765


2.675


0.051

0.043
8(a,o)
Dallas
1980
24
4
32.63





1.294
0.006

0.051
3.436

0.029
0.005
0.023
0.720
0.210

0.020


0.002

0.252

0.240


3.210


0.051

0.030
8(a,o)
St. Louis
1980
24
5
33.76





3.837
0.001

0.021
1.784

0.053
0.001
0.014
0.587
0.291

0.017


0.002

0.095

0.279


4.468


0.058

0.021
18(k,r)
St. Louis
8-9/76
6-12
28.00





1.209
0.001
0.034
0.047
2.817
0.001
0.257
0.009
0.014
1.218
0.392

0.035


0.005
0.098
0.189
0.002
0.533
0.001
0.001
4.470
0.001
0.007
0.475
0.004
0.074
6,7(o,p)* 17
St. Louis St. Louis
9/79-6/81 9/85-8/86
0000-0000
24 24
306 311
12.40 9.90



3.04




0.021
1.499

0.093


0.580


0.019


0.002

0.088

0.200


1.940



SQL

6,7(o,p)*
Steubenville
4/79-4/81
0000-0000
24
499
16.90



4.80




0.010
1.023

0.211


1.610


0.039


0.004

0.043

0.800


2.010



0.002

              References are listed in Table

              * Values for this size fraction
1 Appendix. Associated notes are explained in Table 1.

are calculated from the average measured values reported for the other two size fractions.

-------
       Table 6A-3
Selected Ratios of Mass Components

FM/CM
FM/PM10
Tot Carbon/FM
NH42SO4/FM
EAST
Mean N
2.59 8
0.65 8
0.25 7
0.47 12
WEST
Mean N
0.89 11
0.41 11
0.54 5
0.15 13
CENTRAL
Mean N
1 .06 25
0.51. 25
0.64 5
0.39 28
         N = number of studies contributing to the calculated means.
         Tot Carbon = (OCx1.4 + EC).
April 1995
                                    6A-32     DRAFT-DO NOT QUOTE OR CITE

-------
                                   Table 4a.  Site-to-Site Variability of PM2.5 Concentrations (ug/m3)
OS
u>
 O
 O

 Z
 O


o
 c
 o
 H
 W


 §

 o
 HH
 H
 W
Study Area
No. of Sites
Study Dates
Reference

Fine Mass
OC
EC
Nitrate
Sulfate
Al
Br
Ca
Cl
Cr
Cu
Fe
K
Mn
Ni
Pb
S
Si
Ti
V
Zn
Denver Metropolitan
3,a
11/2/87-1/31/88
14
Mean
19.67
7.25
4.41
3.96
1.55
0.037
0.018
0.058
0.141
0.003
0.017
0.111
0.077
0.012
0.002
0.075
0.642
0.272
0.009

0.031
Spread
2.89
0.79
0.78
0.93
0.16
0.005
0.006
0.001
0.013
0.002
0.008
0.023
0.009
0.003
0.002
0.017
0.077
0.009
0.001

0.008
Phoenix
3,b
10/13/89-1/17/90
11
Mean
29.38
10.09
7.49
3.60
1.33
0.131
0.011
0.167
0.366
0.003
0.015
0.216
0.209
0.023
0.003
0.039
0.436
0.430


0.056
Spread
3.49
2.69
1.71
0.37
0.24
0.015
0.003
0.033
0.356
0.001
0.003
0.035
0.020
0.010
0.001
0.009
0.038
0.066


0.030
Philadelphia
4,c
7/25/94-8/14/94
3
Mean
32.18
4.16
0.69

13.43
0.114
0.009
0.058
0.026

0.007
0.127
0.060
0.003
0.007
0.019
3.251
0.165

0.019
0.041
Spread
2.17
0.94
0.21

0.33
0.009
0.005
0.014
0.007

0.001
0.037
0.008
0.000
0.002
0.010
0.081
0.022

0.003
0.018
San Joaquin Valley
6,d
6/14/88-6/9/89
10
Mean
29.89
4.87
3.24
8.17
3.00
0.152
0.010
0.096
0.094
0.003
0.096
0.180
0.188
0.006
0.016
0.029
1.242
0.460
0.017
0.015
0.078
Spread
10.02
2.70
2.58
2.27
1.33
0.055
0.006
0.050
0.070
0.002
0.036
0.060
0.080
0.003
0.030
0.021
0.565
0.245
0.004
0.028
0.027
Mean = Mean over all sites of the average concentrations determined at each site for the sampling period.

Spread = ABS(Highest Mean Cone. - Lowest Mean Conc.)/2) for all the sites.

a.  Federal, Auraria, and Welby sites in urban Denver.


b.  Central Phoenix, Scottsdale, and Western Phoenix sites.

c.  Broad Street, Castor Avenue, Roxboro, and Northeast Airport sites.

d.  Stockton, Crow's Landing, Fresno, Kern, Fellows, and Bakersfield sites.

-------
                Table 4b.  Site-to-Site Variability of PM10 Concentrations (ug/m3)
Study Area
No. of Sites
Study Dates
Reference

Fine Mass
OC
EC
Nitrate
Sulfate
Al
Br
Ca
Cl
Cr
Cu
Fe
K
Mn
Ni
Pb
S
Si
Ti
V
Zn
San Jose
2,a
12/16/91 -2/24/92
29
Mean
64.95
19.39
9.02
10.90
2.24
0.845
0.012
0.670
0.728
0.003
0.029
0.834
0.823
0.014
0.003
0.035
1.147
2.905
0.088
0.007
0.065
Spread
1.65
0.15
0.42
0.60
0.09
0.035
0.001
0.048
0.032
0.001
0.002
0.027
0.021
0.001
0.000
0.004
0.091
0.045
0.024
0.003
0.005
Phoenix
3,b
10/13/89-1/17/90
11
Mean
62.47
14.55
8.33
4.46
1.70
2.670
0.014
2.096
0.559
0.011
0.036
1.475
0.878
0.054
0.006
0.062
0.615
7.442
0.121

0.090
Spread
7.06
3.48
1.78
0.45
0.29
0.273
0.003
0.317
0.349
0.002
0.009
0.170
0.083
0.014
0.002
0.013
0.041
0.862
0.024

0.034
San Joaquin Valley
6,c
6/14/88-6/9/89
10
Mean
62.92
7.87
3.51
9.44
3.57
2.993
0.012
0.950
0.388
0.009
0.084
1.413
0.720
0.030
0.019
0.039
1.472
7.517
0.128
0.022
0.085
Spread
17.28
4.15
2.76
3.02
1.46
1.570
0.005
0.390
0.225
0.003
0.046
0.445
0.220
0.011
0.032
0.027
0.605
1.765
0.033
0.030
0.029
 Mean = Mean over all sites of the average concentrations determined at each site for the sampling period.
 Spread = ABS(Highest Mean Cone. - Lowest Mean Conc,)/2) for all the sites.
 a. San Carlos St. and Fourth St. sites.
 b. Central Phoenix, Soottsdale, and Western Phoenix Sites.
 c. Stockton, Crow's Landing, Fresno, Kern, Fellows, and Bakersfield sites.
April 1995
6A-34      DRAFT-DO NOT QUOTE OR CITE

-------
  i       7.  HUMAN EXPOSURE TO PARTICULATE MATTER
  2             AMBIENT AND INDOOR CONCENTRATIONS
  3
  4
  5     7.1  INTRODUCTION
  6          The 1982 United States Environmental Protection Agency (U.S. EPA) Air Quality
  7     Criteria Document for Particulate Matter and Sulfur Oxides (PM-SOX AQCD) thoroughly
  8     reviewed the PM exposure literature through 1981.  The later  1986 "Second Addendum to
  9     Air Quality Criteria for Particulate Matter and Sulfur Oxides (1982)" added coverage of
 10     newly available health effects information with references up to 1986.  Consequently,
 11     literature directly concerning human exposure to PM has only  been previously reviewed
 12     thoroughly in an AQCD through 1981 and partially reviewed through 1986.
 13          This new analysis first summarizes key points from the exposure section of the 1982
 14     PM-SOX AQCD, and then reviews thoroughly the PM exposure literature from 1982 through
 15     1993 and includes 1994 literature published and in press through approximately September
 16     1994. Some additional literature available in 1994 and 1995 has also been included.
 17          The U.S. Environmental Protection Agency regulatory authority for PM only extends to
 18     the ambient air, defined in 40CFR as that portion of the atmosphere, external to buildings,
 19     available to the general public.  One major objective of this chapter is to examine the utility
 20     of centralized ambient PM monitoring data as a reasonable surrogate for the average of
 21     personal exposures to ambient PM of people in the surrounding community.  A secondary
 22     objective is to quantify the contribution of ambient air to the personal exposure.
 23          By the operative definition of ambient air, air inside a building or on private property is
 24     not regulated by the NAAQS.  However, it is important to consider total personal exposures
 25     to PM both from the regulated ambient air and non-regulated air.  This is because a variable
 26     fraction of ambient PM penetrates into different non-ambient settings where exposure to PM
 27     of ambient origin also takes place and, independently,  toxic PM can be generated within a
28     non-ambient setting (e.g. cigarette smoke).
29          Personal exposure to PM is important in itself, because it may give us clues as to  which
30     components of PM may be active or inactive biologically.  In addition, personal exposure
31     can act as a confounder in epidemiological studies which use an inferred community exposure
       April 1995                             7-1       DRAFT-DO NOT QUOTE  OR CITE

-------
 1     to ambient PM as a parameter to correlate with community health parameters.  On the other
 2     hand, an individual's personal exposure to total PM is a critical parameter for analysis if that
 3     person is a member of a cohort whose health outcomes are being tracked individually.
 4     Therefore, the chapter examines not only personal exposures to ambient PM, but also
 5     community and individual exposures to PM, which include that portion of ambient PM which
 6     penetrates into indoor microenvironments (pEs). This is to aid in the interpretation of those
 7     acute and chronic epidemiology studies of Chapter  14, in which ambient PM concentrations
 8     are assumed to be an indicator or a surrogate for mean community exposure to PM or an
 9     individual exposure to  ambient PM.
10          This chapter focuses on studies which include information on measurements of
11     simultaneous personal PM exposures, indoor-residential PM concentrations and ambient PM
12     concentrations.  Because people spend an average of 21 h per day indoors (Robinson and
13     Nelson, 1995) the indoor environment is the major exposure  category. Therefore, studies on
14     concentrations of PM indoors  are discussed below in Section 7.6 on Indoor Air.  The reason
15     for this separation of indoor concentration from personal exposure is explained below in the
16     general concepts  Section 7.1.2.
17
18     7.1.1  Ambient PM Concentration as a Surrogate for PM Dosage
19          The health effects of PM on an individual depend upon the mass and composition of
20     those particles which are deposited within the various regions of the respiratory tract during
21     the time interval of interest.  The amount of this potential-dose (FR, Part VI, EPA,
22     Guidelines for Exposure Assessment, May 29, 1992) will be  the product of the concentration
23     inhaled (e.g., the instantaneous personal exposure)  times the  ventilation rate (a function of
24     activity and basal metabolism) times the fractional deposition, which is a function of
25     ventilation rate and mode of breathing (e.g., oral or nasal).  If all people had identical
26     ventilation rates and deposition patterns, then the potential-dosage distribution could be
27     linearly scaled to the personal exposure distribution which  would serve as a suitable primary
28     surrogate.  The usage of ambient PM concentration in health studies as a surrogate for
29     personal PM exposure, and thereby a secondary surrogate for the PM dosage, would be
30     suitable if ambient concentration was also linearly related to  the personal exposure (Mage,
31     1983).

       April 1995                                7-2       DRAFT-DO NOT QUOTE OR CITE

-------
 1           The ventilation rate, which is at a minimum during the night while asleep (~ 6 L/m)
 2      and at a maximum during the day while awake (~  12 L/m) is in phase with PM exposure
 3      which is also lower at night than during the day (Clayton et al., 1993).  Consequently the
 4      product of the 24-h average exposure, the 24-h average ventilation rate, and the average
 5      deposition parameter for the  average ventilation, would seriously underpredict the amount of
 6      particulate matter deposited in the respiratory tract (Mage,  1980).
 7           In practice, when relating human health to pollution variables, one is forced to use
 8      ambient concentration as a surrogate for exposure and dosage because there are typically only
 9      fragmentary data on personal exposures in populations. Data are also limited on ventilation
10      rates as a function of basal metabolism and physical activities.  Furthermore, there are
11      virtually no applicable data on the deposition rates of the particles which people are inhaling
12      since the size distribution is unknown and deposition is influenced by individual physiological
13      parameters which are unmeasured.  According to Hodges and Moore (1977), "even when an
14      explanatory variable (ambient PM concentration) can be measured with negligible error it
15      may often be standing as a proxy for some other variable (dosage) which cannot be measured
16      directly, and so it (dosage) is subject to measurement error". This measurement error can
17      produce a negative bias on the relationship between health effects and PM dosage,  which
18      may be a partial explanation  of why many previous studies have found a positive but not
19      statistically  significant relationship between health effects and ambient PM concentration
20      (Pickles, 1982).
21           In the sections of this Chapter that follow, the  relationships between  ambient  PM
22      concentration, indoor PM concentrations and personal  exposures to PM are discussed in
23      detail.  The reader should keep in mind the following two caveats while going through this
24      chapter:
25
26           1.  Both ambient PM concentration and personal  exposure to PM are surrogates for the
27              amount of PM deposited in people's respiratory tracts.  Even  this quantity is a
28              surrogate for the true (but unknown) species and/or fraction of total PM that is the
29              specific etiological toxic agent(s) that act by an unknown mechanism.  This
30              unknown quantity should be the independent variable for developing the underlying
31              relationship of ambient PM and PM exposure to the health indices used as the
32              dependent variables.
33
34           2.  Virtually all analyses and discussions presented below  are based upon the personal
35              exposure to PM of non-smokers.  Only Dockery and Spengler (1981) included 6

        April 1995                                7.3        DRAFT-DO NOT QUOTE OR CITE

-------
 1              smokers out of 37 subjects.  Smokers are often excluded from these studies because
 2              a personal exposure monitor (PEM) on a smoker will not capture the main-stream
 3              tobacco smoke that is directly inhaled. In Section 7.6 on indoor air pollution, it is
 4              shown that side-stream environmental tobacco smoke (ETS) is the largest
 5              identifiable indoor source of PM.  For the average smoker, the amount of direct
 6              inhalation (several milligrams of PM per cigarette) can be two-to-three orders of
 7              magnitude greater than the microgram amounts of ETS which the PEM captures
 8              (Repace and Lowery, 1980).  The relationships presented below, of ambient PM to
 9              individual PM exposure, only apply to non-smokers.
10
11     7.1.2  General Concepts for Understanding PM Exposure and
12             Microenvironments
13          PM represents a generic class of pollutants which requires a different interpretation of
14     exposure in contrast to that for the other specific criteria gaseous pollutants, such as  CO
15     (Mage, 1985).  Whereas a molecule of CO emitted from a motor vehicle is indistinguishable
16     from a molecule of CO emitted from  a fireplace,  a 1-pim aerodynamic  diameter (A.D.)
17     particle emitted from the motor vehicle and a l-pim A.D. particle emitted from the fireplace
18     can have different shape, different mass, different chemical composition,  and different
19     toxicity. A l-/^m "particle" can be a  single entity, or an agglomeration of smaller particles,
20     such as a small Pb particle bound to a larger crustal particle.  Furthermore,  indoor sources
21     of particles produce a wide variety of particles  of varying size and composition that people
22     will  be exposed to, as shown in Figure 7-1 (Owen et al.,  1992).  Given that the health effects
23     of inhalation of any particle can depend upon its mass and chemical composition, it would be
24     of use to measure PM exposure in terms of mass  and chemical composition as a function of
25     size  distribution (Mage, 1985).
26          The total exposure of an individual to PM during a period of time is composed of
27     exposure to a variety of different particles from a variety of different sources in a variety of
28     different microenvironments (/*E). A /iE was defined by Duan (1982)  as "a chunk of air
29     space with homogeneous pollutant concentration"; it has also been defined (Mage,  1985) as a
30     volume in space,  during a  specific time interval, during which the variance of concentration
31     within the volume is significantly less than the variance between that /xE and its surrounding
32     pEs.
       April 1995                               7-4       DRAFT-DO NOT QUOTE OR CITE

-------
                                                                p
-------
 1          For example, a kitchen with a wood stove can constitute a single ^E for total PM,
 2     when the stove is off, and all people in the kitchen would have similar PM exposures. When
 3     the stove is in operation, the kitchen could have a significant vertical PM concentration
 4     gradient and a child on the floor in a far corner and an adult standing at the stove could be
 5     exposed to significantly different PM concentrations.
 6          In a given /zE, such as one in the kitchen example, the particles may come from a wide
 7     variety of sources.  PM may be generated from within (e.g. the stove, deep frying, burning
 8     toast), from without (ambient PM entering through an open window), from another indoor
 9     juE (cigarette smoke from the living room), or from a personal activity that generates a
10     heterogeneous mix of PM (sweeping the kitchen floor and resuspending a mixture of PM
11     from indoor and outdoor sources that had  settled out).
12          In general, as a person moves through  space and time, they pass through a series of
13     /xEs and their average total exposure (X  /iig/m3) PM for the day can be expressed by the
14     following equation,
15
16                                       X =  EXjtj/Etj                            (7-1)
17
18     where Xj is the total exposure to PM in the ith /xE, visited in sequence by the person for a
19     time interval tj (Mage, 1985).  Individual human activity patterns determine the time-
20     sequence in which these /^Es may be visited  and, therefore, the magnitude of the overall
21     concentration to which a person is exposed.  Let two people on a given day spend 1-h
22     outdoors.  If one person is outside from 7 to 8 a.m and the other is outside from 7 to 8 p.m.
23     they can have significantly different PM exposures, neither of which would be characterized
24     by the midnight-to-midnight 24-h average.
25          With appropriate averaging over sets of 4  classes of /xEs (e.g. indoors,  ambient-
26     outdoors, occupational, and in-traffic we  can simplify the Equation 7-1 as follows (Mage,
27     1985):
28
29                      X  = (Xin tin  +  Xout tout +  Xocc tocc  +  Xtra !,„)  / T             (7-2)
30

       April 1995                               7-6       DRAFT-DO NOT QUOTE  OR CITE

-------
  1      where each value of X is the mean value of total PM concentration in the pE class while the
  2      subject is in it, time (t) is the total time the subject is in that pE during the day, and T is
  3      equal to the sum of all times (usually  1-day).   Similar equations may be written for personal
  4      exposures to particles from specific  sources (e.g. Diesel soot), for  specific chemicals (e.g.
  5      Pb), or for specific size intervals (PM  < 2.5^ AD).
  6           In reference to the situation cited above,  of two people on the same day spending an
  7      identical 1-h as tout at different times during the day, they would have to have two different
  8      values of Xout in their exposure calculation.  This is in accordance  with  the precept of Ott
  9      (1982) that an air pollutant exposure requires the confluence of two variables - the
10      concentration of PM X(x,y,z,t) at a location (x,y,z) and time t, and the position of an
11      individual's breathing zone at x,y,z,t.
12           In the  literature, many excellent  studies have reported data on air quality concentrations
13      in /iE settings that do not meet a rigorous definition of an exposure, which requires
14      occupancy by a person (Ott, 1982).  Section 7.6 on Indoor  Concentrations and Sources, cites
15      Thatcher and Layton (1994) who report that "merely walking into a room increased the
16      particle concentration by 100%  (from  10  to 20 jug/m3)", perhaps by air currents reentraining
17      PM. Consequently, a measurement of air quality in a space that includes time when it is
18      unoccupied may not be a valid measure that can be used to  estimate an exposure.  If this
19      measure includes the periods of time when the space was unoccupied it will tend to be  biased
20      low as a measure of the exposure within  it during periods of occupancy.
21           In the context of exposure, it may be inappropriate to  associate  an  average exposure to
22      a person while cooking at a stove in a kitchen  with a concentration measurement that is
23      influenced by periods  when people were not in the kitchen or when the stove was not in
24      operation.  It is therefore understood that an average concentration measured in an indoor
25      setting - including periods  when people are  not present - may not have direct relevence for
26      computing personal exposure because it is not necessarily the concentration during the
27      portion of time that the subject was  inside that /nE.
28           The literature on 24-h average PM concentrations in indoor juEs, such as those in
29      residential settings, are treated separately  in Section 7.6, as  is done for 24-h average ambient
30      PM concentrations in Chapter 6. In the exposure portion of this chapter, specific reference
31      will be  made to some  of those studies  where simultaneous personal exposures and indoor

        April 1995                                7.7       DRAFT-DO NOT QUOTE OR CITE

-------
 1     measurements have been made so that the relationship between indoor concentration and
 2     personal exposure can be examined.
 3          In practice a cascade sampler  can collect ambient PM samples by size fractionation for
 4     separate chemical analyses, but such a complete definition of personal exposure to PM by
 5     chemistry and size is impossible to  obtain. Although some personal monitors can be
 6     equipped with a cyclone or impactor separator and two filters to capture two sizes (e.g.,
 7     <2.5 />im and 2.5 to 10 /mi), because of the current size of a two filter sampler and the low
 8     mass collected in the two fractions, one almost exclusively obtains a single integrated
 9     measurement of particle mass collected (e.g.,  <2.5 /mi or < 10 jum). Consequently, health
10     studies on individuals are usually only able to  develop associations between their observed
11     health effects and their observed exposure expressed as an integral mass of PM collected and
12     its average chemical  composition.
13          Health studies  on populations  can make multiple measurements of ambient and indoor
14     PM concentrations simultaneously (e.g., PM2 5, PM10, TSP) along with components of PM,
15     such as polycyclic aromatic hydrocarbons (PAHs), to help understand the size distribution
16     and chemistry of the particles in the ambient and indoor atmospheres. However, these data
17     may be weakly correlated with simultaneous personal PM exposure measurements.
18
19     7.1.3  Review of State-of-knowledge Recorded in the 1982 PM-SOX AQCD
20          In 1982 it was  known,  from personal monitoring and indoor monitoring, that SO2 is
21     almost always lower indoors than outdoors because of the virtual absence of indoor sources
22     for SO2 and the presence of  sinks for SO2 in indoor settings (Exceptions can occur if high
23     sulfur coal or kerosene are used as fuel in a poorly vented stove or space heater).  However,
24     this relationship does not hold for PM as  the indoor and personal monitoring data show both
25     higher- and lower-than ambient PM concentrations in indoor settings  as a function of particle
26     size and  human activity patterns.
27          The largest coarse mode particles (>W  /mi), which are generally of nonanthropogenic
28     origin  (wind blown  dust, etc.), require turbulence to provide vertical  velocity components
29     greater than their settling velocity to allow them to remain suspended in the air (Figure 7-1).
30     Outdoor particles enter into  an indoor setting either by bulk flow, as  through an open
31     window, in which all particles can  enter at the inlet condition, or by  pressure driven drafts

       April 1995                                7-8        DRAFT-DO NOT QUOTE OR CITE

-------
  1      and diffusional flows through cracks and fissures in the barriers of the building envelope
  2      when all windows are closed.  In the latter mode of entry, velocities are relatively lower,
  3      thereby  settling out the largest coarse particles in the passage through the barriers.
  4           Indoor settings are usually quiescent (Matthews et al.,  1989), and the larger ambient
  5      particles that do enter indoors quickly settle out  by gravity and electrostatic forces, leading to
  6      the presence of the familiar dust layers on horizontal surfaces and vertical TV screens that
  7      require constant cleaning (Raunemaa et al.,  1989).  However, human activity in indoor
  8      settings, such as smoking,  dusting,  vacuuming and cooking, does generate fine particles
  9      (<2.5 ^m) and coarse particles (>2.5 ^m), and resuspends particles that previously had
 10      settled out (Thatcher and Layton, 1994; Litzistorf et al.,  1985).
 11           There were only three studies of personal PM exposures, compared to ambient PM
 12      concentrations, that were referenced in the PM-SOX criteria document. Binder et al. (1976)
 13      reported that "outdoor air measurements do  not accurately reflect the air pollution load
 14      experienced  by individuals who live in the area of sampling", in a study in Ansonia, CT,
 15      where personal exposures to PM5 were double the outdoor PM concentrations measured as
 16      TSP  (PM27)  (115 versus 58 /xg/m3). Spengler et al. (1980) was cited as reporting that "there
 17      was no correlation [R2 = 0.04] between the outdoor level [of respirable particles] and the
 18      personal exposure of individuals" in a study in Topeka, KS. Figure 7-2,  from Repace et al.
 19      (1980), was  presented as an example of the  variability  of PM exposures which show very
20      little  influence of ambient concentration. Consequently, at the time of writing the  1982 PM-
21      SOX AQCD, two major factors were known to influence the relationship of ambient to indoor
22      PM air quality.  They  were (1) the  variability of indoor concentrations of PM compared to
23      outdoor  concentrations as a function of particle size (e.g., fine indoor > fine outdoor and
24      coarse indoor < coarse outdoor) and  (2) the variation of exposures of individuals related to
25      the different  activities that are involved with the  local generation of particles  in their
26      immediate surroundings (smoking, traffic, dusting and vacuuming at home, etc.).   This
27      understanding was summarized as follows (pg. 5-136,  PM AQCD,  1982):
28
29
30           •  long term personal exposures to fine fraction PM (<2.5 /mi) of outdoor origin, may
31             be estimated by ambient measurements of the  <2.5 /xm PM fraction.
       April 1995                                7.9       DRAFT-DO NOT QUOTE OR CITE

-------
Tl
H
6
o
z
o
H
O
c!
O
H
W
n
HH
H
W
                 280

                 260

                 240

                 220
              oo
               E  200
               §
              O
              160

              140

              120

              100

               80

               60

               40

               20
                     1   I    I   I   I   I   I    I   I


                            • Indoors
                            • In Transit
                            O Outdoors
                                                           1    I   T
                                  V  I   I   I    I
                                                           i    i
                                                                       Well-Ventilated Kitchen
                                        Cafeteria, Smoking Section'


                                     Behind Smoky Diesel Truck I
             Office
Commuting •  ^  ^

       im
                               Outside Cigar
                               Smoker's Office

                                     Cafeteria,
                                    -Nonsmoking
                                     Section
                                  IO..K,,^B    Sidewalk
                                  iSuburbs    BusExhaust
                                     Vehicle     v^
                                       I-In City
                                                                              Office"
                      I   I    I
	Bedroom..

           Street Suburbs, OutdoorCN' / city  Outdoor
            Library Unoccupied Cafeteria   | y> |   |   |
Commuting   Room

       Suburbs
       ogging

LyingjRoOjm  |   |
                 12  1   2  3   4  5   6  7  8  9  10 11  12  1  2   3  4   5  6  7  8  9 10 11 12

              Midnight              A.M.               Noon               P.M.

                                                      Time of Day

Figure 7-2. An example of personal exposure to respirable particles.
2  Source: Repace et al. (1980)

-------
 1          •  Personal activities and indoor concentrations cause personal exposures to PM to vary
 2             substantially.  Ambient measurements appear to be a poor predictor of personal
 3             exposure to PM.
 4
 5          •  Tobacco smoke is an important contributor to indoor concentrations and personal
 6             exposures where  smoking takes place.
 7
 8
 9     7.2  DIRECT METHODS OF MEASUREMENT OF HUMAN
10           EXPOSURE TO PM BY PERSONAL MONITORING
11          Human exposure to  air  pollution can be measured by placing a personal exposure
12     monitor (PEM) close to the breathing zone of an individual.  However, the very act of
13     studying the subjects can  influence the measured value of their exposure and create an
14     erroneous reading.   These influences arise because the subjects become conscious of the
15     study objectives from the  indoctrination required to obtain their written informed consent to
16     participate, and the presence  of the PEM  on their body is a constant reminder.
17          The physical location of the monitor inlet, as worn  by the subject, can also influence
18     the subject's PM exposure and the recorded PM. The movements of the subject's body and
19     sampling flow rate can alter the air currents in  the subjects breathing zone.  "The presence of
20     the body and its movement affect what a personal sampler collects" (Ogden, 1993).  When in
21     close proximity to a source actively emitting PM (within  a meter) a small change in PEM
22     position (e.g. from left side to right side) can vary the PM measurement.
23          These unquantifiable 'errors' in a PM PEM measurement may be greater than the filter
24     weighing errors and flow  rate measurement errors that can be quality controlled through
25     calibration procedures.  This  may be important for interpretation of the PM PEM data in the
26     literature because the expectation is that these errors inflate the variance of the
27     measurements.  In the following section, the individual error components that arise from the
28     measurement process are discussed.
29
30     7.2.1  Personal Monitoring Artifacts
31     7.2.1.1  'The Hawthorne Effect'
32          If subjects carry a personal exposure monitor (PEM) they may change their behavior,
33     subconsciously or consciously, which is known as "The Hawthorne Effect" (Last, 1988).

       April 1995                               7-11      DRAFT-DO NOT QUOTE OR CITE

-------
 1      The name derives from early studies of worker productivity at Western Electric's Hawthorne
 2      plant in Cicero, IL (Mayo, 1960).  For instance a subject may choose not to go to the opera
 3      because the pump noise of the monitor would be disturbing to others.  Even though the
 4      exposure  measured in the alternative activity may be quite accurate, it would be an artifact in
 5      the context of exposure assessment, because the activity not performed (opera attendance)
 6      would have been the correct one to monitor.
 7
 8      7.2.1.2 The Monitor Effect
 9           The filtration of the breathing zone air by a PEM can reduce the PM concentration in
10      the breathing zone by "self dilution" (Cohen et al., 1984). The placement of the personal
11      monitor in the  breathing zone, as well as its flow rate, can alter the air  flow stream lines in
12      the area of the nose or mouth that would exist in its absence.  There may be an electrostatic
13      charge on the plastic cassette filter holder which can possibly affect the collection of charged
14      particles (Cohen et al.,  1982).
15
16      7.2.1.3 Subject Effect
17           A subject may contaminate the personal  monitoring data by  an inadvertent action, such
18      as forgetting to put on the PEM upon awakening in the morning,  or purposefully choosing
19      either not to wear it when going to the opera  (anti-'Hawthorne Effect')  or placing it close to
20      a source.  If such actions are not recorded in  a diary or reported  to the  investigator during a
21      verbal debriefing, the exposure data, although valid per se as an actual concentration
22      measurement, could be treated as a valid exposure of a subject in his/her daily life and
23      related  to an incorrect classification.
24         For example, Sexton et al (1984) reported that one male subject with a personal exposure
25      mean of 77 /xg/m3 had a spouse with a simultaneous personal exposure of 37 pig/m3, with no
26      recorded activity that could explain the higher exposure values.
27
28      7.2.1.4 Non-Response Error
29           In performance of a personal monitoring study, people often refuse to participate. The
30      refusal  rate  increases with the burden on the respondents due to the time required to
31      complete  questionnaires, diaries and the need  to carry the personal monitor with them

        April 1995                                7-12      DRAFT-DO NOT QUOTE OR CITE

-------
 1     throughout the study. If the cohort of people who refuse to participate have significantly
 2     different personal PM exposures than the participants, then the study will produce a biased
 3     estimate of the exposures of the total population.
 4
 5     7.2.2  Characterization of PM Collected by Personal Monitors
 6           The amount of PM collected by different types of personal monitors with the identical
 7     nominal cut-point can be variable.  Small differences in the cut point (AD at 50% efficiency)
 8     and in other portions of the penetration curve can exist under calm wind conditions when the
 9     monitor is stationary, and  these differences can also be a function of the air velocity relative
10     to the monitor.  Consequently, the difference between two PM measurements made by two
11     nominally identical monitors of different design, can be a function of the wind speed and the
12     size distribution of the PM in the air mass being sampled.  Given the knowledge of the
13     sampled size distribution (as collected) and the complete penetration curve, the concentration
14     in the free atmosphere can be predicted (Mage, 1985).  For example, if a sampler  collects 1
15     Mg/m3 °f PM in a size interval that has an overall efficiency of 0, then we can estimate  that
16     the atmosphere contained  1/0 /xg/m3 of PM of the same size range.   A recent field
17     comparison of different types of respirable dust samplers used in occupational settings where
18     coarse mode  paniculate is predominating, by Groves et al.  (1994), shows that there is a
19     considerable  difference between the mass collected by sets of paired cyclones sampling in a
20     concentration range of 0.5 to 6.6 mg/m3, which is much higher than normally seen in non-
21     occupational  personal exposure studies.  This type of comparison study has not been done for
22     personal monitors used in  nonoccupational studies at the ambient and indoor respirable
23     concentrations on the order of 0.01 to 0.1 mg/m3 where the fine mode is more important.
24
25     7.2.3  Microscale Variation and the  Personal  Cloud Effect
26           The tendency for human activity in the home  or at work to generate a 'personal activity
27     cloud' of particles from clothing and  other items (carpet, stuffed furniture, etc.), that will be
28     intense  in the breathing  zone, and diluted near an area monitor located  several meters  away,
29     has also been cited as a contributing factor to the discrepancy between personal measures of
30     exposure and time weighted average (TWA)  exposures using microenvironmental
31     measurements (Martinelli et al.,  1983; Cohen et al., 1984;  Rodes et al., 1991).  The

       April 1995                                7_13       DRAFT-DO NOT QUOTE OR  CITE

-------
 1     difference between sampling for PM at the nose, on the lapel or at a point several meters
 2     away from a person can be appreciable. The microscale variation in PM between a personal
 3     monitor sampling at the nose, versus an identical monitor sampling at the lapel, has been
 4     investigated by Cohen et al. (1982, 1984) who found no net bias resulted from sampling at
 5     either location in a concentration field of uniformly dispersed aerosol.  However, they noted
 6     that spatial concentration variability and resuspended dust from clothing, and to a lesser
 7     extent electrostatic charges  on plastic filter cassette holders  (and plastic eyeglass frames),
 8     could lead to different exposure measurements in the facial  region, with three times as much
 9     mass collected by a filter attached to the clothing as from the air sampled directly  in the
10     breathing zone.
11          Fletcher and Johnson  (1988) also measured metal concentrations (measurement method
12     and size unspecified) in an  occupational exposure situation (metal spraying of spindles on a
13     lathe) and found 50% higher concentrations measured from the left lapel compared to the
14     right lapel, which reflected the orientation of the operator to the lathe.  When a neutrally
15     buoyant tracer gas mixture  was released 0.5 m from an operator in a different work setting
16     there was no variation between the left and right lapels, and the nose, as found by Cohen et
17     al.  (1982).   However, when the experiment was performed with a heavier-than-air tracer gas
18     mixture,  the nose measurement was approximately 25% less than the lapel concentrations.
19     This implies that if submicron particles, which behave like a gas, are emitted close to the
20     subject in a buoyant plume, there may also be significant microscale variations in the
21     breathing zone.
22          Parker et  al. (1990) measured the aerosol distribution in a small test room resulting
23     from a nozzle-jet injection, using a "heated phantom"  (mannequin), and found larger
24     discrepancies between chest mounted monitors and area monitors up to three meters away,
25     by  up to a factor of ten difference.
26
27
28     7.3 NEW LITERATURE ON PARTICLE EXPOSURES SINCE 1981
29          The following sections review studies that measured PEM PM in the general non-
30     smoking population.  In these studies,  the subjects spent time at home and in other indoor
31     environments that include time at work. In the USA, recent data indicate that on a daily

       April 1995                               7-14      DRAFT-DO NOT QUOTE OR CITE

-------
 1      basis, an average US resident spends approximately 21 h indoors (87.2%), 100 minutes in
 2      (or near) a vehicle  (7.2%), and 80 minutes outdoors (5.6%) [Robinson and Nelson, 1995].
 3           The concentration of PM in residential and commercial indoor settings is thoroughly
 4      reviewed and presented in Section 7.6. However, the concentrations of PM and personal
 5      exposures to PM in 'indoor' industrial workplaces (e.g. coal mines) and the "dusty trades"
 6      (e.g. carpentry or machinist), which are covered by occupational standards for air quality,
 7      are not covered in  this document.
 8           Almost all the studies of PM exposure in the general public have been conducted on
 9      urban and suburban residents.  These subjects are often working in occupations that do not
10      require PM monitoring to  assure that occupational standards are being met (e.g. in an
11      office).  However,  PM monitoring in an industrial workplace by a subject - independently of
12      an official corporate industrial hygiene program - can have legal or security implications for
13      an employer.  For  example,  in a study in Tennessee (Spengler et al., 1985) some potential
14      subjects were unable to participate because their employer (Oak Ridge National Laboratory)
15      would not allow them to wear PM monitors at work.  Such exclusion of subjects from
16      exposure studies can negatively bias measured exposure distributions if the reason for the
17      exclusion is related to their potential for high PM exposure.
18           A further complication arises from the fact that industrial exposures tend to be
19      dominated by a specific type of particle.  Coal miners are exposed to coal dust, textile
20      workers are exposed to cotton dust and welders are exposed to metal fumes.  An additional
21      chapter on personal exposures to industrial PM would be needed to describe the various
22      industries and trades that have their own individual PM problems, and the usage pattern and
23      efficiencies of respirators and masks that are required to be worn.  Therefore, occupational
24      PM studies are not presented, and only some selected studies are  cited to illustrate a
25      particular point that is applicable to exposure studies in general, such as microscale variation
26      in PM when close to a source of PM.
27           It may be useful to keep in mind that the baseline exposure of nonsmoking workers in
28      the "dusty trades" or industrial workplaces may be similar to that of other nonsmoking
29      people in their communities.   Their total working day exposure would then be approximated
30      by that baseline exposure plus their incremental workplace exposure for 8 h, in a similar
        April 1995                                7-15       DRAFT-DO NOT QUOTE OR CITE

-------
 1     manner to a smoking increment for smokers.  On a day off from work, these worker's
 2     exposures may be similar to that of the general nonsmoking public.
 3
 4     7.3.1  Review of the Literature
 5     7.3.1.1 Results of U.S. Studies
 6           Dockery and Spengler (1981) compared personal PM3 5 exposures and ambient PM3 5
 7     concentrations in Watertown, MA, and in Steubenville, OH.  In Watertown, 24-h personal
 8     samples were collected on a 1-in 6-day schedule, and in Steubenville, 12-h personal samples
 9     (8 a.m. to 8 p.m.) were collected  on a Monday-Wednesday-Friday schedule.  A correlation
10     coefficient of 0.692 between the mean personal and the mean ambient concentration for 37
11     subjects, 18 in Watertown and  19  in Steubenville, was reported for the pooled data.
12     However, this appears to be an artifact of two separate clusters formed by these data,  each
13     with considerably lower correlation.  When these data are analyzed separately, the regression
14     coefficient between personal and ambient for Watertown is R = 0.01 and for Steubenville it
15     is R = 0.43.
16           Sexton, Spengler and Treitman (1984) studied personal exposures to respirable particles
17     (PM3 5) for 48 nonsmokers during a winter period  in Vermont, where firewood was either
18     the primary or secondary heating source for the subject.  Their results  showed that personal
19     exposures were 45% higher than indoor averages (36 /xg/m3 versus 25  ^g/m3) and indoor
20     averages were 45% higher than outdoor averages (25 /xg/m3 versus 17  /xg/m3). This
21     relationship is consistent with those reported in the 1981 PM-SOX AQCD (Spengler et al.,
22     1980).  Ambient air pollution,  measured by an identical stationary ambient monitor (SAM)
23     outside each residence (a pump contained  in a heated box was connected to an external
24     cyclone and filter),  had no correlation with the resident's personal exposures (R2 = 0.00)
25     and 95% of the subjects had personal exposures greater than the median outdoor
26     concentration. This would not contradict the  first conclusion cited from the 1982 PM-SOX
27     AQCD if there were sources of PM3 5 indoors.
28           Spengler et al. (1985) reported a study of PM3 5 exposures in the non-industrial cities of
29     Kingston and Harriman, Tennessee, during the winter months of February through March,
30     1981. A large TV A coal burning  power plant (Watts Bar) with very tall stacks in the
31     immediate area was not a local source of paniculate pollution.  In this  study, two

       April 1995                              7-16       DRAFT-DO NOT  QUOTE OR CITE

-------
 1     Harvard/EPRI PM3 5 monitors were used for each person.  One stationary indoor monitor
 2     (SIM) remained indoors in the home, in the open air of a first floor activity room, and the
 3     second monitor (PEM)  was carried for 24-h to obtain the personal exposure.  In each
 4     community, identical Harvard/EPRI samplers (SAM) were placed at a central site to
 5     represent ambient PM3  5 concentrations.  The results of the study are shown in Table 7-1. In
 6     both communities, 95% of the subjects had personal exposures to PM3 5 greater than the
 7     average ambient concentrations.  The mean personal exposure and indoor concentrations
 8     (44 ± 3 /ig/m3 and 42  +  3 jug/m3) were more than 100% greater than the mean  ambient
 9     average of 18 +2 /xg/m3  sampled on the same days.
10
11
               TABLE 7-1.  QUANTILE DESCRIPTION OF PERSONAL, INDOOR,
           AND OUTDOOR PM3 5 CONCENTRATIONS (in jtg/m3),  BY LOCATION IN
                               TWO TENNESSEE COMMUNITIES
City
Kingston


Harriman


Total K&Ha


Group
Personal
Indoor
Outdoor
Personal
Indoor
Outdoor
Personal
Indoor
Outdoor
N
133
138
40
93
106
21
249
266
71
95%
99
110
28
122
129
34
113
119
33
75%
47
47
22
54
45
23
48
46
23
50%
34
31
16
35
27
15
34
29
17
25%
26
20
12
24
18
13
26
20
13
5%
19
10
6
15
10
9
17
10
7
Mean
42
42
17
47
42
18
44
42
18
S.E.
2.5
3.5
2.7
4.8
4.1
4.0
2.8
2.6
2.1
      "Includes samples from 13 subjects living outside Kingston and Harriman town limits and from four field
       personnel residing in these communities.
      N = number of samples.
      S.E. = Standard error.
      Source:  Spengler et al. (1985).
1          For the complete cohort, the correlation between PM PEM and PM SAM was r = 0.07
2     (p = 0.30), and between PM PEM and PM SIM was r = 0.70 (p  = 0.0001).  The
3     correlation between simultaneous PM PEM and PM SAM was r = 0.15 for 162 nonsmoke


      April 1995                              7_17      DRAFT-DO NOT QUOTE OR CITE

-------
 1     exposed individual observations (p  = 0.06).  For 63 observations on smoke exposed
 2     individuals, the correlation r = 0.16 was not significant (p = 0.16).
 3          An important finding was that in nonsmoking households, the PM PEM is always
 4     higher than SIM and SAM.  "This  implies that individuals encounter elevated concentrations
 5     away from home, and/or that home concentrations are elevated while they are at home and
 6     reduced while they are away".  This observation is supported by the findings of Thatcher and
 7     Layton (1994), reported in Section  7.6 - merely walking into a room can raise the
 8     concentrations of PM by 100% (from 10 to 20 /xg/m3).
 9          This study is relevant to the analyses by Dockery et al. (1992) of PM mortality in St.
10     Louis, MO, and in Eastern Tennessee counties surrounding the cities of Kingston and
11     Harriman,  which are discussed in Chapter  12 on epidemiology.  Although the Spengler et al.
12     (1985) and Dockery et al. (1992) studies are  not directly comparable, because different years
13     of data were used (1981 versus 1985/1986), this might call into question the meaning of the
14     correlation of ambient PM and mortality,  if individual PEM data are  uncorrelated with the
15     SAM data.  One possible explanation may be that  even though individual PEM of non-
16     smokers as a group are not well correlated with SAM data (r =  0.15, p = 0.06) the mean
17     PEM of the nonsmoking population may be better  correlated with the mean SAM on any
18     given day as discussed in Section 7.5.2 (Mage and Buckley, 1995).
19          Morandi et al. (1988) investigated the relationship between personal exposures to PM
20     and indoor and outdoor PM concentrations, using a TSI Model 3500  piezobalance that
21     measures respirable particles in the range  <3.5 /xm. For  the group of 30 asthmatics in
22     Houston, TX, that were studied, outdoor concentrations averaged 22  /xg/m3,  indoor
23     concentrations averaged 22%  higher than outdoor (27 /xg/m3) and, in motor vehicles, the
24     average concentration of particles was 60% higher than the average outdoors (35/xg/m3).  As
25     for correlations between the various measurement categories, personal 12-h (7 a.m. to
26     7 p.m.) daytime exposures to PM were not predicted as well by fixed site dichotomous
27     sampler ambient monitors (R2 = 0.34) as by the indoor exposures (R2 = 0.57).  However,
28     for 1-h exposures, they found no correlation  (R2 = 0.00) between the personal exposures to
29     PM5 and the indoor exposures measured with a TSI model 5000 stationary continuous
30     piezobalance located in the 'den' area of the  home.  The authors noted that use of home air
31     conditioning and recirculation tended to increase the PM exposures,  and that misclassification

       April 1995                               7-18      DRAFT-DO NOT QUOTE OR CITE

-------
  1      of human activities and microenvironments likely explained the inability to predict personal
  2      exposures from microenvironmental data.
  3           Lioy et al. (1990) reported a study done during the winter (January 1988) in the
  4      industrial community of Phillipsburg, NJ,  where personal PM10 was monitored along with
  5      indoor and outdoor PM10.  They collected PM10 (fine plus coarse particles on a single  filter).
  6      In this study of eight residences of 14 nonsmoking nonsmoke-exposed individuals, the
  7      geometric mean 24-h concentrations were  68, 48 and 42 jug/m3 for personal, outdoor and
  8      indoor sites, respectively.  The arithmetic  mean personal PM exposure of 86 jwg/m3 was 45%
  9      higher than the mean ambient concentration of 60 /xg/m3.  The higher ambient than indoor
10      concentrations in this study, a reversal of the relationships found in the Sexton et al. (1984),
11      Spengler et al.  (1985) and Morandi et al. (1988) studies, may be caused by the local
12      industrial source of coarse particles in that community and the  absence of cigarette smokers
13      in the residences  sampled.  This difference also may be partially explained by the  10 /xm
14      particle sizes in the NJ  study and the 3.5 /xm particle sizes in the other studies, which would
15      suggest that the NJ homes had less influence from the locally generated coarse particles that
16      tend to settle out in the home.  The regression coefficient between personal and ambient
17      PM10 for all 14 people  on the 14 days of the study (n = 191 valid personal values) was 0.19
18      (R2 = 0.037, p = 0.008).  With three personal exposure extreme values removed (n = 188
19      personal values), the coefficient was 0.50  (R2 = 0.25, p = 0.007).
20           In all five studies, the personal PM was measured to be higher than either the indoor or
21      the outdoor PM measurements.  This relationship has also  been found in the two PTEAM
22      studies (Perritt et al., 1991; Clayton et al., 1993) described in detail later in Section 7.3.3.
23      For these PTEAM studies during the day (7 a.m. to 7 p.m.) average personal PM10 exposure
24      data (150 pig/m3) were 57% higher than the average indoor and outdoor concentrations,
25      which were virtually equal  (95 /ig/m3).  At night (7 p.m. to  7 a.m.) average PM10 personal
26      exposures (77 /xg/m3) were higher than the average indoor concentrations (63 /ig/m3) but
27      lower than the average outdoor concentration (86 ptg/m3).  Consequently, considering that the
28      PTEAM subjects were overwhelmingly mostly indoors at night, a time-weighted-average
29      (TWA) of the indoor and outdoor PM concentrations appears to always underestimate the
30      personal exposures to PM.
       April 1995                                7-19      DRAFT-DO NOT QUOTE OR CITE

-------
 1          It has been proposed (WHO, 1982a; Spengler et al., 1985; Mage, 1985) that such a
 2     discrepancy between the TWA and the personal monitoring measurements may be caused by
 3     two factors described as follows:  (1) Human exposure to PM at work and in traffic are only
 4     partially accounted for in a TWA of indoor and outdoor ambient PM values;  and (2) Indoor
 5     and outdoor averages reflect periods of low concentration during which the subject is not
 6     present.
 7          With regard to the first factor,  the PM of occupational exposures, and exposures in
 8     traffic  that also reflect the vehicular  emissions of PM plus the resuspension of roadway dust
 9     from the turbulence of the vehicles,  are not well represented by ambient PM  measurements.
10     Ambient PM monitors are usually sited several meters above ground level, at a location
11     uninfluenced by a single local source, so that the data recorded can represent an average
12     community type concentration (Mage,  1983).  The complexity of the situation is exemplified
13     by the  PTEAM study reported by Clayton et al. (1993). People who were employed had
14     lower daytime exposures to PM than nonemployed people (n = 59, 127 pig/m3 versus
15     n = 111, 162 jiig/m3), and people who had time in traffic had lower PM exposures than
16     those who stayed at home (n = 31,  97/ig/m3 versus n = 121, 163 /*g/m3). A possible
17     explanation may be that people who  stay home generate PM by doing housework and people
18     who go off to work may be driving  to a location where the indoor PM and ambient PM are
19     lower than at their home.
20          With regard to the second factor, the PM pollution generating activities in a home
21     usually occur only when a person is at home, as discussed in Section 7.1.2.  Therefore, the
22     PM in a home will be higher when a person is present than when the home is unoccupied.  A
23     24-h average of the indoor concentration thereby underestimates the average exposure of a
24     person while in that home.
25          Ambient PM is also higher during the day (when industry and traffic are active, and
26     wind speeds are high) than at night when PM generating activities are at a minimum and the
27     air is still. Consequently, a 24-h average ambient PM value generally underpredicts the
28     concentrations during the daylight hours and the exposures of people going outdoors during
29     that period.
30        Therefore, a 24-h TWA personal exposure will always tend to be underpredicted by a
31     simple TWA of 24-h residential-indoor and ambient-outdoor PM concentration data that fails

       April 1995                               7-20       DRAFT-DO NOT QUOTE OR CITE

-------
  1      to account for occupational exposures, in-traffic exposures, and the intensity of personal
  2      exposures during human activity that cannot be recorded by area monitors several meters
  3      away from the subject.
  4
  5      7.3.2  Personal Exposures in International Studies
  6           The Global Environment Monitoring System (GEMS) of the World Health
  7      Organization/United Nations Environment Programme (WHO/UNEP) commissioned a series
  8      of four pilot studies of personal exposure to PM in Zagreb (WHO, 1982a),  Toronto (WHO,
  9      1982b), Bombay (WHO, 1984) and Beijing (WHO, 1985). In these studies, people who
 10      worked in the participating scientific  institutes were recruited to carry a PM sampler and
 11      their exposures were matched to the ambient concentrations measured outside their home or
 12      in their communities. The results of these studies of PM,  expressed as mean personal
 13      exposure (PEM) and mean ambient (SAM), and the regression R2 between them are
 14      presented in Table 7-2.
 15          In addition to the institute personnel, the Toronto study also measured  exposures of
 16      asthmatics to PM25 and obtained R2 values of 0.07 and 0.00 in the summer  and winter
 17      respectively. The net result of these  international  studies is that they confirm the lack of a
 18      consistent relationship between individual personal PM exposures and ambient concentrations
 19      as found in  the U.S. studies.  The results of the Beijing study are of importance because of
 20      (a) the recent paper by Xu et al. (1994), which reports a correlation between mortality in
 21      Beijing and  ambient TSP and SO2 and (b) the very low correlations between personal
 22      exposure to PM and  ambient PM concentrations (not statistically different from zero) found
 23      in the WHO study (R2 = 0.03 and 0.07).  The major fuel in Beijing is coal, and in the
 24      winter the ambient PM27 (TSP) averages up to 500 )ug/m3  (WHO/UNEP,  1992).  During the
 25      GEMS  PM exposure study of 1985, the 24-h ambient concentrations of PM3 5  measured
 26      outside  the subjects homes averaged 420 jug/m3, the indoor concentrations in the homes
 27      averaged 364 /xg/m3, and the personal exposures averaged  191 /ug/m3.  The  subjects who
28      were workers at the institute conducting the study spent little time outdoors and their days
29      were spent at the institute.  The lower relative values of their exposure appear to be caused
30      by lower values at home during the evenings.  Meals are usually prepared during the day
31      when the workers are not there and the indoor exposures during cooking can be quite high.

        April 1995                              7_2i       DRAFT-DO NOT QUOTE  OR CITE

-------
             TABLE 7-2.  SUMMARY OF WHO/UNEP GEMS/HEAL PM, PERSONAL
                              EXPOSURE PILOT STUDY RESULTS
Location
Season PM /mi n
Toronto 25 13
winter
summer
Zagreb 5 12
summer
winter
Bombay 3.5 15
winter
summer
monsoon
Beijing 3.5 20
winter
summer
PEM Mean
m

72
78

12
12

105
102
101

71
40
Time
8-h


1-wk


24-h




24-h
1-wk
+ SE

122+9
124+4

114+?
187+?

127+6
67+3
58±3

177+?
66±?
SAM Mean R2 PEM
+ SE

68+9
78+4

55+?
193±?

117±5
65 ±3
51+2

421 ±?
192+?
vs SAM

0.15
0.10

0.00
0.50

0.26
0.20
0.02

0.07
0.03
P

NS
NS

NS
NR

NR
NR
NS

NS
NS
       n = number of subjects carrying PEM.
       m = total number of observations.
       NR = Not Reported, but listed as significant.
       NS = Not significantly different from 0.
       ? = Not reported.
       *25 pm A.D. computed from flow rate and open filter design.
 1     Smith et al. (1994) reported the cooking exposures to PM10 listed in Table 7-3 for Beijing,
 2     Bangkok and Pune.  In Beijing, personal exposures of the cooks during cooking were 4 to 20
 3     times higher than the 24-h outdoor values on the days that cooking took place. The presence
 4     of high levels of coarse particles in the  ambient air, which do not readily penetrate into the
 5     institute and the homes, may contribute to the significantly higher ambient values that are
 6     uncorrelated with the personal exposures.
 7
 8     7.3.3  The Particle TEAM (PTEAM) Study
 9          In 1986, the  U. S. Congress mandated that EPA's Office of Research and Development
10     "carry out a TEAM  Study of human exposure to particles."  EPA's Atmospheric Research
11     and Exposure Assessment Laboratory (AREAL) joined with California's Air Resources
12     Board (CARB) to  sponsor a study in the Los Angeles Basin.  The study was carried out
13     primarily by the Research Triangle Institute and the Harvard School of Public Health,  with
       April 1995
7-22
DRAFT-DO NOT QUOTE OR CITE

-------
         TABLE 7-3. SUMMARY OF DAILY INDOOR, OUTDOOR, AND PERSONAL
              EXPOSURES TO PM10 DURING COOKING AS A FUNCTION OF
                          FUEL TYPE IN THREE CITIES IN ASIA
City
Beijing, PRC




Pune, India




Bangkok Thailand


Cooking Fuel
Coal (vented)
Gas
LPG
Coal Gas
Natural Gas
Agric. Residue
Wood
Biomass
Kerosene
LPG
Charcoal
LPG near traffic
LPG far from traffic
Indoor
(/ig/m3)
550
400
370
420
410
2800
2000
2100
480
250
330
390
300
Outdoor
(/ig/m3)
550
430
410
440
440
2600
920
1000
340
250
330
450
285
Personal
Cooking
Only (^g/m3)
1900
5000
3300
9100
1600
900
1100
1100
530
420
550
850
3900
      Source: Smith et al. (1994).
1     additional support from Lawrence Berkeley Laboratory, Acurex, and AREAL.  The main
2     goal of the study was to estimate the frequency distribution of exposures to particles for
3     nonsmoking Riverside residents. Another goal was to determine particle concentrations in
4     the participants' homes and immediately outside the homes.
5     7.3.3.1 Pilot Study
6     7.3.3.1.1  Study design
1         A pilot study was undertaken in nine homes in Azusa, CA in March of 1989 to test the
8     sampling equipment (Ozkaynak, et al., 1990). Newly-designed personal exposure monitors
9     (PEMs) were equipped with inhalable (PM10) and fine (PM2 5) particle inlets.  The PEMs
      April 1995
7-23
DRAFT-DO NOT QUOTE OR CITE

-------
 1     were impactors with 4-Lpm Casella pumps (Wiener, 1988).  Two persons in each household
 2     wore the PEMs for two consecutive 12-h periods (night and day). Each day they alternated
 3     inlet nozzles.  The first five households were monitored concurrently for seven consecutive
 4     days (March 6 to 13, 1989); the last four households were then monitored concurrently for
 5     four consecutive days (March 16 to 20, 1989).  This resulted in approximately 100 PEM
 6     samples for each size fraction.
 7          Indoor and outdoor particle concentrations were monitored using microenvironmental
 8     exposure monitors (MEMs).   These monitors were the  Harvard "black boxes" (Wiener,
 9     1989) employing a 10 Lpm pump.  Several indoor MEMs were placed in different rooms  in
10     each house to determine the magnitude of room-to-room variation.  These monitors were
11     capable of monitoring both fine and inhalable particles  simultaneously.
12          A central site with a PEM, MEM, and two EPA  reference methods (dichotomous
13     samplers and high-volume samplers with a 10 /xm size-selective inlet) was also operated
14     throughout the 11 days (22 12-h periods) of the study.

15     7.3.3.1.2 Results
16          Side-by-side comparisons indicated good agreement of all four monitors (Table 7-4).
17     Good agreement was also  noted between outdoor concentrations at the homes and at the
18     central site (Wiener et al., 1990). Room-to-room variation of  particle levels was generally
19     less than 10%. Therefore the several indoor MEM values in a particular house were
20     averaged to provide a single mean indoor value to compare to  the corresponding outdoor
21     value (Table 7-5).  It was  decided that this finding would justify using only one indoor
22     monitor in the subsequent  full-scale study (Clayton et al., 1991).
23          The personal exposures  were about twice as great as the  indoor or outdoor
24     concentrations for both PM10 (Table 7-6a) and PM2 5 (Table 7-6b).  Considerable effort was
25     expended to determine whether this was a sampling artifact, due for example to the constant
26     motion of the  sampler; however, no evidence could be found for an artifactual effect.
27     Nonetheless, to reduce chances for an artifactual finding in the main study, it was decided to
28     use identical PEMs for both the personal and fixed (indoor-outdoor) samples in the  main
29     study.
        April 1995                               7-24       DRAFT-DO NOT QUOTE OR CITE

-------
                TABLE 7-4.  PTEAM PREPILOT STUDY: MEAN OUTDOOR
                           PARTICLE CONCENTRATIONS (/tg/m3)
Central Site PM-10

Mean
SD


Mean
SD
MEM
56.9
28.8

MEM
39.4
23.3
PEM
59.1
25.2
Central
PEM
46.0
23.7
DICOT HIVOL
61.7 56.6
27.6 31.5
Site PM-2.5
DICOT
41.8
20.6
Mean
58.4
28.8
SD
5.7
3.3
Residential
MEM
61.4
26.0
SD
7.4
4.4
Residential
Mean
41.4
22.9
SD
6.2
4.6
MEM
41.7
21.6
SD
7.1
5.8
      Each sampler collected 22 12-h samples over 11 days.
      MEM: Microenvironmental monitor: 10 Lpm impactor.
      PEM: Personal exposure monitor: 4 Lpm impactor.
      DICOT: Dichotomous sampler: 16.67 Lpm virtual impactor.
      HIVOL: High-volume (1130 Lpm) impactor.
            TABLE 7-5.  PTEAM PREPILOT STUDY: TWENTY FOUR-HOUR (24-h)
                           PARTICLE CONCENTRATIONS Gig/m3)


Mean
SD
SE

Indoor
58.7
24.6
3.4
PM-10
Outdoor
62.6
24.9
3.5

Indoor
36.3
18.6
2.6
PM-2.5
Outdoor
42.6
21.6
3.0
1
2
3
4
5
6
7
     Regressions of outdoor on indoor concentrations showed low R2 values (1 to 30%) for
both PM10 and PM2 5 size fractions, as did regressions of daytime indoor on personal
concentrations (R2  = 0 to 18%). Overnight indoor concentrations had somewhat better
ability to explain personal exposures (R2 = 14 to 58%), as might be expected from the fact
that the personal monitor was placed on the bedside table during the sleeping period.
Personal exposures were essentially uncorrelated with outdoor concentrations (R2 = 0 to 2%)

(Ozkaynaket al., 1993).
      April 1995
                                      7-25
DRAFT-DO NOT QUOTE OR CITE

-------
 TABLE 7-6a. PTEAM PREPILOT STUDY:  TWENTY FOUR HOUR (24-h) PM-10
                     CONCENTRATIONS (/*g/m3)
House
1
1
1
1
2
2
2
2
3
3
3
3
4
4
4
5
5
5
6
6
7
7
8
8
9
9
Mean
SD
SE
Day
1
3
5
7
1
3
5
7
1
3
5
7
2
4
6
2
4
6
8
10
9
11
9
11
8
10



Person 1
102
142
158
92
109
99
131
62
98
100
143
76
109
90
99
80
70
80
130
150
209
80
135
97
136
273
117.2
44.9
8.8
Person 2
86
125
150
127
158
140
87
56
107
141
132
103
92
77
122
104
77
78
152
102
126
71
178
151
102
91
112.9
30.8
6.0
Indoors
54
38
49
34
122
37
41
32
86
39
71
36
77
34
36
76
62
54
114
106
46
29
73
38
63
121
60.3
28.5
5.6
Outdoors
132
49
70
49
112
48
70
46
115
45
79
44
102
47
37
99
65
50
39
51
72
39
59
28
43
48
63.0
27.1
5.3
April 1995
7-26
DRAFT-DO NOT QUOTE OR CITE

-------
       TABLE 7-6b. PTEAM PREPILOT STUDY:  TWENTY-FOUR HOUR (24-h) PM-2.5
                                CONCENTRATIONS (/tg/m3)
House
1
1
1
2
2
2
3
3
3
4
4
4
4
5
5
5
5
6
6
7
7
8
8
9
9
Mean
SD
SE
Day
2
4
6
2
4
6
2
4
6
1
3
5
7
1
3
5
7
9
11
8
10
8
10
9
11



Person 1
44
55
55
58
46
51
53
62
109
75
46
118
40
65
59
40
34
71
77
64
111
53
110
178
105
71.2
32.7
6.5
Person 2
96
88
382
53
100
50
66
94
88
61
43
94
40
69
70
56
53
81
75
135
67
100
1453
48
58
140.8
275.5
55.1
Indoors
22
25
21
31
27
28
48
30
39
33
19
31
17
62
35
42
25
56
53
17
32
27
35
70
42
34.7
13.7
2.7
Outdoors
67
39
33
52
43
40
58
35
39
71
29
46
26
96
38
55
28
33
18
27
35
27
35
40
28
41.6
16.8
3.4
1     7.3.3.2 Main Study
2     7.3.3.2.1  Study design
3         A three-stage probability sampling procedure was adopted (Pellizzari et al., 1993a).
4     Ultimately 178 residents of Riverside, CA took part in the study in the fall of 1990.
      April 1995                             7-27      DRAFT-DO NOT QUOTE OR CITE

-------
 1      Respondents represented 139,000 ± 16,000 (S.E.) nonsmoking Riverside residents aged ten
 2      and above.  Their homes represented about 60,000 Riverside homes.
 3          Each participant wore the PEM for two consecutive 12-h periods.  Concurrent PM10
 4      and PM2 5 samples were collected by the stationary indoor monitor (SIM) and stationary
 5      ambient monitor (SAM) at each home.  The SIM and SAM were identical to the PEM except
 6      for the pump,  which was a Medo pump operated off line current.  A total of ten particle
 7      samples were collected for each household (day and night samples from the PEM10, SIM10,
 8      SIM2 5, SAM10, and SAM2 5). Air exchange rates were also determined for each 12-h
 9      period.  Participants were asked to note activities that might involve exposures to increased
10      particle levels  (nearby smoking, cooking, gardening, etc.).  Following each of the two 12-h
11      monitoring periods, they answered an interviewer-administered questionnaire  concerning their
12      activities and locations during that time.
13          Up to four participants per day could be monitored, requiring 48 days in the field.   A
14      central outdoor site was maintained over the entire period (September 22, 1990 through
15      November 9, 1990). The site had two high-volume samplers (Wedding & Assoc.) with
16      lO-pim inlets (actual cutpoint about 9.0 /mi), two dichotomous PM10 and PM2 5 samplers
17      (Sierra-Andersen) (actual cutpoint about 9.5 /*m), one PEM, one PM10 SAM, and one PM2 5
18      SAM.

19      7.3.3.2.2 Results
20          Of 632 permanent residences contacted, 443 (70%) completed the  screening interview.
21      Of these, 257 were asked to  participate and 178  (69%)  agreed.

22      7.3.3.2.3 Quality  of the Data
23          More than 2,750 particle samples were collected,  about 96% of those attempted. All
24      filters were analyzed by X-ray fluorescence (XRF) for a suite  of 40 metals.  More than
25      1,000 12-h average air exchange rate measurements were made. A complete discussion of
26      the quality of the data is found in Pellizzari et al.,  1993b, and in Thomas et al.,  1993. Blank
27      PEM and SIM/SAM filters (N = 51) taken to the field increased in mass by  an average of
28      9.5 pig; this value was subtracted from each field sample.  Limits of detection (LODs), based
29      on three times the standard deviation of the blanks, were on the order of 7 to 10 /ig/m3.   All

        April 1995                               7-28      DRAFT-DO NOT QUOTE OR CITE

-------
 1     field samples exceeded the LOD.  Duplicate samples (N = 363) showed excellent precision
 2     for all types of particle samplers at all locations, with median relative standard deviations
 3     ranging from 2 to 4%.
 4          The collocated particle samplers at the central site showed good agreement, with
 5     correlations ranging from 0.96 to 0.99. The central-site PEMs collected about 12% more
 6     mass than the dichotomous  samplers, which in turn collected about 7%  more than the
 7     modified Wedding hi-vol samplers.  These relative relationships had also been noted in the
 8     pre-pilot study in Azusa. It was noted that the Wedding samplers collected about the same
 9     mass as the dichotomous samplers during the day, but about 13% less mass at night.
10     Multivariate tests indicated  that the Wedding samplers appeared to have a temperature
11     dependency,  amounting to an increase of about 1% per °F (Ozkaynak et al., 1993, Appendix
12     A).

13     7.3.3.2.4  Concentrations
14          Concentrations of particles and elements have been reported (Clayton et al., 1993;
15     Ozkaynak et al., 1993; Pellizzari et al., 1993; Wallace et al., 1993).  Population-weighted
16     daytime personal PM10 concentrations averaged about 150 pig/m3, compared to concurrent
17     indoor and outdoor mean concentrations of about 95 ^g/m3 (Table 7-7). The overnight
18     personal PM10 mean was much lower (77 /*g/m3) and more similar to the indoor (63 jug/m3)
19     and outdoor (86 /xg/m3) means. About 25% of the population was estimated to have
20     exceeded the 24-h  National Ambient Air Quality Standard for PM10 of  150  £ig/m3.  Over
21     90% of the population exceeded the California Ambient Air Quality Standard of 50 /-ig/m3.

22     7.3.3.2.5  Correlations
23          The central site appeared to be a moderately good estimator of outdoor particle
24     concentrations throughout the city.   Spearman correlations of the central-site concentrations
25     measured by all three methods (PEM-SAM, dichot, Wedding)  with outdoor near-home
26     concentrations as measured by the SAMs ranged from 0.8 to 0.85 (p< 0.00001). Linear
27     regressions indicated that the central-site 12-h readings could explain 57% of the variance
28     observed in the near-home 12-h outdoor concentrations (Figure 7-3).
       April 1995                               7-29       DRAFT-DO NOT QUOTE OR CITE

-------
              TABLE 7-7.  POPULATION-WEIGHTED3 CONCENTRATIONS AND
                                 STANDARD ERRORS (/tg/m3)
Sample type
Daytime PM10
Personal
Indoor
Outdoor
Overnight PM10
Personal
Indoor
Outdoor
Daytime PM2 s
Indoor
Outdoor
Overnight PM2 5
Indoor
Outdoor
N
171
169
165
168
163
162
173
167
166
161
Geom.
Mean
129
78
83
68
53
74
35
38
27
37
GSD
1.75
1.88
1.68
1.64
1.78
1.74
2.25
2.07
2.21
2.23
Arith.
Mean + SE
150 + 9
95 + 6
94 + 6
77+4
63 + 3
87 + 4
48+4
49 + 3
36 + 2
51+4
Percentile
90% + SE
260 + 12
180 + 11
160 + 7
140 + 10
120 + 5
170 + 5
100 + 7
100 + 6
83+6
120 + 5
98%
380
240
240
190
160
210
170
170
120
160
      aPersonal samples weighted to represent nonsmoking population of 139,000 Riverside residents aged 10 or
       above.  Indoor-outdoor samples weighted to represent 61,500 homes with at least one nonsmoker aged 10 or
       above.
1          Outdoor 12-h concentrations of PM10 could explain about 25 to 30% of the variance
2     observed in indoor concentrations of PM10, but only about 16% of the variance in 12-h
3     personal exposures to PM10 (Figure 7-4).  This is understandable in view of the importance
4     of indoor activities such as smoking, cooking, dusting, and vacuuming on exposures to
5     particles.  The higher daytime exposures were even less well represented by the outdoor
6     concentrations.
7          Indoor concentrations accounted for about half of the variance in personal exposures.
8     However, neither the indoor concentrations alone, nor the outdoor concentrations  alone,  nor
9     time-weighted averages of indoor and outdoor concentrations could do more than  explain
      April 1995
7-30
DRAFT-DO NOT QUOTE OR CITE

-------
          O)
             600
             500
          §  400
          <•*
          O
          E  300
          "2
          tC
          £
          (0
             200
                                                    Backyard = 1.03*Central + 17.6
                                                    #-0.57   N-323
                 0
50           100          150          200
Central site reference monitor mean (ng/m3)
                            250
       Figure 7-3.  Central-site mean of two dichotomous samplers versus residential outdoor
                   monitors. R2 = 57%.
 1     about two-thirds of the observed variance in personal exposures.  The remaining portion of
 2     personal exposure is assumed to arise from personal activities or unmeasured
 3     microenvironments that are not well represented by fixed indoor or outdoor monitors.

 4     7.3.3.2.6  Discussion
 5          The more than 50% increase in daytime personal exposures compared to concurrent
 6     indoor or outdoor concentrations suggested that personal activities were important
 7     determinants of exposure.  However, the nature of this  "personal cloud" of particles has not
 8     yet been determined.  Scanning electron microscopy was undertaken on 138 personal filters
 9     (Mamane,  1992).  Skin flakes were common on many filters.  A preliminary analysis
10     suggested that the average number of skin flakes per filter was 120,000 to 150,000.  The
11     mass of a small number of personal filters may have been considerably increased by
12     unusually large numbers of skin flakes.  However, attempts to calculate the mass of skin
       April 1995
                 7-31
DRAFT-DO NOT QUOTE OR CITE

-------
              500

              400
           V)
           o
2>  300
           (0
           o
           8-
           75
           c
           
-------
                          20
                              40         60         80        100
                         Percent increase in personal cloud
       Figure 7-5. Increased concentrations of elements in the personal versus the indoor
                   samples.
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
fact that the participants were sleeping for much of the 12-h overnight monitoring period,
and were thus not engaging in these particle-generating or reentraining activities.
     A source apportionment of the personal PM-10 mass during the daytime period is shown
on Figure 7-6.  This chart is derived by subtracting the average SIM and SAM (95  /xg/m3)
from the  mean PEM (150 /ig/m3) given on Table 7-7.  The 55 /xg/m3 difference is shown as
the 37% fraction of the total of 150 ;ug/m3 labelled Personal 37%.  The source of this
'personal cloud' is indeterminable from the SIM, SAM and PEM data.  As discussed
previously, it is likely to consist primarily of resuspended dust that would have a composition
of a mixture of all the other sources. The 15% other-indoor PM represents the indoor mass
that could not be assigned to ETS, cooking or ambient PM.  It is likely that the 52% of
other-indoor plus personal-cloud categories contains an appreciable amount of ambient PM
that came indoors over a long period of time and is resuspended by activity.  If so,  then the
PEM would be at least 50% of ambient origin.
       April 1995
                                        7-33
DRAFT-DO NOT QUOTE OR CITE

-------
                      Outdoor
                       42%
                                                                  Personal
                                                                    37%
                                Smoking
                                  3%
                     N = 166 Samples
               Cooking
                 3%
     Other Indoor
        15%
       Figure 7-6.  Source apportionment of PTEAM PM-10 Personal Monitoring (PEM) Data.
                   'Other indoor' represents PM found by the indoor monitor (SIM) for which
                   the source is unknown.  'Personal PM' represents the excess PM captured
                   by the PEM which cannot be attributed to either indoor (SIM) or outdoor
                   (SAM).
       Source: Clayton et al. (1993).
 1          The PTEAM Study and the 13 PEM studies discussed in this chapter so far are
 2     summarized in Table 7-8.  This table shows that many of the early studies reported no
 3     statistically significant correlation between PEM and SAM.  However, these early  studies
 4     were all characterized by a non-probability sample and a relatively small sample size. The
 5     PTEAM study in Riverside was a probability sample (Clayton et al., 1993) and the Lioy et
 6     al. (1991) study  in Phillipsburg, which was not a probability sample, have large sample  sizes
 7     and achieved significance.  The other studies, such as WHO/UNEP (1982a,b) or Morandi et
 8     al. (1988) are equivocal.  Consequently it is not clear yet what the general pattern is and
 9     why.  In the following sections, PEM/SAM comparisons of some constituents and two means
10     of visualizing the complex relationships of SAM and PEM are discussed.
       April 1995
7-34
DRAFT-DO NOT QUOTE OR CITE

-------
I
VO
Lf»













71
&



w
£>
H
1
0
O
t— ^
2;
O
H
O
O
H
W
TABLE 7-8. COMPARISON OF PEM EXPOSURE OF INDIVIDUALS TO THE SIMULTANEOUS AMBIENT (SAM)
PM CONCENTRATION IN TEN U.S. CITIES AND FOUR FOREIGN CITIES
Reference
Binder et al.
Dockery & Spengler
Dockery & Spengler
Spengler et al.
WHO/UNEP




Spengler et al.

WHO/UNEP


Sexton & Spengler
WHO/UNEP



WHO/UNEP



Morandi et al.
Lioy et al

Perritt et al.

Clayton et al.


Year
1973
1975
1976
1979
1981
Winter
Winter
Summer
Summer
1981

1982
Summer
Winter
1982
1982
Winter
Summer
Monsoon
1985
Winter
Summer

1988
1988

1989

1990


Location
Ansonia
Watertown
Steubenville
Topeka
Toronto
non-asthmatic
non-asthmatic
asthmatic
asthmatic
Kingston/
Harriman
Zagreb


Waterbury
Bombay



Beijing



Houston
Phillipsburg

Azusa

Riverside


PM /xm
5
3.5
3.5
3.5
25




3.5

5


3.5
3.5



3.5



3.5
10

2.5
10
10


n
20
18
19
46

13
13
13
13
97

12


48
15



20



30
14#
14*
9
9
141


Time
24-h
24-h
12-h
12-h

8-h
8-h
8-h
8-h
24-h

1-wk


24-h
24-h




24-h
1-wk

12-h
24-h
24-h
24-h
24-h
24-h


Mean PEM
115
35
57
30

122
124
91
124
44


114
IS7
36

127
67
58

177
66

27
86
76
79
115
113


Mean SAM
59
n
64
13

68
7S
54
80
18


55
193
17

m
65
51

421
192

16
60
60
43
62
84


R2 PEM vs SAM
NS
0.00
0.19
0.04

0.15
0.10
0.00
0.07
0.00


0.00
0.50
0.00

0.26
0.20
0.02

0.07
0.03

0.34
0.04
0.25
0.01
0.01
0.23


P
NS
NS
NR
NS

NS
NS
NS
NS
NS


NS
NR
NS

NR
NR
NS

.09
NS

<.05
.008
.001
NS
NS
NR


n
H
w
n =  Number of individuals carrying personal monitors
NS = Not statistically significant from 0
NR = p Value  not reported, but mentioned as significant
# =  14 subjects carried PEMS for 14 days for 191 valid measurements
* =  Three outliers are removed and regression is for 188 measurements
Year = year study was performed

-------
 1     7.3.4  Personal Exposures to Constituents of PM
 2          The absence of an apparent correlation between ambient PM concentrations and
 3     personal exposures to PM found in most early US and international studies does not
 4     necessarily hold for specific chemical constituents of the PM that have predominantly outdoor
 5     sources.  Suh et al. (1993) measured personal exposures to sulfate (SO4=) and acidity (H+),
 6     and ambient and indoor concentrations in State College,  PA, summer 1991.  The correlations
 7     between personal  and ambient values of sulfate and acidity were R2= 0.92 and
 8     0.38 respectively, which is in marked contrast to the R2  ~ 0 between earlier reported
 9     ambient PM and personal PM (Sexton et al., 1984, Spengler et al., 1985; Morandi et al.,
10     1988).
11          Figure 7-7 shows the consistent relation between ambient and personal sulfate
12     measurements  (slope = 0.78 ± 0.02), and Figure 7-8 shows the improvement in prediction
13     by using the TWA with a correction factor (estimated  personal sulfate = 0.885*TWA,
14     R2 = 0.95 with slope = 0.96 + 0.02). Personal acidity was also computed by the same
15     equation with a correction for personal ammonia (NH3) exposure that gave an R2  = 0.63.
16     As opposed to PM which has both indoor and outdoor sources, the sulfate and acidity are
17     virtually all of outdoor origin.  Consequently, the characteristics of the indoor environment,
18     such as air conditioning and ammonia  sources, modify the personal exposures indoors.
19          Another important consideration  in estimating personal exposures, from the indoor and
20     outdoor environmental measurements,  is that the chemical composition of the excess in
21     personal exposure compared to the TWA exposure calculation may be significantly different
22     than that predicted from the indoor and ambient data alone.  For example, the excess
23     personal PM found by Morandi et al. (1988) appeared to be related to in-traffic exposures
24     which  would have a different chemical composition compared to either the average ambient
25     or average indoor compositions.  Exposures to particles  in vehicles are quite variable,
26     perhaps as a function of traffic density, and do not constitute a simple microenvironment for
27     estimation purposes. For example, Roemmelt et al. (1993) reported an average in-bus
28     concentration of 570 Mg/m3 of TSP (open filter collection at 1 m3/h), with a minute value
29     peak-to-mean ratio of 3:1 using a MINIRAM monitor (« PM5), during an 8.5-h period on
30     16 March 1993 in downtown Munich.  Simultaneous  ambient TSP data were not reported.
        April 1995                               7-36       DRAFT-DO NOT QUOTE OR CITE

-------
     CO
     7S"
     ~o
            I
            CD
            Q_
               600
               500
               400
               300
         200
               100
                            100
                                200
300
400
                                       Outdoor (nmoles/rrr)
500
600
      Figure 7-7.  Personal versus outdoor SO4 .  Open circles represent children living in air
                  conditioned homes; the solid line is the 1:1 line.
1
2
3
4
5
6
1
8
9
     In addition to the two factors cited just above, a microscale 'personal cloud' can be
generated by the person's activities which complicates the exposure measurement process.
This effect is most important in occupational settings where personal exposures are not
readily comparable to weighted area sampling measurements.  For example,  Lehmann et al.
(1990) measured workers exposure to diesel engine exhaust by personal monitoring of PM10
with a range of 0.13 to 1.2 mg/m3, compared to an area estimate range of 0.02 to
0.80 mg/m3. MMWR (1988) reports the exposures of nurses and respiratory therapists  to
the aerosols of ribavirin during treatment of patients by ribavirin aerosols administered inside
an oxygen tent. Bedside area monitors averaged 317 /ng/m3 while personal exposures ranged
April 1995                               7.37      DRAFT-DO NOT QUOTE OR CITE

-------
               500
                                                                               500
                                       Measured (nmoles/nrr)
      Figure 7-8.  Estimated ("best fit" model) versus measured personal SO4 .  Model
                  includes indoor and outdoor concentration and activity data.  Open circles
                  are air conditioned homes; the solid line is the 1:1 line.
1     from 69 to 316 /zg/m3 with an average of 161 /xg/m3.  The implications of these differences
2     in exposure with position relative to a source are discussed in previous Section 7.2.3.
3          Environmental Tobacco Smoke (ETS) is a category of PM that is found in many indoor
4     settings where smoking is taking place or recently had occurred.  As stated in the Indoor Air
5     Section 7.6, ETS is the major indoor source of PM where smoking occurs.  Because of the
6     depth of the discussion and coverage of ETS in Section 7.6, no further discussion will be
7     made here other than to note that ETS adds  on the order of 25 - 30 /zg/m3 to 24-h average
      April 1995
7-38
DRAFT-DO NOT QUOTE OR CITE

-------
  1     personal exposures and residential indoor environments where smoking takes place
  2     (Holcomb, 1993; Spengler et al., 1985).
  3         ETS represents an appreciable contribution to the personal PM exposure of the
  4     nonsmoker which is uncorrelated to the ambient PM concentration.  In many of the PM PEM
  5     studies, smoking status and exposure to ETS was self-identified by the subjects in their
  6     questionnaires and diary responses.  In the absence  of an independent verification by a
  7     measure of nicotine or cotinine,  a subject could be misclassified (Leaderer and Hammond,
  8     1991).  For example, smoking of a non-tobacco product may not be reported.   In the
  9     USEPA Denver and Washington CO exposure studies (Wallace et al.,  1988) there were
 10     several subjects who had high breath CO but their diary and CO PEM record gave no
 11     evidence of an equivalent exposure to CO or methylene chloride (which metabolizes to CO).
 12     A biological measure  of cotinine or nicotine could indicate whether the excessive breath  CO
 13     was from nonreported smoking or endogenously produced.
 14          The random ETS increment will tend to reduce the correlation between PEM and SAM.
 15     If one were able to subtract out the ETS from the PEM PM data, the correlation of SAM
 16     with the non-ETS PEM PM might be improved (Dockery and Spengler, 1981).
 17          As stated as a caveat in the introductory section 7.1, the inhalation of main-stream
 18     tobacco smoke will be a major additive exposure to  PM for the smokers, which dwarfs the
 19     nonsmoker's PEM PM. Therefore the results presented so far apply only to nonsmokers,
 20     and a major proportion of the US population (e.g. smokers) has a total exposure to PM that
 21      is at least one  order of magnitude greater than that of the nonsmokers.
22      7.4  INDIRECT MEASURES OF EXPOSURE
23           The early air pollution literature related health to ambient particulate matter (TSP)
24      concentrations as a surrogate for personal exposures to PM.  Although this relationship has
25      been  shown to be highly questionable for specific individuals, it still is used in studies such
26      as Pengelly et al. (1987) who estimated TSP exposures of school children in Hamilton,
27      Ontario, by interpolation of ambient TSP concentrations  to the school locations.
28           The first usage of a time-weighted-average (TWA)  of environmental exposures to
29      estimate total human personal exposure to an air pollutant (Pb) was by Fugas et al. (1973).

        April 1995                               7.39      DRAFT-DO NOT QUOTE OR CITE

-------
 1     In theory, a human exposure to PM could be estimated by use of Equation 7-2 and
 2     knowledge of the average PM concentration in each microenvironment (/xE) that a person
 3     experiences and the duration of the exposure in each such ^iE (Duan, 1982; Mage,  1985).
 4     For a room with no source in operation, the whole room could be treated as a single /xE.
 5     However, when a PM source is in operation and gradients exist, that very same room may
 6     need to be described by multiple piEs.  These pEs could have dimensions of an order of a
 7     few centimeters close to the source, and dimensions of an order of several meters,  farther
 8     away from the source.
 9          Under research conditions, the complete spatial variation of a gaseous tracer
10     concentration in a test chamber can be mapped (Yost et al., 1994) and measurements at
11     multiple fixed points can be made (Baughman et al.,  1994).   Baughman et al., 1994,
12     proposed that a /xE could be considered to have a uniform concentration if the coefficient of
13     concentration variation within it was less than 10% (standard deviation/mean). In the
14     presence of a source of PM,  these data indicate that tens of such /iEs would have to be
15     defined when a subject is moving about in its immediate vicinity.  A simultaneous video
16     recording of an individual's exact position and activity, while measuring the instantaneous
17     concentration, can be used to visualize a concentration field (O'Brien et al., 1989; Gray
18     et al.,  1992) and could be used to measure the concentration field distortion that occurs from
19     the presence of the person.  These new techniques establish the variability of concentration
20     over small distances and their relations to human activity.  It is clearly impossible to use
21     these research techniques routinely in an exposure survey or to maintain multiple area
22     monitors throughout a normal setting of daily activity while recording human activity in their
23     vicinities.
24          Ogden et al.  (1993) compared exposures from personal sampling and static area
25     sampling data for cotton dust exposures.  The British cotton dust standard specifies static
26     sampling, because the 1960 dose-response study used to set the standard used static sampling
27     data to compute worker exposure and dosage.  Ogden et al. (1993) found median personal
28     exposures of 2.2 mg/m3 corresponding to a mean static background concentration of
29     0.5 mg/m3.  They concluded that "The presence of the body and its movement affect what a
30     personal sampler collects, so static comparisons cannot be used to infer anything about the
31     relationship of the (static) method with personal  sampling." Ingham and Yan (1994)

       April 1995                                7-40      DRAFT-DO NOT QUOTE OR CITE

-------
  1     confirmed this finding by modelling the human body as a cylinder and showing that unless
  2     the personal monitor length/diameter ratio was greater than four, the aspiration efficiency
  3     (the fraction of particles sampled that would be sampled in the absence of the body) could be
  4     greatly affected.
  5          Rodes et al., 1991, compared the literature relationships of personal exposure
  6     monitoring (PEM) to /*£ area monitoring (MEM) for PM, as shown in Figure 7-9, to which
  7     Ogden et al., 1993 is added as a single point.  The authors found that PEM/MEM ratios
  8     ranged from 3 to 10 in occupational settings, and from 1.2 to 3.3 in residential settings.
  9     These combined data show that approximately 50% of all measured PEM PM values are
 10     more than 100% greater than the estimated simultaneous MEM values using the TWA
 11     approach.  Their explanation points to this excess PM as due to the spatial gradient about
 12     indoor sources of PM which are usually well away from area monitors which thus fail to
 13     capture the high exposures individuals may get when in close proximity to a source.  They
 14     suggest that clothing lint and skin dander could only add, at most, a few percent  to the total
 15     PM mass collected by a personal exposure monitor.

 16     7.4.1  Personal Exposure Models Using Time-Weighted-Averages (TWA)  of
 17            Indoor and Outdoor Concentrations
 18         Several studies have used the relationship of Equation 7-2 to compute the time-
 19     weighted-average (TWA) exposure of subjects. The procedure calls for a time-activity diary
 20     to be kept  so that the time at-home, outdoors,  at-work, in-traffic, etc., can be defined.  By
 21      use of fiE monitoring data from the study itself (or literature  values of PM concentrations in
 22     similar /*Es) and concurrent ambient monitoring, one can predict the concentration that would
 23      be  measured  if the subject had carried a PEM.
 24          Because people in the USA spend 21 h indoors each day (Robinson and Nelson, 1995),
 25      the concentration in indoor /^Es is a most important quantity for usage within a TWA PM
 26      model.  The important articles on indoor air qualtity for PM have been reviewed extensively
 27      in Section 7.6.  Many of these articles, such as Quackenboss et al. (1991), estimated TWA
28      PM10 exposures from SIM and SAM measured at subject's homes without collecting
29      simultaneous  PEM for validation of the TWA model.
       April 1995                              7.4!       DRAFT-DO NOT QUOTE OR CITE

-------
    100
     10
  CO
  DC
  LJJ
  LLl
  Q_
    0.1
             5  1.0
30
         ° Stevens (1969)
         A Fletcher and Johnson
         O Parker etal. (1990)
         o Lioyetal. (1990)
         A EPA PTEAM data
         © Ogden et al.
50
      (1988)
70
90   95   98
                                      Data  median   og
                                            13.40    1.98
                                             1.78
              a
              A

              O
              o
              A
              ©
                                             5.70
                                             1.58
                                             1.98
                                             4.40
                    1.17
                    3.40
                    1.53
                    1.62
       2    5  10        30     50     70
                       Cumulative % less than
Figure 7-9. Personal activity cloud and exposure.
                       90  95    98
Source: Rodes et al. 1991
April 1995
    7-42
     DRAFT-DO NOT QUOTE OR CITE

-------
  1           The articles that are discussed here predict PM exposures of non-smokers that include
  2      ETS, and most provide PEM data for comparison.  This may also be interpreted as the
  3      exposure of smokers minus their exposure to direct main-stream smoke and incompletely
  4      diluted side-stream smoke.
  5            As opposed to the gaseous pollutants for which continuous hour-to-hour time series of
  6      SAM data are available, PM SAM monitoring data have been often only available  as a time
  7      series of 24-h SAM measurements.  Consequently, in much of the early PM TWA
  8      literature, the modelers assumed, by necessity, the same ambient PM in the morning and
  9      evening, which might not be accurate (Dockery and Spengler, 1981).
10           Spengler et al.  (1980) in a study of PEM, SAM and SIM in Topeka,  Kansas, found the
11      averages of PEM =  30 Mg/m3, SIM =  24 Aig/m3 and SAM = 13 /*g/m3.  They note "It
12      suggests that somewhere in an individual's daily activities, they are being exposed  to PM at
13      concentrations higher than what is measured either indoors or outdoors". This relationship
14      has been found in almost all other studies, such as PTEAM (Clayton et al., 1993) where
15      daytime PEM averaged 150 ^g/m3 and  SIM and SAM averaged just under  100 /ig/m3.
16      Spengler et al.(1985) measured 24-h PEM, SIM and SAM.  The  resulting relationship based
17      on Equation 1 was:   PEM =  17.7 /ig/m3 + 0.9 TWA.  The authors noted, in addition to the
18      previous suggestion,  that the excess of PEM over TWA may be due to an incorrect
19      assumption that the indoor and outdoor  are constant during the 24-h  sampling period.
20           Morandi et al., (1988) compared PM3 5 PEM with simultaneous SIM  and SAM data.
21      They found  that their TWA model overestimated PEM below 27 /zg/m3 SAM and
22      underestimated PEM above 27 /ig/m3 SAM.  The authors concluded that "this result indicates
23      that there were differences between PM mean concentrations in ^iEs  with similar
24      characteristics ... The implication for air pollution health effect studies is that, for
25      contaminants with significant indoor sources, PEM may be the only adequate measure of
26      exposure when using short-term averaging times".
27          Koutrakis et al. (1992), in a study discussed in Section 7.6 on Indoor  Air report that
28      their source-apportionment  mass-balance model predicts penetration from outdoors to
29      indoors of order 85-90% for Pb and sulfur compounds.  The authors claim  that:
       April 1995                               7.43       DRAFT-DO NOT QUOTE OR CITE

-------
 1           "We can satisfactorily predict indoor fine aerosol mass and elemental concentrations
 2           using the respective outdoor concentrations, source type  and usage, house volume and
 3           air exchange rate."
 4           The authors note that this may be a cost-effective approach to estimating peoples'
 5     exposure while indoors since the necessary ambient data may be available and the housing
 6     profile may be collected with a simple interview.  This technique could possibly correct the
 7     shortcomings noted above by Morandi et al.  (1988).
 8           Colome et al. (1992) measured indoor and outdoor PM-10 at homes of asthmatics in
 9     California.  Their personal monitoring data,  limited to three individuals, confirmed that
10     "some protection from higher outdoor concentration is afforded by shelter if smokers and
11     other particulate sources are not present". This observation may be important for estimating
12     the exposure of elderly and infirm people who are assumed to be the susceptible cohort.
13           Klepeis et al.  (1994) present an up-to-date TWA PM Model that uses, as an input, real-
14     time hourly PM SAM data and a mass balance equation to predict exposures of nonsmokers
15     in various indoor settings based on ambient PM data, presence of PM sources such as
16     smokers, and other variables relating to  air exchange rates.  The addition of the  additive
17     terms that allow for sources, such as cooking and presence of smokers adds to the TWA of
18     Equation 7-2, which in effect is a correction for the underprediction of the /*E concentration.
19
20           In summary, as  described by several authors, the PM PEM exposure of individuals who
21     are not smoke exposed has been shown to be higher than their corresponding TWA of SIM
22     and SAM.   The exact reason for this excess  in PM, sometimes called a 'personal cloud', is
23     not known (Rodes et  al., 1991). It has been thought to reflect the fact that the person's
24     presence itself can stir up loosely settled-dust by induced  air motion and vibration (Ogden,
25     1993).  Thatcher and  Layton (1994) gave an example where merely walking into a room
26     raised the total  suspended dust from 10 to 20 /*g/m3. A study by Litzistorf et al. (1985)  of
27     asbestos type fibers in a classroom showed how fibers (f) were stirred up when it was
28     occupied.  The levels rose from below the detectable level of  10000 f/m3 to 80000 f/m3
29     when occupied,  and they returned to below detectable levels within 1 h after the end of the
30     class. Millette and Hays (1994) present a detailed discussion of the general topic of
31     resuspended dust in their text on settled  asbestos dust.

       April 1995                                7-44       DRAFT-DO NOT QUOTE OR CITE

-------
 1          It may not be a proper procedure to use a 24-h average concentration in a physical
 2     setting, such as a kitchen, to estimate a person's exposure while in the kitchen.  As described
 3     previously in the discussion of the definition of a microenvironment, the same kitchen can
 4     constitute one or more /^Es depending on the source operation pattern. In many studies, such
 5     as Spengler et al. (1985), the SIM sampled the indoor residential setting for 24-h in phase
 6     with the  PEM.  The resulting average SIM will often understimate the person's exposure
 7     while they are at home and may contribute to the difference between a TWA  exposure and
 8     the PEM.
 9          In a similar manner, a person's workplace exposure may be more or less than that in
10     their home. In the PTEAM study (Clayton et al., 1993), there was a general decrease in
11     exposure for those who were employed outside their home.  However, employment in a
12      "dusty trade", such as welding, may increase their PM PEM.  Lioy et al.  (1990) give an
13     example of a subject with a hobby involving welding  which led to a 24-h PEM  reading of
14     971 jig/m3. The variables influencing the contribution to PEM PM from industrial
15     exposures have not been  discussed in this chapter because of their complexity.  The reader
16     should bear in mind that  application of a TWA model to a subject with such an exposure
17     may create a high TWA  estimate with a large uncertainty.
18          Another exposure category that is important for TWA  analysis is that within a vehicle
19     in transit (Rudolf, 1994). In California, people spend approximately 100 minutes per day in
20     or near a vehicle (Jenkins et al., 1992).  In vehicles people are exposed to auto exhaust, road
21     dust resuspended from vehicle turbulence, and PM generated within the vehicle as ETS or
22     exhaust leakage.  Roemelt et  al. (1993) reported a range of TSP in  an urban bus up to 1500
23     Aig/m3 with a mean of 570 /xg/m3 as measured with an optical monitor during an 8.5-h
24     daytime  period.
25          Indirect estimation of a  person's time-weighted-average (TWA) PM exposure may be a
26     cost-effective alternative  to direct PEM PM measurement.  Mage (1992) compared the
27     advantages and disadvantages of the TWA indirect method compared to the direct PEM
28     method.  The primary advantages of the indirect method are the low cost and low burden on
29     the subject, because it uses only a time-activity diary  and no PM PEM is required; the
30     disadvantage is the low accuracy. The primary advantage of the PEM PM method is that it
31     is a high accuracy direct  measurement; the disadvantage  is the high cost and high burden on

       April 1995                               7.45       DRAFT-DO NOT QUOTE OR CITE

-------
 1      the subject (see Section 7.2.1.4).  Mage (1992) proposed a combined study design in which
 2      direct measurements on a subset of subjects can be used to calibrate  the TWA estimates of
 3      the other subjects.  Duan and Mage (1995) present an expression for that includes:  the
 4      optimum fraction of subjects to carry the PEM as a function of the relative cost of the  PM
 5      PEM to the TWA PM estimate and the correlation coefficient between the PEM and the
 6      TWA.
 7      7.5  DISCUSSION
 8      7.5.1. Relation of Individual Exposures to Ambient Concentration
 9           The previous sections discussed the individual PM PEM vs PM SAM relationships of
10      14 studies listed in Table 7-8.  In many  of the PM studies there is no statistically significant
11      linear relationship between PEM and SAM, and in other studies the relationship  is positive
12      and statistically significant.  This section discusses these data in terms of understanding the
13      complex relationship between the SAM concentrations and the individual PEM exposures.  In
14      the following section, the relationship of the SAM to the mean PEM in the community
15      surrounding the SAM will be presented.
16           The principle of superposition is offered as a basis for visualization of the process
17      involved in creating a total exposure.  A linear system will exist for respirable-PM PEM
18      exposures if the expected PEM response to  a source emitting 2 mg/min of PM is exactly
19      twice the PEM response to that identical source emitting 1 mg/min of identical PM.  If
20      superposition applies, then  we can construct the total exposure by adding all the  increments
21      of exposures from the various source classes and activities that a subject performs on a given
22      day.
23           Let the SAM be representative of the macroscale ambient PM concentration in the
24      community as shown on Figure 7-10a. This is the exposure that would be measured for a
25      homeless person if they spent 24-h per day  outdoors near the SAM site.  Neglecting local
26      microscale variation (e.g. backyard barbecue or leaf burning), while people are outdoors they
27      are exposed to 100% of the SAM value  (Figure 7-10b).  Assume  that this exposure is also
28      the baseline PM for a location in traffic  which occurs outdoors. The increment produced by
29      the local traffic is considered later.

        April 1995                                7-46       DRAFT-DO NOT QUOTE OR CITE

-------
                 SAM
                     SAM
                    Outdoors
                 SAM



                   '  Indoors   "•
               6      12     18      24
                Time - Hours
               6      12     18     24
                Time - Hours
               6      12     18     24
                Time - Hours
                                                           SAM
                                                          Traffic
                                                         Increment
                                                          to SAM
                                      6      12     18     24
                                       Time - Hours
                                                    Occupational
                                                      Exposure
                                                      Increment
                                                        to SIM
                                                                       SAM
                                      6      12     18     24
                                       Time - Hours
                                                                        ETS
                                                                     Exposure
                                                                       SAM
                                     6       12     18     24
                                       Time - Hours
SAM
T / X
/ \
SIM
/ Indoors \
/ non-ETS
\ / non-SAM N x
X s | -
- T / -
—if r h—

£.
o>
c


..

°^E
1
20 -*•
Cigarettes
Smoked



;

SAM









- •

























































-


































*


___
6      12     18
  Time - Hours
                                    24
    6       12     18     24
      Time - Hours
Figure lOa-h. Components of personal exposure.
April 1995
                       7-47
DRAFT-DO NOT QUOTE OR CITE

-------
 1          While people are indoors, they are exposed to a variable fraction of time-lagged SAM
 2     PM.  This constitutes an amount of (1) the fresh PM which depends on recent SAM and the
 3     air exchange rate between indoors and outdoors, and the PM deposition sinks (filtration of
 4     recirculated air,  surfaces, etc.), and (2) PM from outdoor sources that had been deposited in
 5     the past but is resuspended due to human activity and air currents.   PTEAM (Clayton et al.,
 6     1993), as cited in Section 7.6, found that outdoor air was  the major source of indoor
 7     particles, accounting for 75% of the fine fraction (<2.5 pirn AD) and 67% of the coarse
 8     fraction (2.5 pm AD to 10 ^im AD) in indoor air. It is noted that these average fractions
 9     will be lower in communities with lower average SAM values.  Lewis (1991) reported an
10     apportionment of indoor air PM in 10  homes within a wood burning community in Boise,
11     Idaho. The results showed that 50% of the fine PM was of outdoor origin (SAM), and in 9
12     of 10 homes, 90% of the sulfur was from outdoors (one home had an anomolous sulfate
13     injection from a humidifier using tap water). This is consistent with indoor sources varying
14     independently  of the SAM in a stationary manner (constant mean and variance), so that the
15     relative contribution of indoor sources  to indoor exposures decreases as SAM increases.
16     Figure 7-10c represents the increment  to PEM from outdoor sources of SAM while the
17     subjects are indoors at home and at work.  The SAM value is  shown as the dotted line for
18     reference in this and all the following  Figures 7-10c and 7-10h.
19          While people are indoors, at home and at work, they are also exposed to PM emitted
20     by  indoor sources - other than ETS from passive smoking and specific occupational sources.
21     These  sources,  such as cooking, lint from clothing and furnishings, mold, insects,  etc.,
22     create PM that agglomerates and deposits as visible dust that can be continuously
23     resuspended, which constitutes an additional PEM increment.   Figure 7-10d shows  the
24     additive effect of this source.
25          In traffic, or near vehicles in a parking garage or parking lot, people are exposed to an
26     increment of PM over and above the SAM value for that location.  Figure 7-10e shows the
27     additive PM for this setting that would be added to Figure 7-1 Ob for the local vehicular
28     emissions.
29          In an indoor setting,  in the presence of a smoker or the wake of a smoker, a PEM will
30     record an increment of ETS  associated with the act of smoking.  Figure 7-1 Of shows the
31     added PM increment for this  source.

       April 1995                               7-48      DRAFT-DO NOT QUOTE OR CITE

-------
  1          At work in a 'dusty trade' such as welder, mechanic, or miner, there will be an
  2     increment of exposure associated with these occupational activities that generate PM.
  3     Figure 7-10g represents the additive PM for these activities which are assumed to take place
  4     'indoors'.
  5          Last, but not least, is the physical  act of smoking itself.  As described previously, the
  6     main stream smoke from a cigarette, cigar, or pipe, bypasses the PM monitor and is inhaled
  7     directly. The mass of PM inhaled from smoking one-pack-per-day of the lowest Tar 'king
  8     soft pack' cigarettes, rated as delivering "1 mg 'tar' per cigarette by FTC method" is 20 mg
  9     per day (Woman's Dav. 14 March, 1995).  If this were distributed into a nominal 20 m3 of
 10     air inhaled per day, it would be an additive increment on the order of 1 mg/m3 to a 24-h
 11     PEM reading. Other cigarettes advertised in the same popular magazine are rated at "16 mg
 12     'tar' per cigarette by FTC method".  Therefore one-pack-per-day smokers can have a PM
 13     exposure standard deviation that is much larger than the mean exposure to PM of non-
 14     smokers, simply from choice of brand.   Figure 7-10h represents the act of smoking as
 15     creating exposures represented by the vertical  spikes with an integral area >  1 mg-day/m3
 16     per day.
 17          For all subjects, by the principle of superposition, the sum of the areas in Figures 7-
 18     lOb and 7-10c represents the exposure of an individual to the PM constituents that are
 19     characterized by a SAM PM concentration. The additional exposure categories that are
 20     independent of the SAM concentration (Figures 7-10d through 7-10g) and are appropriate for
 21      that subject would represent the portion of 24-h PEM PM that is not associated with SAM.
 22      Variance of SAM should explain much of the  variance in the SAM  related PEM fraction as
 23      defined by Figures 7-10b and 7-10c.  The summation over a full day for all categories 7-10b
 24      to 7-10g would be the PEM for any subject, such as is shown in Figure 7-2 (Repace and
 25      Lowery, 1980).
 26           Although there are no data for PEM PM exposures of individuals living in homes
 27      without any indoor sources of PM,  there are data for PEM sulfate as discussed previously in
 28      Section 7.3.4.  Given that there are negligible  sources of sulfur (S) that originate in the home
29      (matches, low-grade kerosine, humidifiers using tap water) the high correlation of PEM S
30      and SAM S (R2 = 0.92) of Figure 7-7 reported by Suh et al. (1993) where no appreciable
31      sources of S were present,  is an indication that the same relationship should hold for all

        April 1995                               7.49      DRAPT-DO NOT QUOTE OR CITE

-------
 1     SAM PM of that size range.  The data of Anuszewski et al. (1992) show that light scattering
 2     particles measured by nephelometry have very high correlation between indoor and outdoor
 3     concentrations (R2 >  0.9). Lewis (1991) and Cupitt et al. (1994) report that PM10 appears
 4     to penetrate with an average factor of 0.5 in Boise homes without woodburning.  The factor
 5     goes up to 0.7 with woodburning, and the authors assume that the factor would go up to 0.9
 6     in the summer when homes are less tightly sealed.
 7          If the variance of the PEM  PM portion which is uncorrelated to SAM (lOd to lOg) is
 8     very large, the percentage of the variance of the PEM PM that can be explained by the
 9     variance of SAM PM will be very small.
10          It may be possible that the 14 different populations sampled, cited in the 14 studies of
11     Table 7-8, have  widely different home characteristics, occupations, mode of commuting, and
12     smoking exposures that contribute to the different PEM vs SAM relationships. In some of
13     the cleaner communities, such as Watertown MA, Topeka KS, Waterbury VT, and Kingston
14     and Harriman TN, SAM averaged less than 20 jug/m3.   The non-SAM increments to PEM
15     exposure in these locales were greater than the SAM and may have been so variable that the
16     PEM PM became insignificantly correlated with the SAM PM data. The exception is
17     Houston, with a SAM = 16 /ng/m3 and a significant R2= 0.34 (0.005 < p < 0.05).
18     However, Morandi et al. (1988) note that deletion of 2 outlier observations would reduce R2
19     and make it nonsignificantly different from 0 (p > 0.2).  This is in contrast to the two large
20     studies in communities with high SAM levels (Clayton et al., 1993, Lioy et al, 1990),  where
21     the relations between PEM and SAM were significant.
22          All discussions above relate to nonsmokers. As for the smoker, the exposure from
23     Figure 7-10h would outweigh the  sum of all the other exposures, 7-10b through 7-10g. This
24     smoking increment may have an important implication for interpretation  of epidemiology
25     studies that relate a surrogate of PEM PM to mortality.
26          In the epidemiology studies of PM  and mortality of Chapter 12, the death counts are
27     usually culled to remove suicides and trauma victims.  This is because the SAM PM is not
28     considered to have been a possibly contributing cause to the accident, violence or voluntary
29     act that resulted  in death.  Consequently, smokers and nonsmokers alike are in the residual
30     mortality counts that are regressed against SAM and weather related variables. The
31     community SAM data are representative  of the exposure to PM of outdoor origin expressed

       April 1995                               7-50       DRAFT-DO NOT QUOTE OR CITE

-------
 1     by Figures 7-1 Ob and 7-10c.  For non-smokers this reflects about 50-75% of their
 2     exposure,  which is an appreciable amount. However, for one-pack-per-day smokers, if their
 3     smoking exposure is more than an order of magnitude higher, the SAM may reflect less than
 4     5.0 - 7.5% of their daily exposure to PM.

 5     7.5.2  Relation of Community Exposures to Ambient Concentrations
 6          Studies of the relationship between ambient PM and mortality/morbidity implicitly
 7     assume that the PM concentration at an ambient monitoring station (SAM) is a surrogate for
 8     the mean PM exposure of people  in the local community.  It can be shown that if
 9     individuals' probability of mortality  from PM exposure is linearly proportional to their PM
10     exposure,  then the expectation of  total PM-related mortality in the community is proportional
11     to the mean personal  exposure to  PM in the community.  Therefore, it may be appropriate to
12     ask the question, "how well does  SAM PM characterize the mean PEM PM in the
13     community?"
14          If all N people in a community carried a PEM, the mean PEM value is obtained by
15     summing all PM PEM values and dividing by N.  If indoor and outdoor PM are equally
16     toxic on a ^tg basis, then no further  information would be contained in PM SAM.  The mean
17     of a random sample of PEM PM measurements on subjects in the community would be an
18     unbiased estimator of the actual community mean PEM PM, and such a mean may be more
19     appropriate for use than a SAM measurement. Mage and Buckley (1995) tested the
20     relationship of the mean PEM PM exposure to the SAM in several locations, and their
21     results are given in the following  section.
22          Figures 7-1 la, 7-12a, 7-13a, and 7-14a show the individual personal  PEM  PM and the
23     corresponding ambient SAM PM from four (4) studies cited in Table 7-8 for which
24     individual data were available. For  example, Figure 7-1 la (Lioy et al, 1990) shows a set  of
25     PEM samples obtained from 14 nonsmoke exposed individuals on 14 consecutive days.
26     Because these 14 subjects were not selected as a probability sample  from the community of
27     Phillipsburg, NJ, we consider their exposures as a biased  sample from the  exposure
28     distribution that we would measure,  had every person in the community carried a PM PEM.
29     The outdoor average is estimated as  the mean of four (4) ambient SAM PM values obtained
30     on each day the people carried the PEM for PM10.

       April 1995                              7-51       DRAFT-DO NOT QUOTE OR CITE

-------
                                    Phillipsburg, NJ (Winter 1988)
                                       (all data included, n-191)
                          Figure a
                                                      Figure b
  1000

  •800
  i
o 600

            Q.
               400
            £. 200
                                                       200
                                                    §150
                                                    o
                                        5:100
                                        CO
                                        I
                                        0)
                                        °- 50
                                        I
              R2-0.333 (p-0.031)
              m - 0.478 (Mdwr- 0.20)
              b - 65.0 (Md«r- 29.3)
              n-14
                  0      50    100   150   200
                   Mean Outdoor PM-10, ug/m3 (4 Sites)
                                              0     50    100    150    200
                                                 Mean Outdoor, ug/m 3 (4 Sites)
       Figure 7-lla,b.  Personal exposure to PM in Phillipsburg, NJ (Winter, 1988)
       Source: Lioy et al. (1990).


 1          The data plot is a vertical profile corresponding to up-to-14 valid PEM values obtained
 2     on that day.  In the regression, each point is weighted equally,  and the R2 value of 0.037
 3     (p = 0.008) would be significant if the sample were unbiased.  The bias of a nonrandom
 4     sample has an expectation of zero and a finite variance if the choice of subjects is not based
 5     on factors related to exposure to PM.  In Figure 7-1 Ib, the mean of the daily PEM values is
 6     plotted.  The R2 now has increased from 0.037 to 0.333  (p=0.031).  The interpretation is
 7     that, on the average, variation in SAM only explains on the order of 4% of the variation hi
 8     an arbitrary individuals PEM,  but that same variation in SAM explains 33% of the variation
 9     in their mean exposure.
10          Because both these PEM and SAM values  are measured with error, an orthogonal
11     regression may be more appropriate,  with inverse variance weighting  for each of the PEM
       April 1995
                                    7-52
DRAFT-DO NOT QUOTE OR CITE

-------
            400
mE300
1?
o
^200
Q.
          I 100
              0
                                    Beijing, PRO (Winter 1985)
                        Figure a
                 R'- 0.064 (p-0.004)
                 m-0.14(*Merr-0.08)
                 b-116(std«T-37)
                 n-45
0  100200300400500600700
    Outdoor PM-10 (ug/m3)
                                     400
                                                 TO

                                                  ^3,
                                                  O
                                                     200
                                   CO
                                   o
                                   y>
                                   0>
                                  Q. 100
                                   I
                                       0
                                                 Figure b
                                                        0  100200300400500600700
                                                            Outdoor PM-10 (ug/m3)
       Figure 7-12a,b.  Personal exposure to PM in Beijing, China (Winter, 1985).
 1     and SAM means.  This would be expected to change the regression slightly but not the
 2     general conclusion that there is much more influence of SAM variation on the variation in
 3     the mean community exposure.
 4          In Beijing (Figure 7-12a,b) with a nonprobability sample of 20, the slope remains
 5     relatively constant at 0.14 as the R2 value increases from 0.064 to 0.23 with the usage of
 6     mean PM PEM exposure.
 7          Figure 7-13a,b for Azusa, California, with a nonprobability sample of 9, the correlation
 8     between PM PEM and PM SAM is negative (-0.01) and R2 is 0.0001. In this case, taking the
 9     mean of personal exposures shows  no significant improvement in the R2 value.  Such a low
10     value of R2, as in several of the studies cited in Table 7-8, may be caused by several factors,
11     such as sampling error (too few observations), a biased (nonprobability) sample, very strong
12     indoor sources of PM, or commuting of the people during the day to locations with
13     significantly different ambient pollution than in their home community.
       April 1995                               7.53      DRAFT-DO NOT QUOTE OR CITE

-------
                                      AZUSA.CA (Spring 1989)
                         Figure a
                    Figure b
ouu

^ 250
rt
E
? 200

o
2 150
0.
1C
o 100
®
50

n

' • R2-0.00




, •

- •
'Jtf-Ii "
"& . f ' :
t •? 1 ! •


1,1,1,1,!.!.
ouu
^»,
eo
E 250
o>
n
o 200
i
S
5: 150
cO
O
! 100
0.
c
§ 50
2
0

R2- 0.00
n-11



-

_ _
^ •• '
'I. ' "
^

. I , [ , I , i i I , I ,

0 20 40 60 80 100 120 1'
                 0  20  40   60  80 100 120 140
                  Mean Outdoor PM-10 (ug/m3)
            Outdoor mean represents 2 to 4 sites
             Mean Outdoor PM-10 (ug/m3)
       Outdoor mean represents 2 to 4 sites
       Personal Mean represents 4 to 6 persons
       Figure 7-13a,b.  Personal exposure to PM in Azusa, CA (Spring, 1989).
       Source:  Perritt et al. (1991).
 1          In Figure 7-14a,b, Riverside, California, with a probability sample of 178 people, the
 2     R2 values are significant, and they improved by a factor of three, from 0.16 to 0.49, while
 3     the regression equation remained essentially the same.
 4          The Riverside CA study and the Phillipsburg NJ study differ in three important aspects.
 5     In Phillipsburg, NJ, 14 nonsmoke-exposed at-home people carried a PEM for 14 days; in
 6     Riverside, CA, each day up to four different people,  some of them who were smoke-exposed
 7     at-home, carried a PEM for one day over the 48-day study period.
 8          1.  The Phillipsburg NJ data have a (potentially) 2-fold more precise estimate of the
 9             mean given by 14 PEM measurements as compared to 4 in Riverside, CA
10             (1/V13  = 0.28,  < 1/V3 = 0.57).
       April 1995
7-54
DRAFT-DO NOT QUOTE OR CITE

-------
   300





i 250
o
.c


S 200



1

§ 150

CL
o  50
a.
                _a
                        jffO go  O    O

                 	'i*7:	°	"	lW	
                     o~ o o  «
                       50     100    150     200     250     300     350

                            Backyard Concentration (24 hours)
          I
          o
            250
            200
          S?

          1 150
            100
             50
LLJ




o



-------
 1          2.  The lack of smoke exposure to the subjects in NJ would reduce the variance of the
 2              PEM measurements, and
 3          3.  The use of a nonprobability sample in NJ (and all other studies in which
 4              nonprobability sampling occurs) limits the statistical applicability of the results.
 5          The improvement in the regression R2 value by taking the mean of the PEM PM data is
 6     not an important finding of itself.  This improvement in the regression coefficient is
 7     predictable from the Central Limit Theorem and the process of regression to the mean of the
 8     observations - as when random measurement errors are removed.  The higher correlation of
 9     categorical exposure assignments has been noted in epidemiological studies.
10            The value of the improvement of the mean PEM relationship to SAM is that it
11     provides a better visualization that helps in understanding how mean PEM varies with SAM.
12     It thus provides a measure of the validity of the use of a daily PM SAM as a surrogate for
13     the mean PM PEM in the community.  It is clear that the uncertainty  in predicting mean
14     personal exposure  PM is  much smaller than the uncertainty in predicting the personal
15     exposure PM for an individual when we note that the means have a much smaller variability
16     about the line as shown in Figures 7-1 Ib, 7-12b, 7-13b,  and 7-14b.
17          There appears to be two distinct categories of exposure studies that are examined:
18          In the first type of study, such as Lioy et al. (1990) and Clayton et al. (1993), there is
19     a significant R2 between individual PM PEM and PM SAM.  In this category, there is an
20     appreciable improvement in correlation between the mean PEM and SAM. It has been
21     suggested that these cases with high correlation of PEM PM and SAM PM may arise where
22     the fine portion of the ambient PM (PM2 5) is highly variable from day-to-day,  and the
23     ambient coarse fraction is relatively constant.  In some locations, the fine  portion of the
24     ambient PM (PM-2.5) is  more variable from day-to-day than that of the ambient coarse
25     fraction.  In an urban area, the fine particle composition and the fine particle concentration
26     are highly correlated from site-to-site on any given day.  This is due,  in part, to the
27     homogeneous gas phase reactions of SOx and NOx to produce sulfates and nitrates, and
28     aerosol droplet formation with the condensation nuclei, such as metals, which are emitted
29     from ubiquitous sources,  such  as automobiles.
30          On the other hand,  ambient coarse particles are generated locally, and they have higher
31     deposition velocities than the fine particles. Their impact may then be limited by fallout to a
       April 1995                               7-56       DRAFT-DO NOT QUOTE OR CITE

-------
  1     locality downwind of their emission point, as they are not readily transported across an urban
  2     area.  Therefore,  during an air pollution episode, people living in an urban area may be
  3     exposed to fine PM of similar chemical composition and concentrations, whereas they will be
  4     exposed to coarse PM with a chemical composition that can depend on the location of the
  5     exposure.  Because PM2 5 could penetrate readily into a nonambient setting, the correlation
  6     between the  mean PM PEM and PM SAM would be high because all the people would have
  7     similar exposure to the ambient fine PM - plus exposure to indoor generated PM and ambient
  8     coarse PM which may have less fluctuation.
  9          In the second type of study, such as Sexton et al. (1982) and Spengler et al.  (1985),
 10     there is negligible correlation between individual PEM PM and SAM PM, and consequently
 11     there will  be little correlation between their mean PEM and the SAM.  In these  cases,  if the
 12     fine fraction is not an appreciable portion of the total PM, or there are  significant indoor
 13     sources, then the correlations between mean PM PEM and PM SAM may not be as
 14     impressive as for the other case.

 15     7.5.3  Implications for PM and Mortality Modeling
 16          PM related mortality may be specific to the most highly susceptible portion of the
 17     population.  Such a cohort may be the elderly people with the most serious chronic
 18     obstructive lung disease (COLD) and cardiac  insufficiency. Smithard (1954) relates the
 19     findings of Dr. Arthur Davies (Lewisham coroner) who  autopsied 44 people who died
 20     suddenly during the  1952 London Fog:
 21           "The great majority of deaths  occurred in people who had pre-existing heart and lung
 22         trouble, that is to say they were chronic bronchitic and emphysematous people with
 23          consequent commencing myocardial damage.  The  suddenness of the deaths,  Dr. Davies
 24          thought, was due to a combination of anoxia and myocardial degeneration resulting in
 25          acute right ventricular  dilation"
 26      Mage and  Buckley (1995) hypothesize that these people with compromised cardio-pulmonary
 27      systems may be relatively inactive, while selecting  to live in homes or institutional settings
28      without sources  of indoor pollution. If their time is spent in clean settings (e.g.  where
29      smoking is prohibited), then they would have  little  exposure to PM other than from the
30      ambient pollution that intrudes into their living quarters.  The  exposure  to PM of this cohort

        April 1995                               7.57       DRAFT-DO NOT QUOTE OR CITE

-------
 1     would be highly correlated with PM SAM, and so would be their mortality, if this PM was
 2     indeed highly reactive in their pulmonary tracts.  However, there have been no exposure
 3     studies done with people who correspond to the Lewisham mortality cohort.  Individual PM
 4     PEM of people outside this cohort, who could be relatively insensitive to PM, might not be
 5     significantly correlated with PM SAM,  as reported in most of the 14 studies cited in
 6     Table 7-8.  This suggests a model to relate PM and mortality as follows. Let any person (j)
 7     on a given day have a probability of mortality, p(m) = kj Xj, where kj is the unit probability
 8     of mortality per /zg/m3 of PM per day,  Xj is the daily average exposure to PM, jig/m3,
 9     independent of kj.  Let us assume that each individual (j) has their own personal value of kj
10     that can vary from day-to-day.
11          The expectation of total mortality  (M) in a community of size N can be shown to be the
12     summation of k X over all individuals (j =  1 to N) as follows:

13                                           M =  E kj Xj                                 (7-3)

14     If kj is independent of Xj,  then we can  define K as (1/N) E kj,  and the mean community
15     exposure X as (1/N) £ Xj, and it follows
16                                          M = N K X                                 (7-4)

17           This implies that, given a linear relationship of mortality with PM PEM exposure (X),
18     the expected mortality  is proportional to the mean community personal exposure to PM.  The
19     individual in the community, on any given day, with the highest probability of dying from a
20     PM  exposure  related condition is that individual with the highest product kj Xj, not
21     necessarily the highest exposed individual with the maximum value of Xj.
22           The Phillipsburg, NJ, data  set is a case in point. In this study, three subjects  had
23     excessively high PEM PM (shown by the three maxima on Figure 7-11 a).  These values
24     were caused by a hobby involving welding in a detached garage (971  /ig/m3), a home
25     remodeling activity (809 Mg/m3)  and usage of an unvented kerosine heater (453 /xg/m3).
26     Excessive PM generating activities are  not expected of elderly people who may have
27     compromised pulmonary systems.  In fact, the elderly and  infirm husband of the remodeler
28     had a personal exposure of 45 /ug/m3 on the day of the remodeling activity.  The indoor
        April 1995                               7-58      DRAFT-DO  NOT QUOTE OR CITE

-------
 1     monitors in the homes of the welder and remodeler only recorded 55 jug/m3 and 19 /-ig/m3,
 2     respectively, during those events, indicating the specificity of the high exposure to only the
 3     individual involved.  If we remove these three  'outliers', as being unrepresentative of the
 4     magnitude of exposures of those nonsmoke-exposed people most at risk from high values of
 5     kj, as defined  by Smithard (1954), then as shown on Figures 7-15a,b the correlation R2
 6     improves markedly, from 0.250 to 0.914.
 7          It is this relation of the PM PEM exposure to PM SAM concentration, as shown in
 8     Figures 7-15a,b that may be a better representation of the true situation underlying the PM vs
 9     mortality relationships because of the "healthy  worker"  effect.  Chronically ill  people who
10     are sensitive to PM might change their behavior to minimize their exposure to  irritants.
11     Consequently, healthy people with high PEM PM measures in occupations and indoor
12     settings can cause the regression R2 between PEM and SAM to be low, but they may not be
13     the individuals at highest risk of the acute effects of PM exposure.

14     7.5.4  Relative Toxicity of Ambient PM and Indoor PM
15          In the previous sections the SAM PM was evaluated as a predictor of PEM PM on the
16     implied basis that the health effects of PM were only mass dependent, and independent of
17     chemical composition.  It was shown in Table  7-8 that many early PM studies  of PEM had a
18     low  correlation between PEM and SAM on an individual basis that was often not
19     significantly different from zero.  But,  in the later studies (Clayton et al., 1993; Lioy et al.,
20     1990), a significant relationship was observed between PEM and SAM.  Further analysis
21     showed that on a daily basis, SAM would appear to be  a good predictor of mean community
22     PEM from the results of the Riverside, CA, and Phillipsburg, NJ studies.
23          However, there may be a significant difference in toxicity of PM per unit mass, as a
24     function of source type and composition, such  that some of the  PM of indoor origin is less
25     toxic than the ambient PM.  If so, then the SAM might be a better choice of surrogate  for
26     the toxicity of an individual's exposure than the PEM which may be  influenced by less toxic
27     materials.  There is some indication that on a unit mass basis, combustion products of fossil
28     fuels (coal  and oil) may be more acutely toxic  to the pulmonary system than combustion
29     products of biomass origin (tobacco, wood).  Furthermore, soil constituents and other
       April 1995                               7.59       DRAFT-DO NOT QUOTE- OR CITE

-------
                                   Phillipsburg, NJ (Winter 1988)
                              Three 'outliers' removed (971,809, & 453)
                         Figure a
                  Figure b
            300
            250

          I 200
            150
          § 100
          2
          S.
             50
                0   50  100  150  200  250 300
               Mean Outdoor PM-10, ug/m3 (4 Sites)
      300
      250
   o 200
    i
   S
   | 150
    o
    ® 100

   1  50
 R!-0.914 (p-0.001)
 m-0.561 (stdwr-0.050)
 b - 38.6 (std err - 7.45)
 n-14
          0   50  100  150  200  250  300
         Mean Outdoor PM-10, ug/m3 (4 Sites)
      Figure 7-15a,b.  Personal exposure to PM in Phillipsburg, NJ with concentration
                       outliers removed.
      Source: Lioy et al. (1990).
1     nonanthropogenic materials (iron oxide, alumina, Mt. St. Helens volcanic ash) also appear to
2     be less toxic than combustion products in general. See Chapter 11 regarding comparative
3     toxicity aspects.
4                      In summary, there is evidence that not all PM constituents have the same
5     toxicity per unit mass.  These differences are due to differences in aerodynamic diameter and
6     chemical composition.  As shown on a Venn diagram [Figure 7-16, Mage (1985)], the
7     focusing of the description of a PM exposure increases the ability to estimate the potential
8     toxicity of the exposure.  In the sequential description given below, the uncertainty in the
9     toxicity of the mixture is decreased as more information is provided.
       April 1995
7-60
DRAFT-DO NOT QUOTE OR CITE

-------
       Figure 7-16. Venn diagram (Mage, 1985) showing focusing of information to more
                   completely specify toxicity of a given PM mixture:  (1) universe of all
                   possible mixtures of PM with concentration of 2 ftg/m3; (2) subuniverse of
                   all combinations of PM with concentration of 2 /tg/m3 in size interval 2.0
                   to 2.5 /-on; (3) subuniverse of all combinations of PM with concentration of
                   2 jig/m3 in size interval 2.0 to 2.5 fim AD with 50% of automotive origin
                   and 50% from indoor sources; and (4) subuniverse of all combinations  of
                   PM with concentration of 2 jig/m3 in size interval  2.0 to 2.5 /tm AD with
                   50% of automotive origin and 50% from indoor  sources; 25% Pb, 25%
                   BaP and 50% unspecified inorganic materials.
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
1.  2 />ig/m3 of TSP.

2.  2 /-tg/m3 of TSP in the size interval 2 to 2.5 /mi.

3.  2 /ig/m3 of TSP in the size interval 2 to 2.5 /mi, 50% of automotive origin and
   50% of indoor source origin.

4.  2 /ig/m3 of TSP in the size interval 2 to 2.5 /mi, 50% of automotive origin and
   50% of indoor source origin, 0.5 /zg/m3 of Pb, 0.5 /*g/m3 of BaP and 1 jtig/m3 of
   unspecified inorganic material.
       April 1995
                                  7-61
DRAFT-DO NOT QUOTE OR CITE

-------
 1          As applied to human exposure to PM, this concept of differential toxicity suggests that
 2     data collections might benefit by providing data that would allow the toxicity of a PM

 3     exposure to be evaluated in terms of information, in addition to the mass collected per unit

 4     volume.

 5

 6     7.5.5  Conclusions

 7          (1)  For any air pollutant, the total exposure of an individual consists of a variety of
 8               sequential exposures to a variety of microenvironments.  They are typically,
 9               outdoor, indoors at-home, at-work, in-traffic and many other  indoor
10               microenvironments.  The principle of superposition is a useful mechanism to
11               visualize the  summation process.
12
13          (2)  For any identified air pollutant, the ambient environment is one source of indoor
14               pollution due to air exchange and infiltration.  Whether the ambient is a significant
15               or dominant source of indoor pollution depends on the relative strength of indoor
16               sources and sinks.
17
18          (3)  For PM, studies have detected a 'personal cloud' related to the activities of an
19               individual which may generate significant levels of airborne PM  in his/her vicinity
20               which may not be picked up by an indoor PM monitor at a distance.
21
22          (4)  For PM, some studies have identified significant sources in the home, e.g. due  to
23               cooking and smoking.
24
25          (5)  For PM of size fractions that include coarse particles, some studies have identified
26               statistically significant relationships between personal exposures and other studies
27               have not, probably due to overwhelming effects of indoor sources, 'personal
28               clouds' and other individual activities.
29
30          (6)  For PM of a fine size fraction - such as sulfates, there seems to  be more of a
31               relationship between ambient concentration and personal exposure, than for
32               coarser PM,  perhaps because of the ability of fine PM to penetrate into indoor
33               settings.
34
35           (7)  For a  study population in which there is a detectable correlation  between personal
36               exposures and ambient concentrations, the ambient concentration can predict the
37               mean personal exposure with much less uncertainty than it can predict the personal
38               exposure of an individual.
39
40           (8)  For fine PM constituents, such as  sulfates, high correlations between ambient
41               concentration and personal exposures have been identified.
42
        April 1995                               7-62       DRAFT-DO NOT QUOTE OR CITE

-------
  1           (9)  For Riverside, CA, where 25% of the population was estimated to have personal
  2               exposures exceeding the 24-h National Ambient Air Quality Standard for PM10 of
  3               150 ^g/m3, at least 50% of this mass is of ambient origin.
  4
  5
  6      7.6  INDOOR CONCENTRATIONS AND  SOURCES OF PARTICULATE
  7           MATTER
  8      7.6.1  Introduction
  9           Although EPA regulates particles in outdoor air, not indoors, it is still important to
 10      consider indoor air.  For one thing, most people spend most of their time indoors.  The most
 11      recent nationwide study of time budgets  (Robinson and Nelson, 1995), based on interviews
 12      with 9,386 respondents in 1993 to  1994, indicates that U.S.  residents spend 87.2% of their
 13      time indoors,  7.2% in or near a vehicle, and only 5.6% outdoors (Figure 7-17).  Secondly,
 14      we need to understand how outdoor particles are affected as they cross building envelopes.
 15      For a home with no indoor sources, how much protection is offered against particles of
 16      various size ranges? How do parameters such as the volume of the house, the  air exchange
 17      rate, cleaning frequency and methods, and materials  in the home affect particle
 18      concentrations? Indoor air studies have grappled with these topics and have the potential to
 19      answer these and other important questions ultimately affecting the health of the general
 20      public.
 21          This section has two parts.  The first part deals with field studies of particles indoors
 22      and outdoors, concentrating particularly on large-scale surveys of many homes and buildings.
 23      Besides presenting the observed indoor and outdoor particle concentrations, contributions of
 24      these studies toward understanding  important parameters such as air exchange rates, source
 25      emission rates, and decay rates are also reported. This section will also discuss a few studies
 26      dealing with inorganic and organic  constituents  of particles (e.g., elements and  PAHs) as
 27      well as other important considerations such as mutagenicity and the role of house dust in
 28      exposure to  metals and pesticides.  However, each of these topics is an entire field of study
29      in itself, and can only be touched on in this section.
 30          The second part of the chapter deals with  indoor air quality models and the experiments
31      and chamber studies performed to validate them.  A crucial parameter for particle studies is
32      the decay rate on surfaces, and a series of recent studies that have given information on this

        April 1995                               7.53      DRAFT-DO NOT QUOTE OR CITE

-------
            69.22
                                                  2.03
                                          Resid. (Indoors)
                                          Resid. (Outdoor)
                                          I" Vehicle
                                          Near Vehicle
                                          Other Outdoor
                                          Office/Factory
                                          Mall or Store
                                          School/Public Bldg
                                          Bar/Restaurant
                                          Other Indoor
                  3.56
                                      2.05
5.57 1-67
       Figure 7-17.  Percentage of time spent in different microenvironments by U.S. residents.
       Source: Robinson and Nelson (1995).
 1     point are reviewed. Since major modeling efforts have been aimed specifically at cigarette
 2     smoking, a special section is devoted to these models.
 3          In keeping with EPA's regulatory responsibilities, we omit the many studies in
 4     industrial workplaces and the "dusty trades".  We also omit studies whose main focus is lead
 5     in indoor locations, since lead is a separate criteria pollutant and such studies are reviewed in
 6     the lead criteria document.  Finally, although particle concentrations indoors are of crucial
 7     importance in determining the impact of radon daughters on health (smokers are at much
 8     higher risk from radon than nonsmokers), we omit studies focused on radon due to the fact
 9     that a different  branch of EPA has regulatory authority over radon.
10
11     7.6.2 Concentrations of particles in  homes and buildings
12          At least six major reviews of field  studies of indoor particles have been published since
13     1980.  However, all of these  reviews, three of which were sponsored  by tobacco companies,
       April 1995
                   7-64
DRAFT-DO NOT QUOTE OR CITE

-------
  1     were concerned with particles mainly as they relate to environmental tobacco smoke (ETS).
  2     Sterling et al. (1982) reviewed studies of ETS byproducts.  The National Academy of
  3     Sciences (National Research Council, 1986) reviewed 16 ETS-related studies,  including 8 in
  4     residences and 5 in offices or buildings.  Repace (1987) reviewed 13 ETS-related studies,
  5     including three in residences and three in offices or hospitals.  Guerin et al. (1992) reported
  6     on 21 field studies of particles worldwide, including 10 in residences and 10 in offices and
  7     public buildings.  The U.S. Environmental Protection Agency (1992) reviewed 27 ETS-
  8     related studies (24 published since 1980) including 10 in residences and 5 in offices.
  9     Holcomb (1993) updated Sterling's review, including 41 studies published in the U.S. or
 10     Canada  since 1980, with 14 studies in homes and 20 in offices or public buildings.
 11          Since the last of these reviews, several important studies have been completed,
 12     including EPA's major probability-based PTEAM Study.  Other large studies were not
 13     included in the earlier reviews for unknown reasons.  And some  studies, such as the Harvard
 14     6-City study, have had very recent summaries of all the data, heretofore scattered in a
 15     number  of publications.  Therefore it is appropriate to provide  a  new review of all major
 16     studies at this time.
 17          Since the two environments where people spend the most time are home  (68 to 70%:
 18     Chapin 1974; Szalai et al., 1972; Robinson and Nelson,  1995)  and work or  school (17 to
 19     20%), we will summarize the studies in these environments in turn.
 20
 21     7.6.2.1   Concentrations in homes
 22          There have been three large-scale studies (greater than 150 homes) of airborne particles
 23     inside U.S. homes. In chronological order, these are:
 24         1.    The Harvard 6-City study, carried out by the Harvard School of Public Health
 25              beginning in 1979 and continuing through 1988, with measurements taken in
 26              1,273 homes;
 27
 28         2.    The New York State ERDA study, carried out by Research Triangle Institute in
29              433 homes in two New York State counties in  1986;
 30
31          3.    The EPA Particle TEAM (PTEAM) Study, carried out  by Research Triangle
32               Institute and Harvard University  School of Public Health in 178 homes in
33               Riverside, California in 1990.
       April 1995                               7.65       DRAFT-DO NOT QUOTE OR CITE

-------
 1          We shall discuss the findings of each in some detail, since these studies present the
 2     most complete investigations to date of indoor and outdoor concentrations of particles.
 3
 4     The Harvard 6-City Study
 5          The Harvard 6-city study is a prospective epidemiological study of the health effects of
 6     particles and sulfur oxides.  Focused on children, it has included pulmonary function
 7     measurements  on more than 20,000 persons in the 6 cities, chosen to represent low (Portage,
 8     WI and Topeka, KN), medium (Watertown, MA and Kingston-Harriman, TN), and high
 9     (St. Louis,  MO and Steubenville, OH) outdoor particle and sulfate concentrations.
10          The study took place in two measurement phases.  The first phase involved monitoring
11     about 10 homes in each city for respirable particles  (PM3 5).  The homes were measured
12     every sixth day (24-h samples) for one to two years. In the second phase, a larger sample of
13     200 to 300 homes was  selected from each city, with week-long PM2 5 samples collected both
14     indoors and outdoors.   Two weeks of sampling in summer and in winter were provided.
15     Ultimately  over 1,200 homes were monitored in this way.
16          Spengler et  al. (1981) described the first 5  years of the Harvard 6-city study,  During
17     that Phase I period, pulmonary function measurements  were administered to 9,000 adults and
18     11,000 children in grades 1 through 6.  A questionnaire asks  about living conditions, type of
19     fuel and heating systems, occupation, and smoking habits  of parents.  Homes  were selected
20     on a volunteer basis, so that no extrapolations to a wider community are warranted.  In each
21     home a 24-h sample (beginning at midnight) was collected every sixth day. The cyclone
22     sampler has a  cut point of about 3.5 pm at a flow rate  of 1.71 Lpm.  About 10 sites in each
23     city were kept in operation for two years.  The annual mean indoor and outdoor RSP
24     concentrations are provided in Figure 7-18.  As  can be seen,  the indoor concentrations
25     exceeded the outdoor levels in all cities except Steubenville, where the outdoor levels of
26     about 46 /xg/m3 slightly exceeded the indoor mean of about 43 /xg/m3.  The authors noted
27     that the  major source of indoor particles is cigarette smoke, and categorized their data  by
28     the number of smokers in the home (Table 7-9).
29          Dockery and Spengler (198la) provided additional data analysis drawn from the same
30     6-city  study but including data from 68 homes compared to the 55 reported on in Spengler
31     et al. (1981).  Annual (every sixth day) mean indoor PM3 5 concentrations (in |ug/m3) were

       April  1995                                7-66       DRAFT-DO NOT QUOTE OR CITE

-------
II IVI
140£

100


« 80
1
3.
tt 60
S


40


20

—


—

—
(376)/r





~



X
V*




L/W







MV
jUj



X
I ywwfcy
Q





(330)
(342)
•i\
J)




X




-










X









S.







>^








pnd
highest
(183)







X










^^M *V< V^VI
( ) No. of samples
B«— Highest site mean
4 — Composite overall mean
•*— Lowest site mean


(186)
^
^

(274)
(355) n |_
(f (|3)0 0
                            P T  K  W SL  S
      P  T  K W SL S
      Figure 7-18.  The annual mean concentration of respirable particles (MRP) for the
                  highest and lowest site from the network of indoor and outdoor monitors
                  in each city. (P-Portage, T-Topeka, K-Kingston/Harriman, W-
                  Watertown, SL-St. Louis, S-Steubenville). Overall composite mean and
                  the number of samples are also shown.
           TABLE 7-9. CONCENTRATIONS OF PARTICLES (PM2 5) IN HOMES OF
             CHILDREN PARTICIPATING IN THE HARVARD SIX-CITY STUDY
Location
Indoors
No smokers
One smoker
Two or more
Outdoors
No. of homes

35
15
5
55
No. of samples

1,186
494
153
1,676
Mean (SD) (Mg/nr*)

24.4(11.6)
36.5 (14.5)
70.4 (42.9)
21.1 (11.9)
      Source: Spengler et al. (1981).
1     20 and 23 in the two "clean" locations (Portage and Topeka); 31 and 36 in the two

2     "medium" locations (Watertown and Kingston-Harriman); and 39 and 47 in the two "dirty"

3     locations (Steubenville and St. Louis). Outdoor PM10 concentrations (/zg/m3) measured by
      April 1995
7-67
DRAFT-DO NOT QUOTE OR CITE

-------
 1     dichotomous samplers running every other day in the cities were 20.1 ± 0.6 (SE) in Portage,
 2     31.9 + 1.1 in Topeka, 35.4 + 1.2 in Kingston, 41.1 ± 1.0 in Harriman,  25.8  ± 0.7 in
 3     Watertown, 37.4 + 1.0 in St. Louis, and 56.6 ± 1.7 in Steubenville (Spengler  and
 4     Thurston, 1983).  Corresponding outdoor PM2 5 concentrations were 12.5  ± 0.4,
 5     12.9 + 0.4, 24.6 ± 0.8,  24.3 ± 0.7, 17.3 ± 0.5, 20.5 ± 0.5, and 36.1  ± 1.2.  A mass
 6     balance model allowed estimation of the impact of cigarette smoking on indoor particles.
 7     Long-term mean infiltration of outdoor PM3 5 was estimated to be 70% for homes without air
 8     conditioners, but only 30%  for homes with air conditioners.  An estimate of 0.88 ^g/m3 per
 9     cigarette (24-h average) was made for homes  without air conditioning, while in  homes with
10     air conditioning the estimate increased to 1.23 /ig/m3 per cigarette.  A residual amount of
11     15 /ig/m3 not explained by the model was attributed to indoor sources such as cooking,
12     vacuuming and dusting.
13          Letz et al. (1984) developed a model of personal exposure to particles based on
14     88 participants who wore  personal monitors in the Kingston-Harriman portion of the 6-city
15     study.  From the 1 to 2 years of indoor-outdoor data on 57 homes  in the 6 cities, they
16     developed an equation relating indoor particle concentrations to those measured  outdoors:
17                           Cin = 0.385 Cout + 29.4 (Smoking)  + 13.8.
18
19     Thus homes with smokers had a PM3 5 ETS component of 29.4 pig/m3. The residual of
20     13.8 /ig/m3 was assumed  to be due to other household activities.
21          Neas et al. (1995) presented summary results for the entire second phase of the 6-city
22     Study (1983 to 1988).  In Phase 2, a total of 1,237 homes containing white never-smoking
23     children 7 to  11 years old at enrollment completed three questionnaires and completed two
24     weeks  of summer and winter monitoring indoors  and outdoors for PM2 5 using the Harvard
25     impactor.  At the beginning of the indoor monitoring study, 55% of the children were
26     exposed to ETS in the home, and 32%  were exposed to two or more smokers.   Household
27     smoking status changed for  173  children, (13% of smoking households ceased to smoke,
28     15% of the nonsmoking households became smoking households.)  The annual (winter and
29     summer) household PM2 5 mean concentration for the 580 children living in consistently
30     smoking households was 48.5 + 1.4 (SE) /ig/m3 compared to 17.3  ± 0.5  /^g/m3 for the 470
31     children in consistently nonsmoking households (Figure 7-19).  Among the 614  exposed

       April 1995                              7-68      DRAFT-DO NOT QUOTE OR CITE

-------
c
1_
-a
                            Never and Former Smoking Households
                                     Mean  17.3|ig/m3
   60
                                            80     100     120
                            Changed Smoking Status Households
                                     Mean  25.6 jig/m3
                           40
   60
                                       80     100    120
                               Consistantly Smoking Households
                                     Mean  48.5 jig/m3
                    20     40     60       80     100
                   Respirable Paniculate Matter (|ig/m3)
                           120
Figure 7-19. Distribution of numbers of children living in households with varying
          respirable particulate matter (PM2.5) as a function of parental smoking
          status.

Source: Neas et al. (1995).
April 1995
7-69
                                    DRAFT-DO NOT QUOTE OR CITE

-------
 1     children for whom complete information on smoking consumption was available, 36% were
 2     exposed to less than one-half pack daily, 40% to between 1/2 and 1 pack daily, and 25% to
 3     > 1 pack daily.  The distribution of household concentrations for children in these smoking
 4     categories is shown in Figure 7-20. (The authors stated that the estimated number of
 5     cigarette packs smoked daily in the home was "highly predictive" of the annual average
 6     PM2 5 concentration, but did not provide the results of any tests for significance.
 7          Spengler et al. (1985) reported on the Kingston-Harriman (TN) portion of the 6-city
 8     study.  An initial study design meant to compare exposures among individuals categorized by
 9     residence (Kingston or Harriman), ETS exposure, and occupational status (office worker,
 1     blue collar, or nonworking) was abandoned due to poor response rates of 30%; the sample
 2     was filled out  with volunteers and  thus cannot be extrapolated to the population of the two
 3     towns. 101 participants took part, with 28 having cigarette smoke exposure at home.  Each
 4     participant had an indoor and personal monitor with cutpoints of 3.5 pm.  Each town had a
 5     centrally located outdoor dichotomous sampler providing two size fractions, with cutpoints of
 6     2.5 /im and 15 /j.m. Both towns had similar outdoor PM2 5 concentrations of 18 /*g/m3, so
 7     the values were pooled for subsequent analyses.  Indoor concentrations averaged 42 +
 8     2.6 (SE) /ig/m3.  Indoor values in homes with smoking averaged 74 ± 6.6 /ig/m3,  compared
 9     to 28 + 1.1 /xg/m3 in homes without smoking (p  < 0.0001).  No difference in exposure
10     between the unemployed and employed population was noted.
11          Lebret et al. (1987) reported on the Watertown MA portion of the  Harvard 6-city
12     study.  265 homes were monitored for two one-week periods.  Homes with smoking
13     averaged 54 pig/m3 (N  =  147 and 152 during weeks 1 and 2), while homes without smoking
14     averaged 21.6 /ig/m3 (N = 70 and 74). The effect of smoking one cigarette/day was
15     estimated at 0.8 /ig/m3  of PM2 5.
16          Spengler et al. (1987) reported on a new round of measurements in three communities
17     within the 6-city study. In each community, about 300 children are selected to take part in a
18     year-long diary and indoor air quality study.  Measurements of PM2 5  were taken indoors at
19     home for two  consecutive weeks in winter and again in summer.  The sampler was the
20     automated Harvard sampler, which collected an integrated sample for  the week except for the
21     8 a.m. to 4 p.m. weekday  period when the child was at school.  During this 40-h period,
22     samples were taken in one classroom  in each of the elementary schools involved.  The three

       April 1995                               7-70      DRAFT-DO NOT QUOTE OR CITE

-------
     120-
     100
 CO
  CD
  €
  CD
 DC
      80-
  CD
  to   60
  o
  "t
  ctf
  0.

  S   40-
      20 H
90th %tile
75th %tile
50th %tile
25th %tile
10th%tile
                 i           i
              Never    Changed
               and      Status
              Former
          i           i           i
          Consistantly Smoking
                                     Pack
                  1/2-1
                   Pack
                 Packs
Figure 7-20.     Distribution percentiles for annual average concentrations of indoor
              respirable particulate matter (PM2 5) by household smoking status and
              estimated number of cigarette packs smoked in the home.

Source:  Neas et al. (1995).
April 1995
    7-71
DRAFT-DO NOT QUOTE OR CITE

-------
 1     communities were Watertown, MA, St. Louis, MO., and Kingston-Harriman, TN.  Results
 2     were presented for smoking and non-smoking homes in each city by season (Figure 7-21);
 3     the authors noted that mean concentrations in homes with smokers were about 30 fjLg/m"
 4     greater than homes without smokers.  The difference was greater in winter than in summer in
 5     all cities.
 6          Santanam et al. (1990) reported on a more recent and larger-scale monitoring effort in
 7     Steubenville and Portage as part of the Harvard 6-city study;  140 homes in each city were
 8     monitored for one week in summer and in winter.  The Harvard impactor sampler was used
 9     with an automatic time unit to collect  PM2 5 samples between 4 p.m. and 8 a.m. on
10     weekdays and all day on weekends, corresponding to likely times of occupancy for school-
11     age children.  Outdoor samples were collected from one site in each city.  Elements were
12     determined by XRF. A source apportionment using principal components analysis (PCA)
13     and linear regressions  on the elemental data was carried out.  Cigarette smoking was the
14     single largest source, accounting for 20 to 27 /xg/m3 indoor PM2 5 in Steubenville and 10 to
15     25 /xg/m3 in Portage (Table 7-10); unfortunately, the authors  do not state the number of
16     homes in the smoking and nonsmoking categories.  Wood smoke was  estimated to account
17     for about 4 /xg/m3 indoors and outdoors in Steubenville in winter, but only for about 1 /-ig/m3
18     indoors and outdoors in Portage. Sulfur-related sources accounted for 8 to 9 jig/m3 indoors
19     and 16 /ig/m3 outdoors in Steubenville in the summer, but were apparently not important in
20     winter.  Auto-related sources accounted for 2 to 5 /xg/m3 in the two cities. Surprisingly,  soil
21     sources accounted for  only about 1 to 3 jwg/m3 indoor and outdoor PM2 5 concentrations.
22     Nonsmoking homes in both cities had indoor mean PM2 5 concentrations very close to the
23     outdoor mean concentrations (ratios of 1.00 and 1.04 in Steubenville,  1.02 and 1.4 in
24     Portage). Homes with smokers exceeded outdoor levels by 25 and 20 pg/m3 in Steubenville,
25     and 24 and 11 /xg/m3 in Portage.
26
27     The New York  State ERDA Study
28          Sheldon et al. (1989) studied PM2 5 and other pollutants in 433 homes in two New
29     York State counties. One goal of the study was to determine the effect of kerosene heaters,
30     gas stoves, wood stoves or fireplaces, and cigarette smoking  on indoor concentrations of
        April 1995                               7-72      DRAFT-DO NOT QUOTE OR CITE

-------
iou -
150-
140-
130-
120-
~ 110-
| 100-
3 90-
"? 80-
2 70-
Q- 60-
50-
40-
30-
20-
10-
0J
S - smoking
N - non-smoking
- - - outdoors



i

h I


1



U


•
T





S N S N S





1
"^
N












1



1-

B


^





I



[•
S N S N





•
1





S N
Winter Summer Winter Summer Winter Summer
Watertown
St. Louis Kingston
       Figure 7-21.  PM2-5 (ng/m3) in smoking (S) and non-smoking (N) homes in three of the
                     Harvard Six-City Study sites.
       Source: Spongier et al. (1987).
 1     combustion products.  A stratified design to include all 16 combinations of the four
 2     combustion sources was implemented, requiring about 22,000 telephone calls.
 3          The sampler was a portable dual-nozzle impactor developed at Harvard University.
 4     Two oiled impactor plates in series were employed to reduce the probability that some
 5     particles larger than 2.5 ^m would reach the filter. Samples were collected in the main living
 6     area and in one other  room (containing a combustion source if possible) using a solenoid
 7     switch to collect alternate 15-min samples over a 7-day period.  Outdoor samples were
 8     collected at a subset of 57 homes.  All samples were collected during the winter (January to
 9     April)  of 1986.
10          PM2 5 mean concentrations indoors were approximately double those outdoors in both
11     counties (Table 7-11).  Of the four combustion sources, only smoking created significantly
12     higher indoor PM2 5 concentrations in both counties (Table 7-12).  Use of kerosene heaters
13     was associated with significantly higher concentrations in Suffolk (N  = 22) but not in
       April 1995
7-73
DRAFT-DO NOT QUOTE OR CITE

-------
           TABLE 7-10a.  RECONSTRUCTED SOURCE CONTRIBUTIONS TO PM2 s
                                 MASS FOR STEUBENVILLE


Source

Soil
Wood smoke
O.C.-I
Tobacco Smoke
Sulfur-related
Auto-related
O.C.-II
Indoor dust
Unexplained
Total
TABLE



Source

Sulfur-related
Auto-related
Soil
Tobacco Smoke
Wood smoke
Unexplained
Total

Smokers'
homes

7.9 (3.45)
9.5 (4.15)
10.3 (4.47)
45.6 (19.9)
NA
NA
NA
NA
26.7(11.6)
100 (43.57)
WINTER
Non-smokers
homes

17.6 (3.45)
21.2(4.15)
22.9 (4.47)
NA
NA
NA
NA
NA
38.3 (7.47)
100 (19.54)

Outdoor
site

9.6(1.79)
23.0(4.31)
24.8 (4.65)
NA
NA
NA
NA
NA
42.6 (7.95)
100 (18.7)
7-10b. RECONSTRUCTED SOURCE


Smokers'
homes

13.2 (4.56)
5.1 (1.78)
3.8(1.31)
71.0(24.6)
2.7 (0.94)
4.2(1.38)
100 (34.6)
MASS
WINTER
Non-smokers
homes

30.7 (4.56)
12.0(1.78)
8.8(1.31)
NA
6.3 (0.94)
42.2 (6.23)
100 (14.8)

Smokers'
homes

NA
NA
NA
53.7 (26.8)
17.8 (8.90)
7.3 (3.65)
8.8 (4.40)
7.4 (3.70)
5.0 (2.4)
100 (49.85)
SUMMER
Non-
Smokers'
homes
NA
NA
NA
NA
33.3 (8.23)
14.8 (3.65)
16.5 (4.07)
15.0 (3.70)
20.4 (5.05)
100 (29.5)

Outdoor
site

NA
NA
NA
NA
52.5 (15.5)
5.3(1.55)
26.0 (7.67)
NA
16.2 (4.78)
100 (29.5)
CONTRIBUTIONS TO PM2 *
FOR PORTAGE

Outdoor
site

39.2 (4.04)
17.3 (1.78)
13.4(1.38)
NA
13.0(1.34)
17.1 (1.80)
100 (10.3)

Smokers'
homes

23.3 (5.80)
18.1 (4.50)
7.5 (1.86)
40.1 (9.99)
NA
11.0(2.75)
100 (24.9)
SUMMER
Non-
Smokers'
homes
38.1 (5.30)
29.6(4.12)
13.4(1.86)
NA
NA
18.9 (2.62)
100 (13.9)

Outdoor
site

45.8 (6.23)
35.6 (4.84)
16.5 (2.25)
NA
NA
2.10(0.28)
100 (13.6)
      All figures in % (fig m-3)
      O.C.-I:  Iron and steel, and auto-related sources.
      O.C.-II: Iron and steel, and soil sources.

      Source:  Santanna et al. (1990)
1     Onondaga (N = 13).  Use of wood stoves/fireplaces and gas stoves did not elevate indoor

2     concentrations in either county.
3          Leaderer et al. (1990) extended the analysis of these data by collapsing the gas stove
4     category, reducing the number of categories from 16 to 8 (Table 7-13). By inspection of the
      April 1995
7-74
DRAFT-DO NOT QUOTE OR CITE

-------
             TABLE 7-11. WEIGHTED SUMMARY STATISTICS BY COUNTY FOR
        RESPIRABLE SUSPENDED PARTICIPATE (PM2 5) CONCENTRATIONS (jig/m3)

Percent Detected
Sample Size
Population Estimate
Arithmetic Mean (/*g/m3)
Arithmetic Standard Error (ng/m
Geometric Mean (/ig/m3)
Geometric Standard Error
Minimum (/ig/m3)
Maximum (jtg/m3)
Percentiles
10th
16th
25th
50th (median)
75th
84th
90th
95th
99th
Main Living
Onondaga
98.9
224
94,654
36.7a
i3) 2.14
25. T
1.07
0.72
172

9.93
11.2
13.5
23.9
48.4
68.0
85.2
112
136
Area
Suffolk
99.6
209
286,580
46.4
2.77
35.9
1.06
2.18
284

13.8
16.8
18.9
33.6
62.8
76.6
89.4
112
155
Outdoors
Onondaga
100
37

16.8
1.00
15.8
1.06
6.32
28.4



12.8
15.1
20.5





Suffolk
100
20

21.8
4.54
18.6
1.11
12.0
106



13.6
16.7
22.3




       Significantly different between counties at 0.05 level.
       Source: Sheldon et al. 1989.
 1     table, it is clear that smoking was the single most powerful source of indoor fine particles,
 2     with geometric means ranging from 28.5 to 61.4 jug/m3, whereas the four nonsmoking
 3     categories ranged from 14.1 to 22.0 /ig/m3.
 4         Leaderer and Hammond (1991) continued their analysis of the New York State data by
 5     selecting a subset of 96 homes for which both nicotine and PM2 5 data were obtained. In the
 6     47 homes in which nicotine was detected (detection limit = 0.1 ^g/m3), the mean
 7     concentration of RSP was 44.1 (± 25.9 SD) /*g/m3 compared to 15.2 (± 7.4) /-ig/m3 in the
 8     49 homes where no nicotine was detected.  Thus homes with smoking had an increased
 9     weekly geometric mean PM2 5 concentration of about 29 /xg/m3.  Imperfect agreement with
10     reported smoking was observed, with nicotine being measured  in 13% of the residences that
11     reported no smoking, while nicotine was not detected in 28% of the residences that reported
       April 1995                              7-75      DRAFT-DO NOT QUOTE OR CITE

-------
     TABLE 7-12. WEIGHTED ANALYSIS OF VARIANCE OF RESPIRABLE
 SUSPENDED PARTICULATE (PM2 5) CONCENTRATIONS Gig/m3) IN THE MAIN
        LIVING AREA OF HOMES VERSUS SOURCE CLASSIFICATION
                        F Value
     Probability
             Coefficient
Onondaga
Model                  20.5
Independent Variables:
  Intercept
  Gas Stove              1.87
  Kerosene Heater         1.06
  Tobacco               81.6
  Wood Stove/Fireplace    2.42

                R2 = 0.17
     0.00
     0.17
     0.30
     0.00
     0.12
             20.3
              5.25
              5.05
             45.1
              7.81
Suffolk
Model                  36.9
Independent Variables:
  Intercept
  Gas Stove              0.13
  Kerosene Heater        12.0
  Tobacco              114
  Wood Stove/Fireplace     0.71

                R2 = 0.21
     0.00
     0.72
     0.00
     0.00
     0.40
             26.1
            -1.52
             30.1
             46.8
              9.88
Source: Sheldon et al. 1989.

   TABLE 7-13.  RESPIRABLE SUSPENDED PARTICULATE CONCENTRATION
                               Suffolk
                       Onondaga
Source
None
W
K
S
KW
SW
SK
SKW
Outdoor
N
30
15
7
61

29
23
6
19
Mean
17.3
18.1
22.0
49.3

38.0
61.4
30.3
16.9
Standard
1.7
1.6
1.6
1.8

1.8
2.0
1.4
1.3
N
45
16
4
80
4
31
4
4
36
Mean
14.1
19.1
21.2
36.5
19.7
33.9
35.3
28.5
15.8
Standard
1.7
1.7
1.0
2.4
1.5
2.2
1.5
1.6
1.5
Abbreviations: W = woodstove; K = kerosene heater; S = smoking.

Source: Leaderer et al. 1990.
April 1995
7-76
DRAFT-DO NOT QUOTE OR CITE

-------
 1     smoking.  A regression of PM2 5 on total number of cigarettes smoked during the week (T)
 2     gave the result:
 3                           PM2 5 = 17.7 + 0.322T (N = 96; R2 = 0.55)
 4
 5     For the homes with measured nicotine, the regression gave the result:
 6
 7                           PM2 5 = 24.8 + 0.272T (N = 47; R2 = 0.40)
 8
 9     Thus each cigarette produces about a 0.3 (±0.03) jug/m3 increase in the weekly mean PM2 5
10     concentration, equivalent to a 2.1  (+0.2) /xg/m3 increase in the daily concentration.
11           Koutrakis et al. (1992) also analyzed the New York State data, using a mass-balance
12     model to estimate PM2 5 and elemental source strengths for cigarettes, wood burning stoves,
13     and kerosene heaters.  Homes with cigar or pipe smoking and fireplace use were eliminated,
14     resulting in 178 indoor air samples.  PM2 5 source strength for smoking was estimated at
15     12.7  ± 0.8 (SE) mg/cigarette; PM2 5 source strengths could not be estimated for wood
16     burning or kerosene heater usage, but only 7 homes in each category were available for
17     analysis.  For a final category of all other residual indoor sources, a source strength of
18     1.16 mg/h was calculated.  For nonsource homes (N  = 49) the authors estimated that 60%
19     (9 Mg/m3) of the total PM2 5 mass was from outdoor sources, and 40% (6 ^g/m3) from
20     unidentified indoor  sources. For smoking homes, they estimated that 54% (26 /xg/m3) of the
21     PM2.5 mass was fr°m smoking, 30% (15 ^g/m3) from outdoor sources, and  16%  (8 /zg/m3)
22     from unidentified sources.  These authors also developed an elemental emissions profile for
23     cigarettes, woodburning,  and kerosene heaters.  For cigarettes, the elemental profile included
24     potassium (160 /xg/cig), chlorine (69 /ig/cig), and sulfur (65 /ng/cig), as well as  smaller
25     amounts of bromine, cadmium, vanadium, and zinc.  The woodburning profile included three
26     elements:  potassium (92 /xg/h),  silicon (44 /xg/h) and calcium (38 /xg/h).  The kerosene heater
27     profile included a major contribution from sulfur (1500 /xg/h) and fairly large inputs of
28     silicon (195 /xg/h) and potassium (164 /xg/h).  A drawback of the mass-balance model was an
29     inability to separately estimate the value of the penetration coefficient P and  the decay rate
30     k for particles and elements; Koutrakis et al. (1992) assumed a constant rate  of 0.36 h"1 for
31     k, and then solved for P.

       April 1995                               7.77       DRAFT-DO NOT QUOTE OR CITE

-------
 1     The EPA Particle TEAM (PTEAM) Study
 2          The EPA sponsored a study of personal, indoor, and outdoor concentrations of PM10
 3     particles,  and indoor and outdoor concentrations of PM2 5 particles in Riverside, CA
 4     (Pellizzari et al.,  1993a,b; Clayton et al., 1993; Thomas et al.,  1993).  The personal
 5     exposure results of this study are discussed in Section 7.3.3.  The main goal of the study was
 6     to estimate the frequency distribution of exposures to PM10 particles for all nonsmoking
 7     Riverside residents aged 10 and above.  178 households were selected using probability
 8     sampling to represent about 61,000 households throughout most of the City of Riverside.
 9     Homes were sampled at the rate of four per day between Sept. 22 and Nov. 9, 1990.  Each
10     home had two 12-h samples  for both size fractions. A central site operated throughout the
11     48 days of the study, producing 96 12-h samples collected by side-by-side reference samplers
12     (dichotomous samplers and modified hi-volume samplers) along with the low-flow (4 Lpm)
13     impactor designed for this study.  The impactors had very sharp cutpoints at 2.5 and 11 /*m.
14     A second filter treated with citric acid to collect nicotine  was placed behind the main Teflon
15     filter.
16          A subset of the homes  was monitored for PAHs (Sheldon et al.,  1992).  125 homes
17     were monitored indoors and  65 of those were monitored outdoors for two consecutive  12-h
18     periods using a 20 1pm pump with an XAD cartridge.
19          Precision of the three types of particle samplers at the central site was excellent,  with
20     median RSDs of about 4 to 5% (Wallace, 1991a).  The low-flow sampler was noted to
21     produce estimates about 12% greater than the dichotomous sampler, which in turn was about
22     7% greater than the hi-vol sampler (Wallace,  1991b).  Part of the difference may be due to
23     the different cutpoints, which are estimated to be 11 ptm for the new sampler, 9.5 for the
24     dichot, and 9.0 for the hi-vol. Part of the difference may also be due to particle bounce
25     (large particles bouncing off the impactor and being reentrained in the flow to the filter),
26     such that  the PM2 5 and PM10 fractions in the low-flow sampler may be contaminated with a
27     small number of larger-size particles; however, particle bounce  was found in laboratory tests
28     to account for less than 7% of the total mass.
29          The population-weighted distributions of personal, indoor, and outdoor particle
30     concentrations are provided in Table  7-14. PM10 mean concentrations (150 /*g/m3) were
31     more than 50% higher than either indoor or outdoor levels (95 pig/m3). Overnight mean

       April 1995                               7-78      DRAFT-DO NOT QUOTE OR CITE

-------
3.
O
TABLE 7-14. WEIGHTED DISTRIBUTIONS OF PERSONAL, INDOOR, AND
           OUTDOOR3 PARTICLE CONCENTRATIONS (/tg/m3)
CD DAYTIME
<2












4j
VO


O
§>
3
H
6

o
H


Sample Size
Minimum
Maximum
Mean
(Std. Error)
Geometric Mean
(Std. Error)
Std. Deviation
Geometric Std. Deviation15
Percentiles
10th
25th
50th (median)
75th
90th
Std. Errors of Percentiles
10th
25th
50th
75th
90th

**Qtatictir»e r»tVn»r than tVi#> canrnlp ci-7P
PM
SAM
167
7.4
187.8
48.9
(3.5)
37.7
(2.5)
37.6
2.07

14.9
23.4
35.5
60.1
102.2

1.6
2.1
4.0
3.9
4.6

• minimum and ti
2.5
SIM
173
2.8
238.3
48.2
(4.1)
35.0
(3.3)
41.2
2.25

11.5
19.3
33.5
61.5
101.0

3.4
1.4
4.5
3.3
6.7

naYimiim arp i

SAM
165
16.2
506.6
94.9
(5.5)
82.7
(4.1)
57.2
1.68

42.8
56.9
84.1
110.8
157.2

2.3
4.5
4.7
4.0
7.2

ralr.nlatfirl nsin
PM10
SIM
169
16.6
512.8
94.7
(5.7)
78.2
(5-0)
61.4
1.88

30.9
49.5
81.7
127.2
180.7

3.4
4.3
8.3
9.4
11.0

IP weighted
NIGHTTIME
PM25
PEM
171
35.1
454.8
149.8
(9.2)
128.7
(8.5)
84.3
1.75

59.9
86.1
129.7
189.1
263.1

4.0
9.4
7.5
10.8
12.0

data: thev mi
SAM
161
3.4
164.2
50.5
(3.7)
37.2
(3.1)
40.3
2.23

14.5
23.0
35.0
64.9
120.7

2.1
2.7
2.4
4.6
5.8

wide estimate
SIM
166
2.9
133.3
36.2
(2.2)
26.7
(1.9)
29.5
2.21

10.0
14.8
25.9
48.9
82.7

0.9
1.3
2.4
5.3
5.8

s for the taree
SAM
162
13.6
222.9
86.3
(4.4)
74.5
(4.0)
47.7
1.74

39.3
53.6
74.1
103.7
167.8

7.4
3.4
4.8
5.1
4.3

:t population c
PM10
SIM
163
14.1
180.3
62.7
(3.2)
53.1
(3.1)
37.4
1.78

25.2
33.5
51.6
84.8
116.9

1.5
2.4
3.5
4.7
5.3

)f person-d

PEM
168
19.1
278.3
76.8
(3.5)
67.9
(3-D
39.7
1.64

36.6
48.1
66.2
98.8
135.0

1.5
3.1
4.3
8.2
10.1

ays
    (PEM) or of household-days (SIM, SAM).
    bln contrast to the other statistics, the gsd is a unitless quantity.
n
3

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14





















personal PM10 concentrations (77 /zg/m3) were similar to the indoor (63 /ig/m3) and outdoor
(86 Mg/ni3) levels. The reason for the higher daytime personal exposures is not completely
understood; it may be due
to the fact that the person is often close to the source of particles,
such as cooking, dusting, or vacuuming. It may also be due to re-entrainment of household
dust. It appears not to be
due to skin flakes or clothing fibers; many skin flakes were found
on filters but their mass does not appear to account for more than 10% of the excess personal
exposure (Mamane 1992).
Mean PM2 5 daytime

concentrations were similar indoors (48 /xg/m3) and outdoors
(49 Mg/m3), but indoor concentrations fell off during the sleeping period (36 /xg/rn3)
compared to 50 /ig/m3 outdoors. Thus the fine particle contribution to PM10 concentrations
averaged about 51% during the day and 58% at night both indoors and outdoors. The
distributions of these ratios are provided in Table 7-15.


TABLE 7-15



Sample Size
Mean
(Std. Error)
Geometric Mean
(Std. Error)
Percentiles
10th
25th
50th (median)
75th
90th
Std. Errors of Percentiles
10th
25th
50th
75th
90th


. WEIGHTED DISTRIBUTIONS3 OF PM2 5/PM10
CONCENTRATION RATIO
DAYTIME NIGHTTIME
Outdoor Indoor Outdoor Indoor
160 167 154 160
0.470 0.492 0.522 0.550
(0.016) (0.021) (0.017) (0.014)
0.444 0.455 0.497 0.517
(0.017) (0.022) (0.019) (0.016)

0.274 0.250 0.308 0.301
0.371 0.347 0.406 0.440
0.469 0.498 0.515 0.556
0.571 0.607 0.646 0.694
0.671 0.735 0.731 0.771

0.018 0.030 0.023 0.023
0.018 0.046 0.028 0.017
0.015 0.020 0.022 0.015
0.019 0.024 0.027 0.023
0.012 0.028 0.016 0.012
"Statistics other than sample size are calculated using weighted data; they provide estimates for the target
 population of household-days.
April 1995
7-80
DRAFT-DO NOT QUOTE OR CITE

-------
  1           Unweighted distributions are displayed in Figures 7-22 and 7-23 for 24-h average PM10
  2      and PM2 5 personal, indoor, and outdoor concentrations.  Most of the distributions were not
  3      significantly different from log-normal distributions, as determined by a chi-square test.
  4      About 25 % of the population of Riverside was estimated to have 24-h personal PM10
  5      exposures exceeding the 150 /xg/m3 24-h NAAQS for ambient air.
  6           The 48-day sequence of outdoor PM10 and PM2 5 concentrations is shown in
  7      Figure 7-24 (Wallace et al., 199la).  At least  two extended episodes of high fine-particle
  8      concentrations occurred.  Also about 6 days of high Santa Ana winds,  with correspondingly
  9      high coarse-particle concentrations from desert sand, were observed.
10           Central-site PM2 5 and PM10 concentrations agreed well with back yard  concentrations.
11      Pearson correlations of the log-transformed data were 0.96 and 0.92 for overnight and
12      daytime PM25, and 0.93 for the overnight PM10 values (Ozkaynak et al., 1993).  The
13      correlation dropped to 0.64 for the daytime PM10 values; however on this day two homes in
14      one part of Riverside showed very high outdoor concentrations of 380  and 500 jug/m3 while
15      two homes in another part of Riverside and the central-site monitor showed more typical
16      concentrations.  It is believed that a local event produced the higher concentrations at those
17      two homes. If they are removed from the data set, the correlation improves to 0.90. This
18      suggests that a single central-site monitor can represent well the PM2 5 and PM10
19      concentrations throughout a wider area such as a town or small city, at least in the Los
20      Angeles basin.
21           Daytime indoor PM10 and PM2 5 concentrations showed low-to-moderate Pearson
22      correlations of 0.46 and 0.55, respectively, with outdoor concentrations (N  =  158 to  173).
23      At night, the correlations improved somewhat  to 0.65 and 0.61, respectively (N  = 50 to
24      168).  Outdoor PM10 concentrations explained about 27%  of the variance of indoor levels
25      (Figure 7-25).
26           Simple regressions of outdoor on indoor  PM10 and PM2 5 resulted in the  following
27      equations:
28                Indoor PM10 =  51 + 0.49 x Outdoor PM10 (day)                 R2 = 0.20
29                Indoor PM10 =  20 + 0.51 x Outdoor PM10 (night)                R2 = 0.41
30                Indoor PM2 5 = 10 + 0.81 X Outdoor PM2 5 (day)                 R2 = 0.49
31                Indoor PM2 5 =  9 + 0.56 x Outdoor PM2 5  (night)                R2 = 0.55

        April 1995                               7-81      DRAFT-DO NOT QUOTE  OR CITE

-------
                     300
                      30
                         -  o
                                               -a- Personal
                                               -A- Indoor
                                                o  Outdoor
                                               J	L
                                                      J	L
                          300
                           25     50    75     90  95  98 99
                                  Cumulative Frequency (%)
                          30
Figure 7-22.  Cumulative frequency distribution of 24-h personal, indoor, and outdoor
             PM10 concentrations in Riverside, CA.
                     200
                  |
                                               o
                                         o
                      20'	*
                           25
                                                ~A~ Indoor
                                                 ° Outdoor
                                                i   i	i  i
                           200
50     75     90  95  98 99
Cumulative Frequency (%)
                           20
Figure 7-23.  Cumulative frequency distribution of 24-h indoor and outdoor PM2 5
             concentrations in Riverside, CA.
April 1995
       7-82
DRAFT-DO NOT QUOTE OR CITE

-------
           200
           150
          ,100
                 D Dichot coarse   —Dichot-10
                         20         40          60         80
                            12-hour period beginning Sept. 22,1990
                                100
Figure 7-24. Forty eight day sequence of PM10 and PM2.5 in Riverside, CA, PTEAM
            study.
         600
       ~ 500
       S 400
       8
         300
       8
       TJ
       ~ 200
       CVJ
         100

                                       Indoor - 0.54*Outdoor + 32
                                       R2-27% (n-309)
                     100       200        300       400       500
                         Average 12-h outdoor concentration (ng/m3)
                                 600
Figure 7-25.  Average indoor and outdoor 12-h concentrations of PM10 during the
             PTEAM study in Riverside, CA.
Source:  Ozkaynak et al. (1993).

April 1995
7-83
DRAFT-DO NOT QUOTE OR CITE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
     Stepwise regressions resulted in smoking, cooking, and either air exchange rates or
house volumes being added to outdoor concentrations as significant variables (Table 7-16).
Homes with smoking added about 27 to 32 /ig/m3 to the total PM2 5 concentrations and about
29 to 37 /ig/m3 to the PM10 values.  Cooking added 13 to 26 /ig/m3 to the daytime PM10
concentration and about 13 /ig/m3 to the daytime PM2 5 concentration, but was not significant
during the overnight period.   At night, air exchange added about 4.5 /ig/m3 to the PM2 5
concentration per unit increase (in air changes per hour) and about 12 /ig/m3 to the PM10
concentration, but was not significant during the day.  By contrast, the house volume was not
significant at night, but was significant during the day, with larger homes resulting in smaller
PM concentrations.
              TABLE 7-16. STEPWISE REGRESSION RESULTS FOR INDOOR AIR
             CONCENTRATIONS OF PM10, PM2 5, AND NICOTINE: COEFFICIENTS
                             (STANDARD ERRORS OF ESTIMATES)
PM10 PM25
Variable
N
R2
Intercept

Outdoor air

Smoking3

No. cigarettes15

Cooking0

Air exchange

House volume*1

All
310
41%


0.52
(0.05)
37
(6)
3.2
(0.7)
20
(5)
5.2
(2.0)
-0.08
(0.02)
Day
158
39%
57
(21)
0.66
(0.09)
29
(8)
3.0
(1.0)
26
(9)


-2.7
(1)
Night
147
58%


0.45
(0.05)
38
(11)
3.9
(0.9)
12
(5)
12
(5)


All
324
55%


0.64
(0.04)
28
(3.5)
2.5
(0-4)
9.4
(2.9)




Day
156
53%
21
(7.8)
0.71
(0.07)
27
(7)
2.4
(0.6)
13
(5)


-2.0
(0.6)
Night
149
71%


0.53
(0.04)
32
(10)
4.0
(0.6)


4.5
(2)


All
222
34%




1.1
(0.1)
0.11
(0.01)






Nicotine
Day
93
28%




1.1
(0.3)
0.1
(0.03)






Night
109
35%
0.28
(0.07)


1.0
(0.3)
0.2
(0.06)






       All listed coefficients significantly different from zero at p < 0.05.
       "Binary variable: 1 = at least 1 cigarette smoked in home during monitoring period.
       bThis variable was interchanged with the smoking variable in alternate regressions to avoid colinearity problems.
       cBinary variable: 1 = cooking reported for at least 1 min in home during monitoring period.
       d Volume in thousands of cubic feet.
        April 1995
                                         7-84
DRAFT-DO NOT QUOTE OR CITE

-------
 1

 2
 3

 4

 5

 6
21

22
     A model developed by Koutrakis et al. (1992) was solved using nonlinear least squares
to estimate penetration factors, decay rates, and source strengths for particles and elements
from both size fractions in the PTEAM study.  In this model, which assumes perfect

instantaneous mixing and steady-state conditions throughout each 12-h monitoring period, the

indoor concentration of particles or elements is given by
                                                 Qis/v
                                                a. + k
                                                                                         (7-5)
7
8
9
10
11
12
13
14
15
16
17
18
19
20

where
Qn
P
a.
COU
Qis
V
k

Fro
smoking
expressio

                   = indoor concentration (ng/m3 for elements, jug/m3 for particles)
                   = penetration coefficient
                   = air exchange rate (h'1)
                   = outdoor concentration (ng/m3 or jig/m3)
                   = mass flux generated by indoor sources (ng/h or jug/h)
                   = volume of room or house (m3)
                   = decay rate due to diffusion or sedimentation (h"1)
             From initial multivariate analyses, the most important indoor sources appeared to be
        smoking and cooking. Therefore the indoor source term Qls was replaced by the following
                                 Qis  = (NcigScig
                                                          'other
                                                                              (7-6)
where
23
24
25
26
27
28
29
30
31
t
Nc\g
sci

T
•*cook
c
°cook

Gother

                       duration of the monitoring period (h)
                       number of cigarettes smoked during monitoring period
                       mass of elements or particles generated per cigarette smoked (ng/cig or
                       Mg/cig)
                       time spent cooking (min) during monitoring period
                       mass of elements or particles generated per hour of cooking (ng/min or
                       /zg/min)
                       mass flux of elements or particles from all other indoor sources (ng/h or
        April 1995
                                         7-85
DRAFT-DO NOT QUOTE OR CITE

-------
 1          With these changes, the equation for the indoor concentration due to these indoor
 2     sources becomes
 3
                              _       t    NcigScig + TcookScook      Qother
                           in
                                 a+k           (a + k)Vt         (a  + k)V
 4
 5          The indoor and outdoor concentrations, number of cigarettes smoked, monitoring
 6     duration, time spent cooking, house volumes, and air exchange rates were all measured or
 7     recorded.  The penetration factor, decay rates, and source strengths for smoking, cooking,
 8     and all other indoor sources (2other) were estimated using a nonlinear model (NLIN in SAS
 9     software).  The Gauss-Newton approximation technique was chosen to regress the residuals
10     onto the partial derivatives of the model with respect to the unknown parameters until the
11     estimates converge.  On the first run, the penetration coefficients were allowed to "float"
12     (no requirement was made that they be < 1).  Since nearly all coefficients came out close to
13     one, a second run was made bounding them from above by one.  The NLIN program
14     provides statistical uncertainties (upper and lower 95 % confidence intervals) for all parameter
15     estimates.  However,  it should be noted that these uncertainties assume perfect measurements
16     and are therefore underestimates of the true uncertainties.
17          Results are presented in Table 7-17 for the combined day and night samples.
18     Penetration factors are very close to unity for nearly all particles  and elements. The
19     calculated decay rate for fine particles is 0.39 ± 0.16 h'1, and for PM10 is 0.65 + 0.28 h'1.
20     Since PM10 contains the PM2 5 fraction, a separate calculation was made for the coarse
21     particles (PM10 to PM2 5)  with a resulting decay rate of 1.0 h"1.  Each cigarette emits 22 +  8
22     mg of PM10 on average, about two-thirds of which (14 ± 4 mg)  is in the  fine fraction.
23     Cooking emits 4.1  ± 1.6  mg/min of inhalable particles, of which about 40% (1.7 + 0.6
24     mg/min) is in the fine fraction.  All elements emitted by cooking were limited almost
25     completely to the coarse fraction.  Sources other than cooking and smoking emit about 5.6 ±
26     3.1 mg/h of PM10,  of which only about 1.1 mg/h ±  1.0 (20%) is in the fine fraction.
27          Similar calculations were carried out for day and night samples separately, and for the
28     logarithms of the data as well as the untransformed data, to obtain more information on the
29     sensitivity of the estimates.  For PM10, the lowest estimate of the source strength of the
       April 1995                                7-86      DRAFT-DO NOT QUOTE OR CITE

-------
TABLE 7-17. PENETRATION FACTORS, DECAY RATES, AND SOURCE STRENGTHS: NONLINEAR ESTIMATES

                               Decay Rate (1/h)
S cook (fig/rain)
                                                                    S_smoke (/ig/cig)
i— VAR
tS PM2 5*
Al
Mn
Br
Pb
Ti
Cu
Sr
P
Si
Ca
Fe
K
S
Zn
Cl
PM10"
Al
^nj
i Mn
oo ,,
•Cj Br
Pb
Ti
Cu
0 Sr
53 P
>> Si
hrl
Ca
6 ?
O IT
s
o ^f
2 C1

^ aMass units in
mean
Too
1.00
0.87
0.90


1.00
0.97
1.00
0.98
1.00
1.00
1.00
1.00
0.71
0.50
1.00
1.00
1.00
1.00
1.00
1.00
0.83
1.00

1.00
1.00
1.00
1.00
1.00
1.00
0.94

195
0~89
0.95
0.78
0.81


0.56
0.93
0.98
0.75
0.65
0.76
0.81
0.97
0.57
0.28
0.85
0.80
0.80
0.90
0.89
0.80
0.62
0.83

0.81
0.68
0.80
0.83
0.96
0.81
0.44

u95 mean
1.11 0.39
1.05 0.03
0.95 0.23
0.99 0.28
fail to converge
fail to converge
1.44 1.63
1.01 0.07
1 .02 0.04
1 .20 0.54
1.35 0.61
1.24 0.70
1.19 0.16
1.03 0.16
0.86 0.78
0.72 0.64
1.15 0.65
1.20 0.80
1 .20 0.69
1.10 0.21
1.11 0.14
1.20 0.60
1 .05 0.77
1.16 0.62
fail to converge
1.19 0.62
1.32 0.63
1.20 0.66
1.17 0.46
1.04 0.21
1.19 0.37
1.43 2.36

195"
0.22
-0.03
0.07
0.15


0.38
0.01
0.02
0.04
-0.02
0.11
-0.04
0.12
0.31
0.05
0.36
0.38
0.30
0.11
0.01
0.22
0.18
0.28

0.26
0.06
0.26
0.17
0.17
0.10
0.48

u95
0.55
0.09
0.38
0.41


2.88
0.12
0.06
1.05
1.25
1.29
0.37
0.19
1.25
1.24
0.93
1.21
1.07
0.32
0.26
0.98
1.36
0.97

0.97
1.20
1.06
0.75
0.26
0.64
4.24

mean 195b
1.7 1.0
0.9 -1.4
0.1 -0.1
0.1 0.0


0.6 0.0
0.0 0.0
0.1 -0.1
6.1 -8.6
11.9 -0.6
4.5 -3.3
0.0 -4.4
1.0 -3.9
0.4 -0.5
5.9 0.1
4.1 2.6
69.5 16.6
0.9 0.1
0.1 0.0
0.0 -0.3
4.0 0.3
0.5 0.0
0.3 0.0

149.3 26.9
118.7 37.3
46.7 8.5
17.6 0.1
6.8 -0.7
1.2 -0.2
45.7 17.6

u95
2.3
3.1
0.2
0.2


1.2
0.0
0.3
20.9
24.4
12.3
4.4
5.9
1.2
11.6
5.7
122.4
1.7
0.3
0.3
7.8
1.1
0.5

271.8
200.1
84.8
35.2
14.3
2.5
73.9

mean
13.8
9.0
0.2
1.9


3.7
0.1
2.0
14.4
165.6
23.8
121.3
27.1
2.9
102.6
21.9
97.6
1.1
1.8
2.1
10.0
3.5
2.6

296.4
800.0
73.0
215.7
68.0
4.0
320.2

195"
10.2
-2.5
-0.4
1.3


0.2
-0.1
1.1
-58.3
72.0
-16.3
85.7
2.4
-1.5
54.0
13.6
-159.0
-2.7
1.2
0.4
-8.4
0.4
1.2

-293.9
329.0
-109.8
116.9
29.3
-3.0
107.0

u95
17.3
20.5
0.8
2.5


7.2
0.2
2.9
87.2
259.1
63.9
156.9
51.7
7.4
151.2
30.2
354.2
4.9
2.5
3.9
28.4
6.5
3.9

886.6
1271.0
255.9
314.5
106.7
11.0
533.4

mean
1.1
3.0
0.5
0.6


3.8
0.1
0.8
57.3
34.1
23.8
8.9
4.0
7.5
20.6
5.6
154.5
1.2
0.4
0.0
10.3
3.2
0.9

237.8
107.6
51.5
43.6
22.7
7.4
148.4

195b u95
0.0 2.1
-3.7 9.8
0.2 0.9
0.3 0.9


1.4 6.3
0.0 0.2
0.2 1.3
12.5 102.0
3.4 64.8
1.8 45.7
-0.5 18.3
-3.7 11.7
4.2 10.9
7.2 34.0
2.6 8.7
52.0 257.0
-0.2 2.6
0.1 0.6
-0.6 0.6
2.6 18.1
1.3 5.1
0.3 1.5

16.1 459.6
-27.0 242.3
-15.5 118.5
8.6 78.5
10.4 34.9
3.4 11.4
49.4 247.4

mg for PM2 5 and PM10 only.
O bA negative lower confidence
H
6
n
H
W




interval implies a nonzero mean is not statistically












significant.








































-------
 1     cigarettes was 9 + 4 mg/cig and the highest estimate was 29 + 7 mg/cig.  The estimates for
 2     cooking ranged from a low of 1.5 ± 0.6 mg/min to a high of 4.9 ± 1.3 mg/min.  The
 3     estimate for other sources ranged from 2.5 +  0.9 mg/h to 12 ±4 mg/h.
 4          Decay rates for elements associated with the fine fraction were generally lower than for
 5     elements associated with the coarse fraction, as would be expected.  For example, sulfur,
 6     which has the lowest mass median diameter of all the elements, had calculated decay rates of
 7     0.16 ± 0.04 and 0.21  + 0.04 h"1 for the PM2 5 and PM10 fractions, respectively.  The
 8     crustal elements (Ca, Al, Mn, Fe), on the other hand, had decay rates ranging from 0.6 to
 9     0.8 h4.
10          Based on the mass-balance model,  outdoor air was the major source of indoor particles,
11     providing about 3/4 of fine particles and 2/3 of inhalable particles in the average home.
12     It was also the major source for most elements, providing 70 to 100% of the  observed indoor
13     concentrations for 12 of the 15 elements. Only copper and chlorine were predominantly due
14     to indoor sources in both the fine particle and inhalable particle fractions.  It should be noted
15     that these conclusions are applicable only to Riverside, CA.  In five of the six cities studied
16     by Harvard and in both New York counties, outdoor air could not have provided as much as
17     half of the indoor air particle mass,  because the observed indoor-outdoor ratios of the mean
18     concentrations were > 2.
19          Unidentified indoor sources accounted for most of the remaining particle and elemental
20     mass collected on the indoor monitors.  The nature of these sources is not yet understood.
21     They do  not include smoking, other combustion sources, cooking, dusting, vacuuming,
22     spraying, or cleaning, since all these sources together account for less than the unidentified
23     sources.   For example, the unidentified sources accounted for 26%  of the average indoor
24     PMio particles, whereas smoking accounted for 4% and cooking for 5% (Figure 7-26).
25          Of the identified indoor sources, the two most important were smoking and cooking
26     (Figures 7-27 and 7-28).  Smoking was estimated to increase 12-h average indoor
27     concentrations of PM10 and PM2 5 by 2  to 4 ^g/m3 per  cigarette, respectively. Homes with
28     smokers  averaged about 30 /xg/m3 higher levels  of PM10 than homes without  smokers.  Most
29     of this increase was in the fine fraction. Cooking increased indoor  concentrations of PM10
30     by about 0.6 /*g/m3 per minute of cooking,  with most of the increase in the coarse particles.
        April 1995                               7-88       DRAFT-DO NOT QUOTE OR CITE

-------
                                      Cooking
                                        4%
                                             Other Indoor
                                                14%
                                                  Smoking
                                                    5%
                   Outdoor
                     76%
             N - 352 Samples from 178 homes
                Outdoor
                 66%
                                      Cooking
                                        5%
                                                 Other Indoor
                                                    26%
               Smoking
                 4%
             N - 350 Samples from 178 homes

Figure 7-26. Sources of fine particles (PM2 s) and respirable particles (PM10) in all
           homes (Riverside, CA).
April 1995
7-89      DRAFT-DO NOT QUOTE OR CITE

-------
               Outdoor
                60%
                                               Indoor
                                                   Smoking
                                                     30%
             N - 61 Samples from 31 homes
                                     Cooking
                                       3%
               Outdoor
                56%
                                             Other Indoor
                                                 16%
                                                  Smoking
                                                   24%
             N - 61 Samples from 31 homes

Figure 7-27. Sources of fine particles (PM2-5) and respirable particles (PM10) in homes
           with smokers (Riverside, CA).
April 1995                          7-90     DRAFT-DO NOT QUOTE OR CITE

-------
                                                Cooking
                                                    Other Indoor
                ^  ,     .                  	,      8%
                Outdoor
                 62%
                                                  Smoking
                                                    5%
              N = 62 Samples from 33 homes
                                                Cooking
               Outdoor
                 56%

                        \             V      /
                                                   Other Indoor
                                                       16%

                                            Smoking
                                               4%
              N = 62 Samples from 33 homes

Figure 7-28.  Sources of fine particles (PM2 5) and respirable particles (PM10), top and
            bottom panels respectively, for homes with cooking during data collection
            (Riverside, CA).

Source: Ozykaynak et al. (1993).

April 1995                            7-91      DRAFT-DO NOT QUOTE OR CITE

-------
 1           Emission profiles for elements were obtained for smoking and for cooking.  Major
 2      elements emitted by cigarettes were potassium, chlorine, and calcium. Elements associated
 3      with cooking included aluminum, iron, calcium,  and chlorine.
 4           Other household activities such as vacuuming and dusting appeared to make smaller
 5      contributions to indoor particle levels.  An interesting finding was that commuting and
 6      working outside the home resulted in lower particle exposures  than for persons staying at
 7      home.
 8           As with the particle mass, daytime personal exposures to 14 of 15 elements were
 9      consistently higher than either indoor or outdoor concentrations. At night, levels of the
10      elements were  similar in all three types of samples.
11
12      Comparison  of the three large-scale studies
13           The three studies had somewhat different aims and therefore different study designs.
14      The Harvard study selected homes based on various criteria, in particular the requirement
15      that a school-age child be in the home, but did not employ a probability-based study.
16      Therefore the results strictly apply only to the homes in the sample and not to a wider
17      population; however, the very large number of homes suggests that the results should be
18      broadly applicable to homes with school-age children in the six cities. The New York study
19      used a probability-based  sample, but stratified on the basis of combustion sources.  Therefore
20      there are likely to be a higher number of homes  with kerosene heaters, wood stoves, and
21      fireplaces in the sample than in the general population.  The PTEAM study used a fully
22      probability-based procedure, and therefore its results are the most broadly  applicable to the
23      entire population of Riverside  households.  However, the participants were limited to
24      nonsmokers,  and therefore homes with only smokers were excluded; therefore indoor
25      concentrations  are likely to be slightly underestimated.
26           The three studies employed different monitors with different cutpoints; therefore exact
27      comparisons  are not possible.  However, large differences between the PM3 5 and PM2 5
28      cutpoints are not likely, and therefore these results can  be more readily compared. In what
29      follows, we will use the  term  "fine particles" to  refer to the PM3 5 and PM2 5 size fractions
30      collected in the three studies.
31

        April 1995                                7-92      DRAFT-DO NOT QUOTE  OR CITE

-------
1 Indoor-outdoor relationships. Outdoor concentrations of fine particles in five of the
2 six cities and the two New York counties were relatively low, typically in the range of 15 to
3 25 /-ig/m3 (Table 7-18). Only Steubenville, with an annual mean of 40 /xg/m3 (but a range
4 among the outdoor sites of 20 to 60 /Ltg/m3) approached the mean outdoor level of 50 jug/m3
5 observed in Riverside. It is interesting to note that indoor concentrations exceeded outdoor
6 concentrations in the seven sites with low outdoor levels, (indoor/outdoor ratios were
7 contained in a small range between 1.9 and 2.4), but were slightly less than outdoor
8 concentrations in the two sites with high
9
10
outdoor levels (ratios of 0.9).





TABLE 7-18. INDOOR-OUTDOOR MEAN CONCENTRATIONS (jtg/m3) OF FINE
PARTICLES IN THREE LARGE-SCALE STUDIES
Study Name
Harvard 6-City Study
Portage, WI
Topeka, KN
Kingston-Harriman, TN
Watertown, MA
St. Louis, MO
Steubenville, OH
New York State ERDA Study
Onondaga County
Suffolk County
EPA Particle TEAM Study
Riverside, CA
Homes Out In
11 10 20
10 10 22
8 18 44
8 15 29
10 18 42
8 45 42
224 17 37
209 22 46
178 50 43
In/Out
2.0
2.2
2.4
1.9
2.3
0.9
2.2
2.1
0.9
       Harvard:  PM3 5 measured using cyclone sampler.  Samples collected every sixth day for one year (May 1986
                to April 1987).
       NYS:     PM2 5 measured using impactor developed at Harvard.  Samples collected for one week at each
                household between January and April 1986.
       PTEAM:  PM2 5 measured using Marple-Harvard-EPA sampler.  Samples collected for two 12-h periods at each
                home between September and November 1990.

       Source: Harvard data—Spengler et al., 1981.
              NYS data—Sheldon et al., 1989.
              PTEAM data—Pellizzari et al., 1993.
1           Effect of smoking.  All three studies found cigarette smoking to be a major source of
2      indoor fine particles.  All three studies compared fine particle concentrations in homes with

       April 1995                                  7.93       DRAFT-DO NOT QUOTE OR CITE

-------
 1     smokers to those in homes without smokers.  The annual mean (every sixth day)
 2     concentration in the 35 six-city  homes with no smokers was 24.4 /xg/m3, compared to a
 3     mean of 36.5 /ig/m3 in 15 homes with one smoker and a mean of 70 /-tg/m3 in five homes
 4     with two or more smokers.  Neas et al. (1990) derived a difference of 31.2 /ug/m3 in
 5     smoking homes compared to nonsmoking homes based on all 1237 homes in Phase II of the
 6     6-City Study.  In the New York State study, 62 Onondaga homes with smokers had a mean
 7     indoor PM2 5 concentration of 69 jiig/m3 compared to 24 jug/m3 in 182 nonsmoking homes;
 8     the indoor value in 80 Suffolk homes with smokers was 76 /xg/m3 compared to 30 jug/m3 in
 9     129 homes without smokers. Thus homes with smokers had PM2 5 weekly mean increases of
10     about 45 to 46 /*g/m3. A similar calculation was done for the PTEAM Study, resulting in a
11     mean of 36 /xg/m3 in 119 homes with no smokers, 63 /xg/m3 in 18 homes with one smoker
12     and at least one cigarette smoked during the 24-h monitoring period, and 69 ptg/m3 in 15
13     homes with two or more smokers and at least  one cigarette smoked during the monitoring
14     period.  The latter result is somewhat muddied by the fact that outdoor concentrations were
15     somewhat higher at the homes with smokers (about 60 /xg/m3 in 33 homes with smokers and
16     only 49 ^g/m3 in 119 homes without smokers).  Thus homes with smokers were about 20
17     ng/m3 higher (after allowing for higher outdoor concentrations) than homes without smokers.
18     A similar correction for outdoor concentrations cannot be made in the New York State study,
19     due to the lack of sufficient matched outdoor measurements.  These bivariate calculations  are
20     not highly trustworthy, since they assume that all other sources affecting indoor air are
21     independent of smoking status,  which may not be the  case. For example,  smokers are more
22     likely to be found in the  lower socioeconomic brackets. Therefore they may have smaller
23     homes, which would lead to higher indoor concentrations in general.  They may also be
24     located in less desirable (more polluted) parts  of town, which would lead to higher outdoor
25     concentrations, as was observed in the PTEAM Study. This would lead to overestimates  of
26     the effect of smoking on indoor concentrations.  On the other hand, persons in homes with
27     smokers may open windows or  otherwise ventilate more extensively to reduce perceived
28     smoke or particle loadings.  This would lead to underestimates of the effect of smoking.
29     Therefore multivariate regressions accounting  for the effect of volume, air exchange, and
30     outdoor concentrations are more trustworthy than the above bivariate calculations.
        April 1995                              7-94      DRAFT-DO NOT QUOTE OR CITE

-------
  1           Multivariate calculations in all three studies result in rather similar estimates of the
  2      effect of smoking on fine particle concentrations.  Spengler et al. (1981) estimated an
  3      increase of about 20 /xg/m3 per smoker based on 55 homes from all six cities.  Since homes
  4      with at least one smoker probably average at least 1.3 smokers per home, this corresponds to
  5      about 26 /xg/m3 per smoking home.  Spengler et al. (1985) found a smoking effect of about
  6      32 /xg/m3 for smoking homes in multivariate models based on the Kingston-Harriman data.
  7      Sheldon et al. (1989) found  an increase of 45 (Onondaga) and 47 (Suffolk) /xg/m3 per
  8      smoking home in a multivariate model of the New York State data.   Ozkaynak et al. (1993)
  9      found an increase of about 30 to 35 /xg/m3 in smoking homes  in a multivariate regression
10      model of the PTEAM data.   Thus the estimates of the effect of a smoking home on indoor
11      fine particle concentrations range from about 26 to about 47 /xg/m3.
12           Dockery and Spengler  (1981) found an effect of 0.88 /xg/m3 per cigarette for homes
13      without air conditioning, and 1.23 /xg/m3 per cigarette for homes with air conditioning, based
14      on 68 homes from all six cities.  Lebret found an effect of 0.8 /xg/m3 per cigarette for homes
15      in the Watertown area.  Leaderer et al. (1991) found an effect ranging between 0.27 and
16      0.33 /xg/m3 per cigarette smoked over a week's time, corresponding to 1.9 to 2.3 /xg/m3 per
17      cigarette contribution to a 24-h average.   In a series of stepwise regressions on the PTEAM
18      data,  Ozkaynak et al. (1993) found an effect ranging between  2.5 and 4.7 /xg/m3 per cigarette
19      per 12-h monitoring period;  this corresponds to an effect ranging between 1.2 and 2.4  /xg/m3
20      per cigarette smoked during  a 24-h period.  Taking the midpoint of these ranges  leads  to
21      estimates for the three studies of about 1.1, 2.1, and 1.8 /xg/m3 increases in fine  particle
22      concentrations per cigarette  smoked in the home over a 24-h period.
23           Both  the New York State study and the PTEAM study were able to estimate source
24      strengths for different variables using a mass-balance model.  The estimates for PM2 5
25      emissions from cigarettes were very comparable, with Koutrakis et al. (1992) estimating
26      12.7 mg/cig compared to the PTEAM estimate of 13.8 mg/cig (Ozkaynak et al.,  1993).
27      Both  studies also found similar elemental profiles for smoking, with potassium  and chlorine
28      being emitted in substantial amounts.
29
30          Effect of other variables. In the PTEAM Study, the second most powerful indoor
31      source of PM10, and possibly PM2 5 particles, was cooking.  Quite large emission strengths

        April 1995                               7.95       DRAFT-DO  NOT QUOTE OR CITE

-------
 1     of several mg/minute of cooking were determined from the mass-balance model, while
 2     multiple regressions indicated that cooking could contribute between 10 and 20 jwg/m3 PM10,
 3     and somewhat smaller amounts of PM2 5, to the 12-h concentration.  Apparently neither of
 4     the other studies analyzed for the effect of cooking, although Dockery and Spengler (1981)
 5     suggested cooking as one possible reason for the observed 15 /ig/m3 residual in their indoor
 6     particle model.
 7          In the New York State study, homes with kerosene heaters had increased PM2 5
 8     concentrations of 15.8 /xg/m3 in Suffolk (p < 0.03) and 3.65 ^g/m3 in Onondaga (p <  0.3,
 9     not significant).  Both the New York State and PTEAM studies also measured air exchange
10     in every home, and both studies  found that air exchange significantly affected indoor particle
11     concentrations. In the PTEAM study, increased air exchange led to increased indoor air
12     concentrations for both PM2 5 and PM10 at night only, perhaps because outdoor
13     concentrations were larger than indoor levels at night. In the New York State study,
14     increased  air exchange led to decreased RSP concentrations in Onondaga (p < 0.02) but no
15     effect was noted in Suffolk (p  < 0.90). In both of these counties, indoor levels generally
16     exceeded  outdoor levels, so increased air exchange would generally reduce indoor
17     concentrations.
18          Both the New York State and PTEAM studies found a very small but significant effect
19     of house volume.  In Onondaga, PM25 concentrations decreased by —1.1 /zg/m3 per
20     1,000 cubic foot increase in volume; in Suffolk concentrations decreased by —0.75 /-ig/m3
21     per 1,000 cubic feet.  In the PTEAM study, PM2 5 daytime concentrations decreased by
22     -2.0 /xg/m3  per thousand cubic feet. Probably because of a significant negative correlation
23     between house volume and air exchange rate,  the two variables did  not both reach
24     significance in the same regression.
25
26     Other Studies
27          Several other large-scale  studies of homes have taken place in other countries, and a
28     number of smaller studies have occurred in the U.S.  These will be discussed in order of the
29     number of homes included in the study.
30          Lebret et al. (1990) carried out week-long RSP measurements (cutpoint not described)
31     in 260 homes in the cities of Ede and Rotterdam, the Netherlands, during the winters  of 1981

       April 1995                               7-96      DRAFT-DO NOT QUOTE OR CITE

-------
 1     to 1982 and 1982 to 1983, respectively. 60% of the Ede homes and 66% of the Rotterdam
 2     homes included smokers.  Diary information collected during the measurement period
 3     indicated that, on average, 1 to 2 cigarettes were smoked during the week, presumably by
 4     guests, even in the nonsmoking homes.  Homes with one smoker averaged 7 cigarettes
 5     smoked per day at home in Ede (N = 53) and 11 per day in Rotterdam (N  = 35).  Homes
 6     with two smokers averaged 21 cigarettes per day in Ede (N  = 23) and 25 cigarettes per day
 7     in Rotterdam (N = 15).
 8          Geometric means  for the combined smoking and nonsmoking homes were similar in the
 9     two cities (61 and 56 jug/m3, respectively), with maxima of 560 and 362 /-ig/m3. Outdoor
10     concentrations averaged about 45 /ig/m3 (N not given).  Indoor concentrations in the homes
11     with smokers averaged  about 70 /*g/m3 (calculated from data in the paper), compared to
12     levels in the nonsmoking homes of about 30 /xg/m3.  Multiple regression analysis indicated
13     that the number of smoking occupants explained about 40%  of the variation in the  log-
14     transformed RSP concentrations—family size, frequency of vacuuming, volume of  the living
15     room,  type of space heating, and city (Ede versus Rotterdam) had no significant effect on
16     RSP concentrations.  In a second regression, the number of  smoking occupants was replaced
17     by the number of cigarettes and cigars  smoked during the  week. The  regression equation
18     was
19
20                    log(RSP)  =1.4 + 0.37 log(# cigarettes) +  0.53 log(# cigars)
21                                     +0.03 log(family size)
22                          R2 = 0.49; d.f. =250 F = 83.7  p  < 0.0001
23
24     From this equation, the authors estimated that one cigarette smoked  per day would increase
25     weekly average indoor RSP concentrations by 2 to 5 jig/m3, whereas one cigar smoked per
26     day would increase indoor levels by  10 pig/m3.
27          Instantaneous RSP concentrations were made using a TSI Piezobalance on the day the
28     technicians were setting up the equipment.  The influence  of smoking  on these measurements
29     may be seen in Table 7-19:
30
       April 1995                              7-97       DRAFT-DO NOT QUOTE OR CITE

-------
          TABLE 7-19. INFLUENCE OF RECENT CIGARETTE SMOKING ON INDOOR
       	CONCENTRATIONS OF PM (SIZE UNSPECIFIED)
       Time Since Smoking                               N      RSP (geom. mean)
       No smoking                                        98                  41
       More than 1 h ago                                   18                  52
       Between 1/2 and 1 h ago                              7                  76
       Less than 1/2 an hour ago                            27                 141
       During the measurements                             54                 191
       Source:  Lebret et al. (1990).
 1          Diemel et al. (1981) measured particles in 101 residences as part of an epidemiological
 2     study related to a lead smelter in Arnhem, the Netherlands. The indoor sampler was adapted
 3     from a small aquarium-type pump, collecting samples at a  flowrate of 1  to 1.5 1pm.  The
 4     authors state that particles smaller than 3 to 4 jum in diameter should have been sampled
 5     efficiently, but present no laboratory data on measured cutpoint size. The outdoor samplers
 6     (number not given) were high-volume samplers. The 28-day average levels indoors ranged
 7     from 20 to 570 /*g/m3, with an arithmetic mean of 140 jig/m3 (SD not presented) and a
 8     geometric mean of 120 ng/m3;  corresponding outdoor concentrations (2-mo averages of 24-h
 9     daily samples) ranged from 53.7 to 73.3 /ig/m3 (N not given), with nearly identical
10     arithmetic and geometric means of 64 /xg/m3.
11          Kulmala et al. (1987) measured indoor and outdoor air in approximately 100 dwellings
12     (including some office buildings) in  Helsinki, Finland between 1983 and 1986.  Samples
13     were collected on Nuclepore filters using a stacked foil technique.  Mean concentrations were
14     presented in Table 7-20 for fine (< 1 /xm) and coarse (> 1  /mi) particles (standard deviations
15     not provided):
16
17
18
       April 1995                              7-98      DRAFT-DO NOT QUOTE OR CITE

-------
            TABLE 7-20  INDOOR AND OUTDOOR PM IN BUILDINGS IN HELSINKI,
                   FINLAND, AS A FIUNCTION OF SEASON AND LOCATION.

Fine particles




Coarse particles




Location

Urban
Urban
Suburban
Suburban

Urban
Urban
Suburban
Suburban
Season

Summer
Winter
Summer
Winter

Summer
Winter
Summer
Winter
Outdoors

40
30
15
30

75
30
20
25
Indoors

25
30
20
20

20
15
25
20
       Source:  Kulmala et al. (1987).
 1     The authors noted that the geometric mean for the combined fine particle samples indoors
 2     was 16 ^g/m3, with a 95% range of 4 to 67 /xg/m3. Corresponding values for the indoor
 3     coarse particles were 13 ^ig/m3 with a range of 3 to 63 ^g/m3.  Outdoors, the fine particles
 4     had a geometric mean of 20 /zg/m3 with a 95% range of 5 to 82 jug/m3, and the coarse
 5     particles had a geometric mean of 16 /xg/m3 with a range of 3 to 91 /ig/m3.
 6          Quackenboss et al. (1989) reported PM10 and PM2 5 results from 98 homes in the
 7     Tucson, Arizona area.  Homes were selected as part of a nested design for an
 8     epidemiological  study.  The Harvard-designed dual-nozzle indoor air sampler (Marple et al.,
 9     1987) was employed for indoor air measurements.  Outdoor air was measured within each
10     geographic cluster by the same instrument; supplementary data were obtained from the Pima
11     County Air Quality control District, but these data did not include PM2 5 measurements, and
12     some data were apparently PM15. Homes were classified by tobacco smoking and by use of
13     evaporative ("swamp") coolers; these coolers apparently act as a significant removal
14     mechanism for particles (Table 7-21).  Homes without smoking averaged about 16 /xg/m3
15     PM2 5,  compared to 24  iig/m3 for homes reporting less than a pack a day, and 51 iig/m3 for
16     homes reporting more than a pack a day.  PM2 5 particles formed the bulk of the PM10
17     fraction even in nonsmoking homes, ranging from nearly 80% in those homes to  nearly
18     90% in homes with heavy smoking. This is somewhat surprising in view of the statement
19     made in the paper that the bulk of the PM in Tucson is silica quartz averaging around 5 /xm
       April 1995                              7.99      DRAFT-DO NOT QUOTE OH CITE

-------
           TABLE 7-21. INDOOR AVERAGE PM2 5 AND PM10 (/tg/m3) BY REPORTED
                 SMOKING IN THE HOME AND EVAPORATIVE COOLER USE
                                   DURING SAMPLING WEEK
Smoking
Cigarettes/day
None


1-20


>20


Evaporative
Cooler
Yes
No
Total
Yes
No
Total
Yes
No
Total

Mean
8.8
20.3
15.2
19.3
32.3
27.3
36.2
82.7
60.8
PM2.5
S.D.
5.0
19.0
15.5
8.8
28.5
23.6
32.9
55.4
50.8

Homes
(20)
(25)
(45)
(10)
(16)
(26)
(8)
(9)
(17)

Mean
21.0
38.4
30.3
33.9
53.4
46.2
47.4
102.5
75.0
PM10
S.D.
9.7
22.9
19.9
12.0
33.9
29.1
39.6
60.6
57.2

Homes
(20)
(23)
(43)
(10)
(17)
(27)
(9)
(9)
(18)
       PM2 5: Significant (p < 0.01) main effects for smoking and evaporative cooler use; two-way interaction nearly
             significant (p = 0.06).
       PM10: Significant (p < 0.01) main effects for evaporative cooler and smoking.
       Source:  Quackenboss et al. (1989)
 1     in mean aerodynamic diameter; that should result in the coarse fraction being larger than the
 2     fine fraction.  Outdoor PM10 particles were not strongly correlated with indoor levels
 3     (R2 = 0.18; N about 90). Although about 30%  of homes in the larger sample reported using
 4     fireplaces, no estimate of an effect on indoor air quality was made in the paper.
 5          Quackenboss et al.  (1991) extended the analysis of the Tucson homes over three
 6     seasons.  Median indoor  PM2 5 levels in homes with smokers were about 20 /xg/m3 in the
 7     summer and spring/fall seasons compared to about 10 /zg/m3 in homes without smokers in
 8     those seasons (Table 7-22).  In winter, however, the difference was considerably increased,
 9     with the median level in  24 homes with smokers at about 36 /ig/m3 compared to  13 /zg/m3  in
10     26 homes without smokers.
11          Revsbech et al.  (1987) studied 44 apartments in Aarhus, Denmark.  All were retrofitted
12     or tight apartments in three-story brick buildings. Particles  were measured by an open filter
13     surface directed downwards.  The authors considered that the low suction rate and the
14     downward facing surface would result in collecting predominantly the respirable fraction, but

       April 1995                              7-100      DRAFT-DO NOT QUOTE OR CITE

-------
                TABLE 7-22.  INDOOR PM10 AND PM2 5 (jig/m3) BY SEASON AND
                           ENVIRONMENTAL TOBACCO SMOKE (ETS)
Smokers at home

PM10


PM2.5


Season3
Summer
Spring/Fall
Winter
Summer
Spring/Fall
Winter
Median
35.
43.
80.
20.
20.
35.
3
3
3
5
1
7
Qi
29.3
28.7
40.8
13.7
12.8
26.7
Q3
49.8
64.3
104.9
27.0
43.6
77.8
N
49
38
24
49
39
24
No smokers at home
Median
17
30
31
8
10
13
.5
.0
.4
.9
.6
.4
Qi
14.1
20.5
24.3
5.9
8.7
10.2
Q3
24.4
40.1
42.9
11.9
14.8
19.9
N
49
37
26
50
37
26
       "Seasons:  Summer = May through September;
                Spring/Fall = March, April, October, November;
                Winter = December through February
       QI, Q3:  1st and 3rd quartiles of distribution
       N:       number of households
 1     presented no evidence of work done to determine the sampling efficiency curve or cutpoints.
 2     Sampling occurred at the rate of 3 1pm for 16 h beginning in the evening and including the
 3     overnight sleeping period.  Outdoor levels were not recorded, although existing outdoor
 4     measurements were quoted as having yearly averages of 30 to 38 Mg/m3> with 84th
 5     percentiles at 41 to 51 /zg/m3.  Ventilation rates were measured by introducing CO2 at
 6     5,000 ppm and recording the decay with an infrared gas analyzer for at least 1 h.  The
 7     median number of cigarettes smoked was 8; in eleven homes (25% of the total) no cigarettes
 8     were smoked. The median ventilation rate was 0.23 h"1, with an interquartile range of
 9     0.19 to 0.31 h"1.  In the 11 dwellings without smoking, the median concentration was
10     91 /ig/m3 (IQR 57 to 107 /*g/m3); in the homes with < 10 cigarette equivalents,  the median
11     (IQR) was 169 (49 to 338) /zg/m3; and in homes with more than 10 cigarette equivalents, the
12     median (IQR) was 475 (309 to 587) /ig/m3.  The particle concentration correlated positively
13     with tobacco consumption (rs = 0.716, p< 0.001) and length of daily residence  (rs  =
14     0.405, p < 0.01) but not with the frequency of cleaning (rs  = 0.203, n.s.), time of airing of
15     living rooms (n.s.), or the ventilation rate (rs = 0.277, n.s.).
16          Sexton et al. (1984) reported on a study in Waterbury VT.  This study included  24
17     homes,  19 with wood-burning appliances, and none with smokers.  24-h  samples (0800 to
18     0800) were collected in each home every other day for two weeks,  providing 163 valid

       April 1995                              7_10l      DRAFT-DO NOT QUOTE OR CITE

-------
  1      indoor samples. Indoor RSP levels ranged from 6 to 69 /*g/m3 with a mean value of 25
  2      fJig/m3.  Outdoor levels ranged from 6 to 30 |Ltg/m3 with a mean value of 19 /zg/m3.  Indoor
  3      concentrations were not correlated with outdoor concentrations (r  = 0.11, p> 0.16.)
  4           Kim and Stock (1986) reported results from 11 homes in the Houston  area.  (The year
  5      and the season  were not supplied in the paper.)  For most homes, two 12-h PM2 5 samples
  6      (day and night) were collected for approximately one week.  Sampling methods were not
  7      fully discussed, but apparently involved samples collected using a mobile van near each
  8      home.  The mean weekly concentrations in the five smoking homes averaged 33.0 +
  9      4.7  (SD) /ig/m3, versus  mean outdoor concentrations averaging 24.7 ± 7.4  jug/m3 (calculated
10      from data presented in paper). Indoor concentrations in the six nonsmoking homes averaged
11      10.8 + 4.9 /ig/m3 compared to outdoor levels of 12.0  ± 5.9 jig/rn3.
12          Morandi,  Stock and Contant (1986) reported on 13 Houston homes monitored during
13      1981 as part of a larger personal monitoring study of 30 nonsmoking participants.  The TSI
14      Piezobalance (cutpoint at about PM3 5) was employed for the personal monitoring, with
15      technicians "shadowing" the participants and taking consecutive 5-min readings.  At the
16      homes,  dichotomous samplers (cutpoints at PM2 5 and PM10) were used for  two 12-h daytime
17      samples (7 a.m. to 7 p.m.) both inside and outside the  homes for seven consecutive days.
18      Little difference was  noted in the indoor concentrations at homes (25  ± 30 (SD) /zg/m3) and
19      at work or school (29 ± 25 /Ag/m3).  The authors noted that the highest overall respirable
20      suspended particle  (RSP) concentrations occurred in the presence of active smoking
21      (89 /xg/m3), significantly different from mean RSP values measured in the absence of
22      smokers (19 /xg/m3; p < 0.0001). Among homes with smokers, those homes with central
23      air conditioning were significantly (p< 0.0001) higher (114 versus 52 /zg/m3) than those with
24      no air conditioning; in this case, the outdoor differences only increase this contrast.  Cooking
25      was associated with significantly higher RSP concentrations (27 pig/m3 compared to 20
26      Mg/m3, p < 0.01).  The single highest RSP concentration (202 /*g/m3) was found in  a home
27      with no smokers and no air conditioning but with active cooking.  The authors concluded that
28      cooking was a more important source of indoor RSP than smoking, at least in the small
29      number of homes they studied.
30          Coultas et al. (1990) measured PM2 5 and nicotine in 10 homes containing at least one
31      smoker.  The authors used the Harvard aerosol impactor with sodium bisulfate-treated filters

        April 1995                              7-102      DRAFT-DO NOT QUOTE OR CITE

-------
  1      for nicotine collection, with analysis by GC-FID using a modified Hammond method.
  2      Samples were collected for 24 h every other day for 10 days and then for 24 h every other
  3      week for 10 weeks, resulting in  10 samples per household.  The mean concentrations of
  4      PM25 ranged from 32.4 +  13.1 (SD) to 76.9 + 32.9 /xg/m3; of nicotine, from 0.6 ± 0.7 to
  5      6.9  ± 8.2 ,ug/m3. Outdoor particle concentrations were not reported; thus it is difficult to
  6      calculate the portion of the observed PM2 5 that might be due to ETS. The authors employed
  7      a regression technique together with questionnaire  variables indicating the periods of smoking
  8      to derive an estimate of 17 /xg/m3 as the contribution of smoking to indoor PM2 5; however,
  9      the 95% confidence bounds for this estimate ranged from —3 to 38 ^g/m3, indicating that it
 10      is not significantly different from zero. The Spearman correlation between total PM2 5 and
 11      nicotine was 0.54 (N  = 99).
 12           Kamens  et al. (1991) measured indoor particles in three homes without smokers in
 13      North Carolina in November and December 1987  (no measurements of outdoor particles
 14      were taken).  Two dichotomous samplers (PM2 5 and PM10), several prototype personal
 15      samplers (also PM2 5 and PM10, and three particle sizing instruments including a TSI
 16      electrical aerosol mobility analyzer (EAA) with  10 size intervals between 0.01 and 1.0 /mi,
 17      and  two optical scattering devices covering the range of 0.09 to 3.0 and  2.6 to 19.4 /mi were
 18      employed.  Air exchange measurements were made using SF6 decay over the course of the
 19      seven 8-h (daytime) sampling periods.  There were also three 13-h (evening and overnight)
 20      sampling periods.  For the entire study, 37% of the estimated total mass collected was in the
 21      fine fraction, and another 37% in the fraction greater than 10 /mi.  The remainder (26%)
 22      was  in the coarse (PM10 - PM2 5) fraction.  However, considerable variation was noted in
 23      these size distributions.  For example,  on one day  with extensive vacuuming, cooking, and
 24      vigorous exercising of household pets, 52% of the total mass appeared in the fraction larger
 25      than 10 /mi, with only 18% in the fine fraction. The peak in particle mass on that day
 26      coincided with vacuuming and sweeping of the carpets and floors.  On another day, stir-fried
 27      vegetables  and rice produced a large number of small particles,  with those less than 0.1 /mi
28      accounting for 30% of the total EAA particle volume, much larger than the  normal amount.
29      The  cooking contribution of that one meal  to total  8-h daytime particle volume exposure was
30      calculated to be in the range of 5 to 18%.  The authors concluded that the most significant
31      source of small particles (<2.5 ^m) in all  three of these nonsmoking homes was cooking,

        April 1995                               7_103      DRAFT-DO NOT QUOTE OR CITE

-------
 1     while the most significant source of large particles (> 10 /xm) was vacuum sweeping.  Coarse
 2     particles (PM10 - PM2 5) appeared to be of largely biological (human dander and insect parts)
 3     and mineral (clay, salt, chalk, etc.) origin.
 4          In a test of a new sampling device (a  portable nephelometer), Anuszewski, Larson and
 5     Koenig (1992) reported results from indoor and outdoor sampling at nine Seattle homes
 6     sampled for an average of 18 days each during the winter of 1991 to 1992.  The
 7     nephelometer is a light-scattering device with rapid (1-min) response to various household
 8     activities such as sweeping, cigarette smoking, frying, barbecuing, and operating a fireplace.
 9     Homes with fewer activities showed high correlations of indoor and outdoor light-scattering
10     coefficients, both between hourly averages  and 12-h averages.  However,  homes with
11     electrostatic precipitators, with weather-stripped windows or  doors, and with gas cooking or
12     heating devices showed weak 12-h indoor-outdoor correlations.  One home with a fireplace
13     produced the strongest indoor-outdoor light scattering relationship (R2  = 0.99); this is
14     presumably due to the very high air exchange rate produced  by the fireplace.
15          Chan et al. (1995) studied particles and nicotine in seven homes with one smoker each
16     in Taiwan.  Sampling was carried out in summer and winter of 1991.  Each home had one
17     indoor PM5 sampler in the living room and another in the yard.  In the winter study, two
18     homes had PM10 samplers added inside and outside and at two central  sites. Indoor mean
19     PM5 concentrations  averaged 44 + 32 (SD) ^ig/m3 in summer compared to outdoor
20     concentrations of 27 ± 15 /ig/m3. Corresponding winter values were  107 ± 44 /ig/m3 and
21     92 + 40 Mg/ni3. Mean cigarette butt counts were only 11.1 in the summer and 6.1 in the
22     winter, and nicotine levels were only 0.4 to 0.5 ^ig/m3 in the two seasons; thus the homes
23     appeared to include  fairly light smokers.  It was  calculated that these smokers  produced an
24     average of about 16 /ig/m3 of PM5 daily.
25          Daisey et al. measured RSP, PAH, and extractable organic  matter (EOM) in seven
26     Wisconsin homes with wood stoves; one 48h (1,000 m3) sample was collected during
27     woodburning and a  second sample was collected when no woodburning occurred.  Five of
28     seven homes had somewhat higher RSP levels during woodburning, but the mean difference
29     was not significant.   On the other hand, all homes had from 2 to 40 times higher levels of
30     PAH during woodburning, with the means being  significantly different. Mean values of
31     EOM were also significantly higher during woodburning.

       April 1995                              7-104      DRAFT-DO NOT QUOTE OR CITE

-------
  1           Highsmith et al. (1991) reported on the results of two 20-home studies in Boise, ID and
  2      Roanoke, VA.  The Boise study was designed to assess the effects of wood burning on
  3      ambient and indoor concentrations in the area.  Ten homes with wood burning stoves were
  4      matched with 10 homes without wood burning stoves.  One matched pair of homes was
  5      monitored from Saturday through Tuesday  for eight consecutive 12-h periods.  Ambient
  6      PM2 5 concentrations increased by about 50% at night, suggesting an influence of
  7      woodburning.   Indoor PM2 5 concentrations also were increased (by about 45%)  in the homes
  8      with the wood burning stoves compared to those without (26.3 versus 18.2 /xg/m3), although
  9      coarse particles showed no increase  (10.2 versus 9.7 Mg/m3).  The Roanoke study, designed
10      to assess the effects of residential oil heating, showed no effects on indoor  levels of fine or
11      coarse particles.
12           Lofroth et al. (1991) measured particle emissions from cigarettes, incense sticks,
13      "mosquito coils," and frying of various foods.  Emissions  were 27 and 37 mg/g  for two
14      brands of Swedish cigarettes,  51 and 52 mg/g for incense  sticks and cones, and 61 mg/g for
15      the mosquito coil.  Emissions from pork, hamburgers, herring, pudding, and Swedish
16      pancakes ranged from 0.07 to 3.5 mg/g. Extracts of some of the foods showed mutagenic
17      activity. Other studies of mutagenicity of foods from cooking were referenced:  Berg et al.
18      (1988) and Teschke et al. (1989).  The authors concluded  that indoor air pollution from
19      cooking requires further study.
20           Mumford et al. (1991) measured PM10, PAH, and mutagenicity in eight mobile homes
21      with kerosene heaters.  Each home was monitored for 2.6  to 9.5 h/day (mean of 6.5 h) for
22      three days a week for two weeks with the kerosene heaters off and for two weeks with them
23      on (average on-time of 4.5 h).  Mean PM10 levels were not significantly increased when the
24      heaters were on (73.7 ± 7.3 (SE) ^g/m3 versus 56.1  ± 5.7 ^tg/m3), but in two homes levels
25      increased to 112 and 113 /ig/m3 when the heaters were on. Outdoor concentrations averaged
26      18.0 ± 2.1 jiig/m3.  PAH and CO concentrations and mutagenicity were significantly
27      increased when the heaters  were on.
28           Colome et al. (1990) measured particles using PMi0  and PM5 (cyclone) samplers inside
29      and  outside homes  of 10 nonsmokers, including eight asthmatics, living in Orange County,
30      CA.  Indoor PM10 samples were well below outdoor levels for all homes (mean of 42.5  ±
       April 1995                              7-105      DRAFT-DO NOT QUOTE OR CITE

-------
 1     3.7 (SE) ^g/m3 indoors versus 60.8 ± 4.7 /tg/m3 outdoors). No pets, wood stoves,
 2     fireplaces, or kerosene heaters were present in any of these homes.
 3          Lioy et al. (1990) measured PM10 at eight homes (no smokers) for 14 days in the
 4     winter of 1988 in the town of Phillipsburg,  NJ, which has a major point source consisting of
 5     a grey iron pipe manufacturing company. The Harvard impactor was used indoors to collect
 6     14 24-h samples beginning at 4:30 p.m. each day;  Wedding hi-vol PM10 samples  were
 7     deployed at  three outdoor sites.  A fourth outdoor site was located on a porch directly across
 8     the street from the pipe sampler.  The first three sites showed little difference from one
 9     another,  whereas on day 4 and day 6 of the study, the outdoor sampler on the porch had
10     readings that were considerably (about 40 ng/m3) higher than the other outdoor samplers,
11     suggesting an influence of the nearby point source.  The geometric mean outdoor PM10
12     concentration was 48 ^g/m3 (GSD not provided) compared to 42 /*g/m3 indoors.  A simple
13     regression equation for all homes (N = 101  samples) explained 45% of the variance in
14     indoor PM10:
15
16                            Indoor PM10 =  0.496 Outdoor PM10 + 21.5
17
18          Thatcher and Lay ton (1994) measured particle size distributions inside and outside a
19     residence in the summer.  Measured deposition velocities for particles  between 1 and 5 jim
20     closely matched the calculated gravitational  settling velocities; however,  for particles
21      >5 fjim, the deposition velocity was less than the calculated settling velocity, perhaps due to
22     the non-spherical nature of these particles.  The authors calculated a penetration factor of 1,
23     agreeing with the findings of the PTEAM Study.  They also determined  that resuspension of
24     particles had a  significant impact on indoor particle concentrations. Merely walking into a
25     room increased the particle concentration by 100% for particles > 1 jim. The authors
26     calculated a resuspension rate of 2 to 8 X 10~3 h"1 for four persons in  a house performing
27     normal activities (i.e., close to 0.01%) of the total collected house dust could be resuspended
28     in an hour.  For a surface mass of house dust per unit area of 220 /xg/cm3, this rate resulted
29     in an airborne concentration due to resuspension of about 10 /zg/m3, compared to an outdoor
30     contribution of similar magnitude.
        April 1995                               7-106      DRAFT-DO NOT QUOTE OR CITE

-------
 1          Because house dust can be resuspended, it will contribute to total airborne exposure to
 2     particles and constituents such as metals and pesticides.  Roberts et al. (1990) studied
 3     42 homes in Washington State. Geometric mean lead concentration in 6 homes where shoes
 4     were removed on entry was 240 /xg/m2 on carpets, compared to 2,900 /ig/m2 on carpets in
 5     homes where shoes were kept on.  This finding suggests that most of the carpet dust in a
 6     home enters via track-in on shoes rather than infiltration by air.
 7
 8     7.6.2.2   Studies in buildings
 9          The single largest study of particles in buildings was carried out by the Lawrence
10     Berkeley Laboratory (LBL) for the Bonneville Power Administration (BPA) (Turk et al.,
11     1987,  1989).  38 buildings were chosen from two climatic regions in the Pacific Northwest:
12     Portland-Salem, Oregon (representing mild coastal conditions), and Spokane-Cheney,
13     Washington (representing extreme inland conditions).  The buildings were studied for a
14     variety of pollutants to determine how ventilation rates affect indoor air quality.   Buildings
15     were measured in winter (21 buildings in both regions),  spring (10  buildings in both regions)
16     and  summer (nine buildings in the inland region only).  All but four buildings were
17     government or public properties, and therefore cannot be considered to represent the full mix
18     of building types.
19          Each building was monitored for 10 working days  over a two-week period.  From 4 to
20     8 particle sampling sites were chosen in each building according to size. The  sampler was
21     an LBL-developed flow controlled device with a 3/mi cutpoint. The pumps sampled only
22     during hours the building was occupied.  If filters had to be changed due to excessive
23     loading, the combined weight  of all filters from one site was determined—thus all values are
24     approximately 10 working-day (80-h) averages.
25          Buildings had varied types of smoking policies, from relatively unrestricted to very
26     tightly controlled, as in one elementary school. In most buildings,  an attempt was made to
27     site  at least one monitor in an area where smoking was allowed.  Data was obtained from
28     smoking areas in about 30 of the 38 buildings.
29          Results comparing smoking and non-smoking areas are provided in Table 7-23 and
30     Figure 7-29. Mean RSP concentrations in the smoking areas were  more than three times
31     higher than in the non-smoking areas (70 versus 19 jwg/m3).  Since  these arithmetic means

       April 1995                               7-107      DRAFT-DO NOT QUOTE OR CITE

-------
               TABLE 7-23. SMOKING, NON-SMOKING, AND OUTDOOR RSP
                             CONCENTRATIONS AND RATIOS

Outdoor
(/igm 3)
Building No.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30"
31
32
33
34
35
36"
37
38
39
40
AM
ASD
GM
GSD
ND
ND
ND
8
BD
35
35
8
8
9
8
ND
10
6
BD
10
7
7
7
18
17
20
11
11
68
32
52
65
29
33
13
ND
ND
16
18
20
19
14
11
11
19
16
14
2.2

Indoor

0*gnr3)
Arithmetic Mean (Range)
Non-Smoking Smoking0 Meand
25(19-36)
19(18-21)
ND
7(6-8)
13(13)
12(11-13)
38(32-44)
7(7-8)
11(11)
65(53-74)
23(9-49)
10(10)
5(5-6)
ND
11(7-14)
9(8-11)
11(10-13)
ND
ND
11(10-11)
11(9-12)
18(18)
9(BD-20)
44(10-77)
35(32-38)
45(20-70)
36(33-38)
36(29-43)
10(8-12)
24(20-30)
12(8-18)
13(10-17)
ND
13(10-16)
20(6-35)
14(9-18)
21(12-32)
7(BD-9)
8(8-9)
10(8-12)
19
14
15
1.9
ND
ND
20(16-25)
ND
14(14)
35(23-59)
39(39)
ND
16(13-20)
95(67-127)
209(209)
63(63)
ND
30(26-34)
12(12)
73(73)
105(105)
19(19)
20(11-29)
ND
ND
57(22-165)
ND
24(24)
109(109)
82(55-123)
61(33-89)
BD
144(144)
113(113
268(268)
36(21-52)
29(12-74)
54(13-117)
50(50)
72(17-127)
27(11-62)
308(308)
13(11-14)
26(11-40)
70
73
44
2.7
"Repeat test of building #1 1 .
bRepeat test of building #17.
cSmoking within 10 m radius of site.
25(19-36)
19(18-21)
20(16-25)
7(6-8)
13(13-14)
28(11-59)
38(32-44)
7(7-8)
15(11-20)
86(53-127)
63(9-209)
36(10-63)
5(5-6)
30(26-34)
11(7-14)
31(8-73)
40(10-105)
19(19)
20(11-29)
11(10-11)
11(9-12)
50(18-165)
9(BD-20)
37(10-77)
60(32-109)
67(20-123)
48(33-89)
24(BD-43)
32(8-144)
37(20-113)
64(8-268)
21(10-52)
29(12-74)
28(10-117)
23(6-50)
28(9-127)
25(11-62)
46(BD-308)
11(8-14)
15(8-40)
30
19
24
2.0
NA
ND
BD

Ratios
Indoor Non- Indoor
Smoking •*• Smoking •*•
Outdoor Outdoor
NA
NA
NA
0.9
NA
0.3
1.1
0.9
1.3
7.0
2.9
NA
0.5
NA
NA
0.9
1.6
NA
NA
0.6
0.7
0.9
0.8
4.0
0.5
1.4
0.7
0.6
0.3
0.7
0.9
NA
NA
0.8
1.1
0.7
1.1
0.5
0.7
0.9
1.2
1.3
0.9
2.0
= Not applicable.
= No data collected.
= Below detection limit.
NA
NA
NA
NA
NA
1.0
1.1
NA
2.0
11.0
26.1
NA
NA
5.0
NA
7.3
15.0
2.7
2.9
NA
NA
2.9
NA
2.2
1.6
2.6
1.2
NA
5.0
3.4
20.6
NA
NA
3.4
2.8
3.6
1.4
22.0
1.3
2.4
6.0
7.2
3.6
2.6



Indoor
Mean -H
Outdoor
NA
NA
NA
0.9
NA
0.8
1.1
0.9
1.9
9.6
7.9
NA
0.5
5.0
NA
3.1
6.1
2.7
2.9
0.6
0.7
2.5
0.8
3.4
0.9
2.1
0.9
0.4
1.1
1.1
4.9
NA
NA
.8
.3
.4
.3
.3
.0
.4
2.3
2.2
1.7
2.3


      dArithmetic average of all sites in building.
1     showed evidence of being driven by one or two high values, the geometric mean (averaged
2     across all sites in a building) may be a better comparison.  Here the ratio is very close to
      April 1995
7-108
DRAFT-DO NOT QUOTE OR CITE

-------
80
70
60

50
     CO
     I
     I  30
        20
        10
         0
                            Mean Concentrations
            Smoking areas  Nonsmoking areas   Outdoors
     CO

     "oi
50

40

30
     fe  20
     DC

        10
                             Geometric Means
            Smoking areas  Nonsmoking areas   Outdoors
Figure 7-29. Comparison of respirable particles in smoking and non-smoking areas of
          38 buildings in the Pacific Northwest.

Source: Turk et al. (1987).
April 1995
                        7-109
DRAFT-DO NOT QUOTE OR CITE

-------
 1     3 to 1 (44 versus 15 /ig/m3).  If each of the 70 smoking and 106 nonsmoking sites is allowed
 2     to contribute equally to the average, the ratio of geometric means declines (34 versus
 3     14 /ig/m3).  However, the geometric standard deviation (GSD) for the smoking sites is
 4     2.5 compared to only 1.9 for the nonsmoking sites—thus, the 97.7th percentile is likely to be
 5     more than 6 (2.52) times the geometric mean in the smoking areas (i.e., about 200 /ig/m3)
 6     compared to only about 4 (1.92) times the geometric mean in the nonsmoking areas (i.e.,
 7     about 56 /ig/m3). Outdoor results at 30 sites had the identical arithmetic mean as the indoor
 8     non-smoking sites: 18.9 /ig/m3.
 9          Repace and Lowrey (1980) sampled 19 establishments allowing smoking (seven
10     restaurants, three bars, church bingo games, etc.) and 14 where no smoking occurred
11     (including five residences and four restaurants) between March and early May of 1978.
12     Sampling occurred for short periods of time (2 to 50 min) using a TSI Piezobalance to
13     measure PM3 5.  Indoor concentrations ranged from 24 to 55 /ig/m3 in the areas without
14     smoking, and from 86 to  697  /ig/m3 in places with active smoking.  Five of the locations
15     with active smoking were sampled at a time or place when no smoking was occurring; the
16     matched concentrations (smoking/no smoking) were 279/30;  110/55; 109/30; 86/51; and
17     107/30 /ig/m3.  Because ventilation conditions may have changed between visits, these values
18     cannot be considered fully quantitative.
19          Miesner et al. (1989) sampled particles and nicotine in  57 locations within 21 indoor
20     sites in  Metropolitan Boston between July 1987 and Feb.  1988.  PM2 5 was sampled using
21     Harvard aerosol impactors. Nicotine was sampled using sodium  bisulfate-impregnated filters
22     placed downstream from the Teflon filters for the particles.  Sampling times ranged from
23     about 3 h in a bus station to 16 h in a library, depending partly on how "clean" the
24     environment was perceived to be. PM2 5 concentrations ranged from 6 /ig/m3 (in the library)
25     to 521  /ig/m3 in a smoking room in an office building. For 42 measurements in
26     non-smoking areas, the mean PM2 5 concentration was 25  + 30 (SD) /ig/m3. Six of these
27     measurements  included a  classroom with visible levels of chalk dust on the impactor, four
28     measurements  in subways, and the bus station.  The remaining 36 nonsmoking areas had a
29     mean PM2 5 concentration of 15  ± 7 /ig/m3. The  15  smoking areas ranged from 20 to
30     520 /ig/m3 with a mean of 110 ± 120 /ig/m3.
       April 1995                              7-110      DRAFT-DO NOT QUOTE OR CITE

-------
  1           Turner,  Cyr, and Gross (1992) measured RSP using the TSI Piezobalance (cutpoint
  2      about 3.5 /*m) in 585 office environments during 1989.  The offices were selected because
  3      they had contracted with a commercial group (Healthy Buildings International) to perform
  4      indoor air quality evaluations— thus they cannot be considered a probability sample.  Each
  5      office was measured for 1 h (ten 6-min measurements).  If smoking was observed during the
  6      hour,  the office was so  classified.  Mean RSP concentrations were 46  + 57 (SD) pig/m3 for
  7      331 smoking offices, 20 + 17.6 /ig/m3 for 254 nonsmoking offices. Further discriminant
  8      analysis reclassified the smoking offices into light smoking (mean smoking density of
  9      0.075 cig/m2h) and heavy smoking (0.30 cig/m2h).  This analysis  suggested that particle
 10      concentrations in the light smoking offices were very similar to those in the nonsmoking
 11      offices (19 ± 9.2 versus 17 ± 9.5 /xg/m3) while concentrations were much higher in the
 12      heavy smoking offices (85 + 72 /*g/m3).  It must be pointed out that funding for this work
 13      was supplied ultimately by tobacco companies, and that serious allegations have been made
 14      by Congressional staff (U.S. House of Representatives, 1994) regarding data irregularities,
 15      including possible data alteration and fabrication, systematic misreporting of room size,
 16      misclassification of smoking and nonsmoking rooms, and other charges.
 17           Vaughan and Hammond (1990) measured particles  and nicotine before and after a large
 18      corporation adopted a smoking policy limiting smoking to the cafeteria on the 32nd floor.
 19      Nicotine levels in offices on various floors ranged from  1.6 to 24  /xg/m3 before the policy
 20      was instituted, but dropped to 0.1  to 0.5 /^g/m3 afterwards,  an improvement by 84 to 98%.
 21      Particle levels dropped from a range of 20 to 270 jig/m3 before the policy to ND-35 jug/m3
 22      afterwards (only three measurements due to loss of several samples). The authors noted
 23      some evidence that ETS vapors were spilling over from the snack  bar to offices on the same
 24      floor and two adjacent floors on the same  air handler (offices on these  floors had nicotine
 25      levels about 4 times higher than those on more distant floors).
26           Sheldon (1988a,b) reported on the EPA 10-building study of hospitals, homes for the
27      elderly, schools, and office buildings.  Although the main focus of the  study was  VOCs,
28      particle measurements were taken in all buildings.  Measurements  were taken in six buildings
29      using a National Bureau of Standards portable particle  sampler (NBS 1982; McKenzie et al.,
30      1982)  to collect two size fractions: PM3 and a coarse fraction between  PM3 and PM15.  The
31      sampler employed two filters in series: an 8.0 pm Nuclepore filter for PM15 and a 3
       April 1995                               7_1H      DRAFT-DO NOT QUOTE OR CITE

-------
 1     Ghia Zefluor Teflon filter for fine particles.  The flow rate was 6 L/min for a 24-h sample.
 2     Three consecutive 24-h samples were collected at each building.  Additional particle
 3     monitoring was provided at certain locations  (e.g., smoking lounge, cafeteria) using a
 4     Piezobalance (PM3 5) and a dichotomous sampler (PM2 5 and PM10).
 5          In areas without smoking, indoor concentrations of both size fractions were generally
 6     lower than outdoor levels; for example, the coarse fraction ranged from 0.2 to 0.66 of the
 7     outdoor level (13 to 17 /xg/m3) in the three buildings with no smoking.  The fine fraction was
 8     present at higher indoor-outdoor ratios, ranging from 0.56 to 0.99 in the same three
 9     buildings (outdoor fine fraction ranged from  16 to 33  /xg/m3).  The fine fraction was elevated
10     in the regions of smoking (range of 14 to 56  /xg/m3).  Piezobalance  results for several
11     buildings showed uniformly low (7 to 29 /xg/m3) for 800 min of monitoring in nonsmoking
12     areas.
13          Concentrations in the areas allowing smoking were more  often in  the 40 to 60 /xg/m3,
14     with short-term peaks as  high as 345 itg/m3 (Figure 7-30).  It was possible to use the
15     observed declines in PM3 5 following cessation of smoking to calculate  an effective air
16     exchange rate and thus a source strength for PM3 5 emissions from cigarettes. Four
17     estimates gave an average value of about 6 mg/cigarette, somewhat below the chamber study
18     estimates of 10 to 15 mg/cig.  An estimate due to Repace and  Lowrey (1980) of
19     concentrations of respirable particulates due to smoking was also tested, with good
20     agreement.  The Repace and Lowrey equation is
21
22                                         C = 25.6 Pala
23
24     where  Pa is smoking occupancy in  persons per 1,000  square feet and a is the air exchange
25     rate.  The equation was developed  assuming  1/3 of the occupants are smokers who smoke
26     two cigarettes per hour.  Assuming a background concentration of 15 /xg/m3, the estimates
27     for the smoking lounge for 0, 3, and 9 smokers were  10, 78, and 284 /xg/m3, respectively.
28     Repace's equation predicts 0, 99, and 296 /xg/m3, respectively. In two of the homes for the
29     elderly, apartments with smokers and nonsmokers were measured for three  consecutive days
30     using the NBS samplers.  In one building, the smoker's apartment had  a 3-day PM3 average
       April 1995                               7-112      DRAFT-DO NOT QUOTE OR CITE

-------
           100
                          Home #1
                                                  Home #2
                       Smoker's Room
                                             Nonsmoker's Room
      Figure 7-30.  Respirable particles in smoking and non-smoking areas of homes for the
                 elderly (arithmetic mean for 72 h).
1
2
3
4
of 39 £ig/m3, compared to 9.2 /xg/m3 in the nonsmoker's apartment; in the other home for
the elderly, where two smokers shared one apartment, the average 3-day PM3 concentration
was 89 /ig/m3 compared to 8.6 mg/m3 in the nonsmoking apartment.
     April 1995
                                  7-113
DRAFT-DO NOT QUOTE OR CITE

-------
 1     7.6.2.3  Studies in locations other than homes and buildings
 2          Nagda et al. (1990) measured RSP in aircraft cabins using both gravimetric and optical
 3     methods.  Although the methods did not agree well, they were averaged to produce an
 4     estimate of 75 /ig/m3 in the smoking sections, 54 /xg/m3 in the several "boundary rows"
 5     bordering the smoking sections, and 31 to 35 jig/m3 in the middle and remote seats.
 6     Average concentrations on nonsmoking flights were 35 to 40 jug/m3. Nicotine concentrations
 7     were 13.4 pig/m3 in the smoking section, but very low in all other sections (0.04 to 0.26 in
 8     nonsmoking sections of smoking flights, 0.00 to 0.08 in all areas  on nonsmoking flights).
 9          Oldaker et al.  (1990) measured PM3 5 and nicotine in 33 restaurants  in the Winston-
10     Salem  area during the summer of 1986 and the winter of 1988 to 1989.  (In the winter
11     season, the cutpoint was changed to PM2 5.) A wide range of particle concentrations was
12     noted,  from 18 to 1,374 /xg/m3 in the summer, and  <25 to 281 /ig/m3 in  winter.  Nicotine
13     concentrations also ranged over wide intervals, from 0.9 to 25.6 /xg/m3 in the summer,  and
14      <0.1 to 35.2 jug/m3 in the winter.
15          Lowrey et al.  (1993) measured PM3 5 using a TSI Piezobalance in a  number of outdoor
16     and indoor locations in Budapest.  Outdoor concentrations  measured over 6 to  16-min periods
17     ranged from 28 to 150 /ig/m3, with three of the four values above 100 /ig/m3 associated with
18     roadways (tram or bus stops). Indoor concentrations in seven areas without visible smoking
19     ranged from 42 to 100 /zg/m3; in 19 areas with active  smoking the range was from 56 to
20     650 ^g/m3.
21
22     7.6.3   Indoor air quality models and supporting experiments
23          Indoor concentrations of particles are a function of penetration of outdoor particles and
24     generation of particles indoors.  The concentrations are modified  by air exchange rates and
25     decay rates of the particles onto indoor surfaces.
26
27     7.6.3.1  Mass Balance Models
28          Mass balance models have been used for more than a century  in various branches  of
29     science.  All such models depend on the law of the conservation of mass.  They simply state
30     that the change in mass of a chemically inert substance in a given volume  is equal to the
31     amount of mass entering that  volume minus the amount leaving the  volume. Usually they

       April 1995                               7-114      DRAFT-DO NOT QUOTE OR CITE

-------
  1      are written in the form of first-order linear differential equations.  That is, consider a volume
  2      V filled with a gas of mass  m. The change in mass Am over a small time At will simply be
  3      the difference between the mass entering the volume (mgain) and the mass leaving the volume
  4      fatoJ:
  5

                                         Am = mgain - mloSS                            (7'8)
  6
  7
  8      Taking the limit as At approaches zero, we have the differential equation for the rate of
  9      change of the mass:
10
                                      dm/dt  = d/dt(mgain-mloss)                          (7-9)

11
12
13          If we require that the  mass  be uniformly distributed throughout the volume at all times,
14      we have  a condition that the physical chemists call "well-mixed".  We assume that any mass
15      gained or lost in the volume V is instantaneously distributed evenly throughout the volume.
16      We may  then replace the mass terms by the concentration C = m/V:
17
18                                   VdC/dt = d/dt (mgain - m[oj                        (7-10)
19
20          The above equation is the basis for all such mass-balance models. It takes on many
21      forms depending on the type of processes involved in transporting mass into or out of the
22      volume being considered.  A large class of models assume that the volume is a single
23      compartment. More complex models assume multiple compartments.  As an example of a
24      single compartment model,  we may consider a room of volume  V that exchanges air with the
25      outside at a constant flow rate Q.  We also assume that a mass of gas has been released in
26      the room at time t — 0, and that the outdoor concentration of this gas is 0.  (This is the
27      situation, for example,  when a tracer gas such as SF6 is released to determine the air

        April 1995                              7-115      DRAFT-DO NOT QUOTE OR CITE

-------
 1      exchange a.)  In this case, the gain in mass mgain is zero and the loss in mass is equal to the
 2      flow rate Q out of the house times the concentration C, so that equation (7-10) becomes:
 3
 4                                          VdC/dt = -aVC                               (7-11)
 5      Integrating this equation by separation of variables, we have
 6
 7                                           C = C0 eat.                                  (7-12)
 8
 9      where C0 is the concentration at time / = 0, and a =  Q/V is the air exchange rate.
10           Thus we find that the original concentration of tracer in the room decays with a time
11      constant a: the air exchange rate.
12           For a nonreactive gas with a nonzero outdoor concentration (e.g., carbon monoxide),
13      the mass balance equation takes the form
14
15                                       dCfr/dt = a(COM - Cin)                            (7-13)
16
17      where Cin is the indoor concentration, and cout is the outdoor concentration.
18
19           Depending on the variation with time of Cout, this equation has a number of solutions.
20      If Cout is constant, for example, and the initial indoor  concentration is zero, then the indoor
21      concentration rises at a rate determined by the air exchange rate to approach an asymptotic
22      value equal to the outdoor concentration:
23
24                                         Cin = Cout(l - eat)                              (7-14)
25
26           An early effort at developing an indoor air quality model was made by Shair and
27      Heitner  (1974).  This was a mass balance model in which the building was represented as a
28      single well-mixed chamber.  A single first-order linear differential equation represented the
29      change in mass of a pollutant due to infiltration, exfiltration, recirculation, source generation,
30      and removal due to filters in the circulation system or deposition on surfaces. Shair and
        April 1995                                7-116      DRAFT-DO NOT QUOTE OR CITE

-------
 1      Heitner (1974) solved the equation for certain simple inputs, such as a linearly increasing or
 2      decreasing outdoor concentration:
 3
 4                                          Cout = mt + b                               (7-15)
 5
 6           Since the outdoor concentration normally is a slowly-varying function, Shair and
 7      Heitner's linear approximation is actually an excellent approximation for time intervals of
 8      moderate length.
 9           If an indoor source 5(0 exists, it enters the mass balance model in the following way:
10
11                                  dCin/dt = a(Cout  - Cin) + S(t)/V                       (7-16)
12
13      where S(t) has the units of mass per unit time.
14           If the source has a constant generation rate (e.g., CO2 emissions from a person at rest),
15      then S(t) is a constant value 50 and the equation becomes
16
17                                   dC-Jdt = a(Cout - Cin) +  So/V                       (7-17)
18
19           If the substance of interest reacts or is adsorbed on surfaces while indoors, the equation
20      becomes
21
22                                dCin/dt = aCout - (a+k)Cin + S(t)/V                     (7-18)
23
24      where k represents the loss of the substance due to chemical reaction, adsorption on surfaces,
25      sedimentation, etc.  The decay rate k has the same units as the air exchange rate a (I/time);
26      their  sum (a+k) may be thought of as an effective air exchange rate.  The decay rate  k is
27      often used to apply to particles, which disappear faster indoors than a nonreactive gas such as
28      CO.  Since particles experience more difficulty than of a gas in penetrating the building
29      envelope, a penetration factor/(f <  1) is often applied that multiplies the outdoor
30      concentration in Equation 7-18 above.
        April 1995                                7-117      DRAFT-DO NOT QUOTE OR CITE

-------
 1          Alzona et al. (1979) applied the mass-balance equation with terms including a
 2     penetration coefficient/through the building envelope, adsorption on surfaces given by a rate
 3     k proportional to the indoor concentration Cin, and a resuspension rate R:
 4
 5                               dCin/dt = af Cout - aCin - kCin  + R.                     (7-19)
 6
 7     Setting dCin/dt = 0, the equilibrium solution is
 8
 9                                    Cin = (afCout + R)/(a+k)                         (7-20)
10
11     The equation was tested against a series of experiments in which elements known to be of
12     outdoor origin were collected under various experimental conditions and analyzed using
13     X-rays.  The authors concluded that/was of the order of 0.3 for many elements, and that
14     resuspension did not appear to be particularly important.  Measurement errors were fairly
15     large (15 to 25%) and limited their ability to estimate values of these parameters.
16          As described above, Koutrakis et al. (1991) used least-squares analysis to solve a
17     simplified form of the mass-balance model to determine source emission rates for particles
18     and elements due to cigarettes, woodsmoke, and kerosene heater use.  Koutrakis assumed a
19     value for k in order to solve the equation for/and the source emission rates.  Ozkaynak
20     et al. (1993) improved on Koutrakis'  approach by using least-squares analysis of the PTEAM
21     results to solve the equation simultaneously for k, / and source emission rates for PM2 5 and
22     PM10 particles and associated elements for smoking and for cooking.
23          Axley  and Lorenzetti (1991) developed an  indoor model using an element-assembly
24     computer language (STELLA) that is  capable of handling any number of compartments and
25     air flows between compartments.  The model is  based on an earlier model (CONTAM88)
26     developed under EPA sponsorship.
27          Sparks et al.  (1991) developed a more user-friendly version of the Axley model, with
28     menus prompting the user to insert the necessary parameters.
29          Traynor et al. (1989) developed a "macromodel" based on Monte Carlo simulations
30     using global input data such as house  volumes, air exchange rates,  and emissions from
31     combustion sources to assess residential concentrations of combustion-source pollutants such

       April 1995                               7-118      DRAFT-DO NOT QUOTE OR CITE

-------
  1      as CO, NO2, and respirable suspended particles. For a home with only one combustion
  2      source during winter in upstate New York, at an outdoor temperature of 0 °C ,  and an
  3      outdoor RSP geometric mean concentration of 19 /*g/m3, the model predicted geometric
  4      mean concentrations of about 80 jug/m3 in a home with smoking, 75 /zg/m3 for a radiant
  5      kerosene heater, about 60 /-ig/m3 for a convective unvented gas space heater and a non-
  6      airtight wood stove, and about 40 /ig/m3 for a radiant unvented gas space heater.  An airtight
  7      wood stove was predicted to produce a geometric mean about 30 ptg/m3. Gas ovens, dryers,
  8      hot water heaters,  boilers, and forced-air furnaces were predicted to result in low indoor
  9      concentrations of 10 to 15 /-ig/m3, unless the gas oven was used for heating, in which case
10      the predicted geometric mean was about 20 ^g/m3.
11           At present, one of the most complete forms of the mass-balance indoor air quality
12      model has been presented by Nazaroff and Cass (1989).  These authors developed the model
13      to allow for changes in particle size and chemical composition, including terms  for
14      homogenous turbulence, natural convection, thermophoresis,  advection, and Brownian
15      motion.  Coagulation of particles is also included.  The computer form of the model required
16      40 to 60 min of CPU time to determine an 11-h evolution of an aerosol mixture of 16
17      different sizes.  The model was partially validated by checking it against the results  of a
18      chamber study using cigarette-generated aerosol to determine the effectiveness of air cleaners
19      (Offermann et al., 1985).
20           A simplified  form of the model was employed in a study of indoor air soiling potential
21      in three California museums, two with new HVAC systems and one with only natural
22      ventilation (Nazaroff et al., 1990a).  Measured values of elemental carbon in fine (PM2)
23      particles were 0.63, 0.16, and 5.6 /xg/m3 in the three museums compared to model estimates
24      of 0.62 to  0.83, 0.22 to 0.23,  and 4.9 jug/m3-  The authors were able to predict the fate
25      (i.e., the main removal process and the rate of removal) of particles  of various sizes in the
26      three museums (Figure 7-31).  In the two newer museums, most particles below 10  pirn were
27      removed by filters; in the  older museum,  most PM10 particles were removed by ventilation.
28      In all three museums, the dominant fate of the larger particles was gravitational settling onto
29      upward-facing surfaces. The authors concluded that perceptible soiling would occur in less
30      than a year for the older museum, but would require  10 to 40 years for the newer museums.
       April 1995                               7-119      DRAFT-DO NOT QUOTE OR CITE

-------
                                    Norton Simon Museum
                                          *'*  •  i  H-
                                                  • Filtration
                                                    Ventilation
                                                    Deposition
                                                  x Coagulation (-)
                                                  + Coagulation (+)
                  0.001
                            0.1
                                   Particle Diameter (urn)


                                        Scott Gallery
                                        i—i  i i i r	i  • i  i  i i i i 11
                                                  • Filtration
                                                    Ventilation
                                                  • Deposition
                                                    Coagulation (-)
                                                  + Coagulation (+]
                                                    Ventilation
                                                  • Deposition
                                                  x Coagulation (-)
                                                  + Coagulation (+
                                   Particle Diameter (u,m)
Figure 7-31.  Predicted fate of particles penetrating into buildings of three California
              museums as a function of particle size.  The ordinate of each point

              represents the fraction of the mass that is removed by the indicated

              process.


Source: Nazaroff et al. (1990a).
April 1995
7-120       DRAFT-DO NOT QUOTE OR CITE

-------
  1      The hanging of pictures at a slight downward-facing angle was predicted to greatly decrease
  2      the rate of soiling.
  3           A crucial unknown parameter in the mass-balance model for particles is the rate of
  4      decay to surfaces. This rate of decay (k) may be expressed as the product of a deposition
  5      velocity kd with the surface to volume ratio in the room or building:
  6                                             k = kdS/V                                  (7-21)
  7
  8      The deposition velocity will vary with particle size.
  9           Both the Nazaroff study above and the series of studies by Weschler  and colleagues
 10      below have provided useful data on deposition velocities for important anions such as
 11      sulfates.
 12           A series  of studies, also concerned with the effects of  indoor particles on materials,
 13      were carried out by Weschler and colleagues at AT&T Bell Laboratories (Weschler et al.,
 14      1989; Sinclair et al., 1988,  1990, 1992).  Studies of buildings with low occupancy, large
 15      amounts of electronic equipment, and high-quality filtering and HVAC systems succeeded in
 16      determining deposition velocities for coarse particles and various fine particle ions. For
 17      coarse particles, these velocities were about equal to velocities predicted for gravitational
 18      settling, similar to the results of Nazaroff et al. (1990a) described above.   For fine particles,
 19      however, the deposition velocity was greater than that predicted for gravitational settling
20      alone.  For sulfates, the dominant anion in fine particles, deposition velocities at four
21      buildings in Wichita, Lubbock, Newark, and Neenah were 0.004, 0.005. 0.005, and
22      0.004 cm/s, respectively  (Sinclair et al., 1992).
23           Nazaroff et al. (1993) reviewed these and other studies of deposition  velocity.  The
24      authors pointed out that the studies by Weschler and colleagues and also one study in
25      Helsinki (Raunemaa et al., 1989)  had produced values of 0.003 to 0.005 cm/s for fine-mode
26      sulfate, but that studies by Nazaroff and colleagues (Ligocki et al.,  1990; Nazaroff et al.,
27      1990) resulted in much smaller values of 0.00002 to 0.001 cm/s. It is not  clear whether the
28      differences are due to the many differences in surface materials and filtration systems in the
29      different types of  buildings (museums versus telephone equipment buildings) or to the
30      different methods  of determining deposition velocities. However, the discrepancy  is clear
31      evidence that further work is needed.

        April 1995                               7-121      DRAFT-DO NOT  QUOTE OR CITE

-------
 1          Because all large-scale studies of indoor air quality agree that the most important indoor
 2     source of fine particles is tobacco smoke, a brief review of models and chamber studies
 3     dealing with ETS is provided here.
 4
 5     7.6.3.2  Models of ETS
 6          Using the mass balance approach, a number of efforts have been undertaken to model
 7     mathematically  the pollutant concentrations from tobacco smoke in indoor locations.  For
 8     example, Brief  (1960) proposed a simple graph to determine transient concentrations for
 9     pollutants in indoor settings that is based on an exponential decay as a function of time.
10     Turk (1963) proposed a general equation for calculating the concentrations in a chamber that
11     includes both exterior and interior sources, as well as the removal effect of pollutants by air
12     treatment systems.   Bridge and Corn (1972) reported that a solution to the equations
13     proposed by Turk (1963) adequately  predicts tobacco smoke in occupied spaces. Jones and
14     Pagan (1974) used Turk's equation to calculate carbon monoxide (CO) concentrations versus
15     time from cigarette smoke  in an office building and a single-family dwelling.  Ishizu (1980)
16     examined experimentally the  inclusion of a mixing factor in these models, and Repace and
17     Lowrey (1980) developed a modification of the Turk equation incorporating a mixing factor.
18     They model the concentration as a function of time (assuming an initial concentration of
19     zero) as :
20
                                     C = CM [(l-exp(-(a+k)mt)]                        (7-22)

21
22     where
23     CQO  = G/Vm (a+k) is the equilibrium concentration of ETS  particles;
24     m = a "mixing factor" to account for imperfect mixing'
25     G = the generation rate, a function of the number of cigarettes being smoked and the total
26     particle mass emitted from sidestream and exhaled mainstream smoke.
27           Using an  estimated value for G of 24  mg/cigarette, Repace (1978b) arrived at an
28     estimate for the steady-rate concentration of PM3 5 (in /ng/m3) due to smoking:
29

       April  1995                               7-122      DRAFT-DO NOT QUOTE  OR CITE

-------
  1                                        PM35 = 217Dhs/a                             (7-23)
  2
  3      where Dhs is the density of "habitual smokers" (i.e., those who smoke two cigarettes per
  4      hour) per 100 m3, and the value of 217 incorporates a mixing factor and a particle decay rate
  5      based on empirical observations in a number of locations.
  6           As reported in Repace and Lowrey (1982), equation 7-23 was partially validated by a
  7      chamber  experiment reported later by Leaderer et al. (1984), in which the measured
  8      equilibrium concentration was 620 /ng/m3 compared to 478 /^g/m3 predicted by the equation,
  9      a difference  of about 23 %.  Since equation 7-23 assumes a nation-wide sales-weighted
 10      average tar content, a small experiment using one or two brands of cigarettes would be
 11      expected  to give somewhat  different values (sse discussion on page  7-49).
 12           The concentrations of pollutants from ETS in a large mixing volume, such as a room,
 13      have been observed to increase once a cigarette starts to burn and to decay exponentially
 14      once the cigarette is put out (Brief,  1960; Ishizu, 1980; Repace and Lowrey,  1980,  1982;
 15      Leaderer  et al., 1984; Repace, 1987).  These exponential  functions  are solutions to the mass
 16      balance equation for the case  of a source that emits at a fixed rate when it is on—and at zero
 17      rate when it  is off—with a fixed air  exchange rate.  This source can be viewed as a
 18      "rectangular" input "tune series" (concentration as a function of time) to the mass balance
 19      model.
20           Smokers ordinarily engage in a sequential smoking "activity pattern" over time: one
21      cigarette is smoked after another, with a recovery period between each cigarette.  A person
22      in a room with a smoker (an office,  an automobile, a smoking lounge, a restaurant) is
23      exposed to a time series of  concentrations resulting from a succession of cigarettes reflecting
24      the smoking  activity patterns of the smoker.
25           In recent research, the basic mass-balance model was adapted  to the case of a sequence
26      of cigarettes  smoked one after another, and its effectiveness in predicting the pollutant
27      concentrations as a function of time  (the  "time series" of concentrations) was  tested using
28      real-time  monitoring instruments (Ott et al., 1992). This work has developed a model for
29      computing the time series of pollutants generated by sequential cigarette smoking activity
30      patterns, the  Sequential Cigarette Exposure Model (SCEM).  If tj is the time at which
        April 1995                               7-123       DRAFT-DO NOT QUOTE OR CITE

-------
 1      cigarette i begins, and S; is the duration of cigarette i, then during the time the cigarette is
 2      being smoked, the particle concentration rises according to:
 3
                               S(l-exp(-(a+k)(t-ti)))                             (
                          C  =  	     u^	1—  +C2i_2exp(-(a+k)(t-tj))      C
 4
 5     whereas between cigarettes the concentration declines:
 6
                                    C = C^expHa+kXt-Sj-tj))                    (7-25)
 7
 8
 9     In these equations, C2i_2 = C(f,) and C2i.! = C(f; + ^).
10          A computer program for the SCEM has been written in Microsoft QuickBASIC Version
11     4.5 programming language to apply to equations 7-24 and 7-25. This research also has
12     derived theoretical equations for the minimum, maximum,  and mean pollutant concentration
13     in a well-mixed microenvironment for any cigarette smoking activity pattern. General
14     expressions also have been derived for the case  of the habitual smoker (uniform cigarette
15     duration and  same time between cigarettes)  and  for the case of multiple habitual smokers.
16          The  equations used to derive the SCEM are general and are consistent with earlier ETS
17     indoor air quality models that were derived for special cases (for example, Repace and
18     Lowrey,  1980). Repace (1987), for example, described  a person with uniform smoking
19     activity (a constant rate of smoking per unit time) as an  "habitual smoker."  He considers the
20     special case in which the habitual smoker smokes two cigarettes per hour, which is based on
21     a national average smoking rate.  The SCEM considers the general case in which each
22     habitual smoker can have  any smoking rate, and concentration is measured on a  "real-tune"
23     (that is, continuous)  basis.
24          Solutions to the mass balance equation provide a theoretical basis for calculating all
25     parameters of the model—air exchange rate, source strength, and sink removal terms—in a
26     single experiment.  Because of the SCEM's fine time resolution, experiments to validate the
27     model require monitoring instruments that operate with fine tune resolution (minutes or
28     seconds).   The air exchange rate is determined from the exponential decay of concentrations

       April 1995                              7-124       DRAFT-DO NOT QUOTE OR CITE

-------
  1      seconds).  The air exchange rate is determined from the exponential decay of concentrations
  2      in the microenvironment (Ott et al., 1992).  The source strength is determined from the
  3      equilibrium concentration with continuous smoking.  The sink removal term for pollutants
  4      that adhere to surfaces, such as particles, is determined by subtracting the particle decay rate
  5      from the decay rate for a pollutant that has no surface sinks, such as CO.
  6           Klepeis et al. (1995) applied an extension of the SCEM model (now called the Multiple
  7      Cigarette Exposure Model, or MCEM) to smoking lounges at the San Francisco, CA and San
  8      Jose, CA airports.  Three Piezobalances were placed at each end and the center of the
  9      lounges.  One investigator counted the number of lit cigarettes every minute while the other
 10      took readings from the Piezobalances. A Langan CO monitor took carbon monoxide
 11      readings continuously and logged them automatically.   Five visits to each airport were made.
 12      During the fifth visit to each airport,  an experiment was done to calculate the air exchange
 13      rate when there were few or no smokers present: several cigars were smoked and the decay
 14      of the CO level was measured, giving air exchange rates of 10.7 and 13.0 ach at the two
 15      airports.  The calculated PM3 5 source strengths for cigarettes during these two visits were
 16      identical at 1,340 ptg/min.  Air exchange rates were not determined at the other visits, but
 17      assuming the same rates resulted in an average PM3 5 emission rate of 1,450 /xg/min.  The
 18      decay rate of the ETS-related PM3 5 was estimated to be 0.048 and 0.034 min"1, or 2.88 and
 19      2.04 h-1.
 20           Ott et al. (1995a) tested the model in a tavern before and after smoking was prohibited.
 21      During 26 visits over a period of two years while smoking was  allowed, indoor
 22      concentrations averaged 56.9 /xg/m3 above  outdoor concentrations, compared to 5.9 /ig/m3
 23      above outdoor levels on 24 visits in the first six weeks  after smoking was prohibited.
 24      A second set of follow-up visits (matched by time of day, day of week, and season to the
 25      earlier visits) yielded an average concentration 13.1 pig/m3 above outdoor levels.  Using
26      cigarette emission rates from the literature, the measured tavern volume of 521 m3, and a
27      measured air exchange rate under  "typical" conditions,  a mass-balance model predicted 42.5
28      jug/m3 for  an average "continuous  smoking" count of 1.17 cigarettes, comparing favorably
29      with the observed average of 43.8 /ig/m3.
30          Ott et al. (1995b) used Laplace transforms to apply the model to an experiment in
31      which three Kentucky reference 2R1 cigarettes were smoked one after another in a 25.7 m3

        April 1995                              7-125      DRAFT-DO NOT QUOTE OR CITE

-------
 1     bedroom.  Resulting PM3 5 concentrations were measured in the bedroom and living room
 2     with a Piezobalance.  Peak values reached 5,500 /*g/m3, and about 2,000 /xg/m3 in the living
 3     room.  The living room window was opened 2 h later, but concentrations were  still in excess
 4     of 100 /ig/m3 after 4 h.
 5
 6     7.6.4  Summary and Conclusions
 7          At low outdoor levels of fine (PM3 5 or PM2 5) particles  (as in most of the cities in the
 8     Six-City and New York State studies),  mean indoor concentrations have been found to be
 9     twice as high as outdoor levels.  At high outdoor levels, mean indoor concentrations have
10     been about 10% lower than the mean outdoor concentrations in the two areas studied
11     (Steubenville and Riverside). Indoor concentrations are considerably higher during the day,
12     when people are active, than at night.  Based on a mass-balance model, outdoor air was the
13     major source of indoor particles in the PTEAM study, providing about 3/4 of fine particles
14     (PM2 5) and 2/3 of inhalable particles (PM10) in the average home.
15          The three largest studies of indoor air particles in U.S. homes have all found that the
16     single largest indoor source of fine particles is cigarette smoke,  for homes with smokers.
17     (EPA's NHAPS data show that 31% of U.S. homes have a smoker, down from 50 to 60% in
18     years past).  Estimates  of the impact of a smoking home range from about 30 to 45 /ig/m3,
19     and a of a single cigarette from  1 to 2 /ig/m3 for a 24-h period. Homes without smoking
20     have indoor particle concentrations (both PM10 and  PM2 5) that are sometimes below and
21     sometimes above the outdoor levels. At low outdoor levels (as in most of the cities in the 6-
22     City and New York State studies) indoor concentrations are generally higher—at high
23     outdoor levels, they are slightly lower.  Indoor concentrations are considerably  higher during
24     the day, when people are active, than at night.
25          The second largest identified indoor source of particles,  as determined by  the PTEAM
26     Study and several smaller studies, is cooking.  Slightly less than half of the PTEAM homes
27     reported cooking on the day they were monitored.   Estimates  of the effect of cooking ranged
28     from about 10 to 20 ^ig/m3.  A  few small studies confirm the effect of cooking on indoor
29     particle levels,  both PM10 and PM2 5.  The two other large-scale studies did not directly test
30     for the effect of cooking, although the residual indoor concentrations in multivariate
31     calculations led the authors to suggest that cooking could have contributed to the  residual.

       April 1995                               7-126      DRAFT-DO NOT QUOTE OR CITE

-------
 1          Kerosene heater use was determined to contribute about  15 jug/m3 to indoor
 2     concentrations in one county but not the other in the New York State study.  Also, a later
 3     effort using mass-balance calculations was unsuccessful in determining a contribution to
 4     particle mass from kerosene heater use in either county, although a somewhat smaller set of
 5     homes may have been responsible for this result.  Gas stoves, wood stoves, and fireplaces
 6     were found to have no noticeable impact on total concentrations of particles, although many
 7     studies show an increase in PAH concentrations and some show an increase in mutagenicity
 8     of indoor air due to these combustion sources.
 9          Vacuuming, dusting, and sweeping were found to contribute slightly but with doubtful
10     significance to indoor levels in the PTEAM Study. House volume had a significant but small
11     effect on particle concentrations, with values of -1 to  -2 /xg/m3 per 1,000 cubic feet.  Air
12     exchange rates were also significant at times, but with different impacts depending on the
13     relative indoor and outdoor concentrations—at high outdoor concentrations,  increased air
14     exchange resulted in increases in the indoor air particle levels.
15          Unknown indoor sources were found to account for a substantial fraction (25%) of
16     indoor concentrations (both PM2 5 and PM10) in the PTEAM study.  This suggests a need for
17     further research to determine the source or sources of these particles.
18          Decay rates for fine (PM2 5) particles were determined to be about 0.4 h"1  compared to
19     1 h"1 for coarse particles, with an intermediate estimate of 0.65 h"1 for PM10. For a home
20     with no indoor sources whatever and a typical air exchange rate of about 0.75 h-1, this
21     would imply that fine particles indoors would be about 0.757(0.4+0.75) =  65% of the
22     outdoor value at equilibrium, indoor PM10  would be about 54% of outdoor levels, and indoor
23     coarse particles would be about 43% of outdoor  levels.  Since very few homes were
24     observed to have concentrations  this low, it can be inferred that very few homes are  free of
25     important indoor sources of particles.
26          Studies in buildings also indicated that smoking was the  major indoor  source of
27     particles, with a geometric mean of 44 versus 15 jig/m3 (arithmetic mean of 70  versus
28     18 jiig/m3) observed  for smoking versus nonsmoking areas in 38 Pacific Northwest buildings.
29     This difference of 29 to 52 pig/m3 is similar to the difference of 30 to 45 pig/m3 estimated
30     from the three major studies of U.S. homes.
        April 1995                               7-127      DRAFT-DO NOT QUOTE OR CITE

-------
 1          Indoor air quality models have been employed with increasing success to estimate
 2     source emission rates and particle decay rates.  Cigarettes in homes with normal activities
 3     appear to emit about 14 mg/cigarette, a result that agrees well with various chamber studies
 4     using smokers or smoking machines. Cooking  was  estimated to emit 4 mg/min, a result that
 5     needs confirmation by other studies. Elemental emission profiles have been determined for
 6     both smoking and cooking, with potassium and  chloride being major markers for smoking,
 7     and iron and calcium for cooking.  Particle decay rates have been estimated for homes to
 8     range between 0.4 and 1.0 h"1.  Studies in telephone equipment buildings and museums have
 9     established particle deposition velocities for sulfates  and other ions, although differences in
10     the estimates suggest that further research is needed.
11
       April 1995                              7-128      DRAFT-DO NOT QUOTE OR CITE

-------
 1     REFERENCES

 2     Alzona, J.; Cohen, B. L.; Rudolph, H.; Jow, H. N.; Frohliger, J. O. (1979) Indoor-outdoor relationships for
 3             airborne particulate matter of outdoor origin. Atmos. Environ.  13: 55-60.

 4     Anuszewski, J.; Larson, T. V.; Koenig, J. Q. (1992) Simultaneous indoor and outdoor particle light scattering
 5             measurements at nine homes using a portable nephelometer. Presented at: meeting of the American
 6             Association for Aerosol Research; paper no. 3A.5.

 7     Axley, J. W.;  Lorenzetti, D. (1991) IAQ modeling using STELLA™:  a tutorial introduction building technology
 8             program. Cambridge, MA: Massachusetts Institute of Technology.

 9     Baughman, A. V.; Gadgil, A. J.; Nazaroff, W. W. (1994) Mixing of a point source pollutant by natural
10             convection flow within a room. Indoor Air 4:  114-122.

11     Berg, I.; Overvik,  E.; Nord, C.-E.; Gustafsson, J.-A. (1988) Mutagenic activity in smoke formed during broiling
12             of lean pork at  200, 250, and 300 °C. Mutat.  Res. 207: 199-204.

13     Binder, R.  E.; Mitchell, C.  A.; Hosein, H. R.; Bouhuys, A. (1976) Importance of the indoor environment in air
14             pollution exposure. Arch. Environ.  Health 31: 277-279.

15     Bridge, D.  P.; Corn, M. (1972) Contribution to the assessment of exposure of nonsmokers to air pollution from
16             cigarette and cigar smoke in occupied spaces. Environ. Res. 5: 192-209.

17     Brief, R. S. (1960) A simple way to determine air contaminants. Air Eng. 2: 39-51.

18     Chan, C.-C.; Huang, S.-K.; Chen, Y.-C.; Wang, J.-D. (1995) Personal, indoor, and outdoor concentrations of
19             respirable suspended particulates (RSP) and nicotine in seven smoker's families in Taiwan. Atmos.
20             Environ.:  submitted.

21     Chapin, F.  S., Jr. (1974) Human activity patterns in the city. New York, NY: Wiley-Interscience Publishers.

22     Clayton, C. A.; Pellizzari, E. D.; Wiener, R. W. (1991) Use of a pilot study for designing a large-scale
23             probability study of personal exposure to aerosols. J. Exposure Anal. Environ. Epidemiol.  1: 407-422.

24     Clayton, C. A.; Perritt, R. L.; Pellizzari, E. D.; Thomas, K. W.; Whitmore, R. W.; Ozkaynak, H.; Spengler,
25             J. D.;  Wallace, L. A. (1993) Particle total exposure assessment methodology (PTEAM) study:
26             distributions of aerosol and elemental  concentrations in personal, indoor, and outdoor air samples in a
27             southern California community. J. Exposure Anal. Environ.  Epidemiol. 3: 227-250.

28     Cohen,  B. S.;  Chang, A. E.; Harley, N. H.;  Lippmann, M.  (1982) Exposure estimates from personal lapel
29             monitors.  Am.  Ind.  Hyg. Assoc. J.  43: 239-243.

30     Cohen,  B. S.;  Harley, N. H.; Lippmann, M.  (1984) Bias in air sampling techniques used to measure inhalation
31             exposure.  Am.  Ind.  Hyg. Assoc. J.  45:  187-192.

32     Colome, S. D.; Kado, N. Y.; Jacques,  P.; Kleinman,  M. (1990) Indoor-outdoor relationships of particles less
33             than 10 /*m in aerodynamic diameter (PM10) in homes of asthmatics. In: Indoor  air '90: precedings of
34             the 5th international conference on indoor air quality and climate, volume 2, characteristics of indoor air;
35             July-August;  Toronto, ON, Canada. Ottawa, ON, Canada: International Conference on Indoor Air
36             Quality and Climate, Inc.; pp. 275-280.
         April 1995                                    7-129      DRAFT-DO NOT QUOTE- OR CITE

-------
 1     Colome, S. D.; Kado, N. Y.; Jaques, P.; Kleinman, M. (1992) Indoor-outdoor air pollution relations: paniculate
 2             matter less than 10 /mi in aerodynamic diameter (PM10) in homes of asthmatics. Atmos. Environ. Part A
 3             26: 2173-2178.

 4     Commission of the European Communities. (1992) Exposure assessment report number  1, COST 613/2.
 5             Luxembourg: Commission of the European Communities; report no. EUR 14345 EN. (Series on air
 6             pollution epidemiology).

 7     Coultas, D. B.; Samet, J. M.; McCarthy, J. F.; Spengler, J. D. (1990) Variability of measures of exposure to
 8             environmental tobacco smoke in the home. Am. Rev. Respir. Dis. 142: 602-606.

 9     Cupitt, L. T.;  Glen, W. G.; Lewtas, J. (1994) Exposure and risk from ambient particle-bound pollution in an
10             airshed dominated by residential wood combustion and  mobile sources. In: Symposium of risk assessment
11             of urban air: emissions, exposure, risk identification, and risk quantitation; May-June 1992;  Stockholm,
12             Sweden. Environ. Health Perspect.  102(suppl. 4):75-84.

13     Daisey, J. M.; Spengler, J. D.; Kaarakka,  P.  (1987) A comparison of the organic chemical composition of
14             indoor aerosols during woodburning and non-woodburning periods. In: Seifert, B.; Esdorn, H.; Fischer,
15             M.; Rueden, H.; Wegner, J., eds.  Indoor air '87: proceedings  of the 4th international conference on
16             indoor air quality and climate, v. 1, volatile organic compounds, combustion gases, particles and fibres,
17             microbiological agents;  August; Berlin, Federal Republic of Germany.  Berlin, Federal Republic of
18             Germany: Institute for Water, Soil and Air Hygiene; pp. 215-219.

19     Diemel, J. A.  L.; Brunekreef, B.; Boleij, J. S. M.; Biersteker, K.; Veenstra, S. J. (1981) The Arnhem lead
20             study:  II, indoor pollution, and indoor/outdoor relationships. Environ. Res. 25:  449-456.

21     Dietz, R. N.; Cote, E. A.  (1982) Air infiltration measurements in a home using a convenient perfluorocarbon
22             tracer technique. Environ. Int. 8: 419-433.

23     Dockery, D. W.; Spengler, J. D. (1981a) Indoor-outdoor relationships of respirable sulfates and particles.
24             Atmos. Environ. 15:  335-343.

25     Dockery, D. W.; Spengler, J. D. (1981b) Personal exposure to respirable  particulates and sulfates. J. Air Pollut.
26             Control Assoc. 31: 153-159.

27     Dockery, D. W.; Schwartz, J.; Spengler, J. D. (1992) Air pollution and daily  mortality: associations with
28             particulates and acid aerosols. Environ. Res. 59: 362-373.

29     Duan, N. (1982) Models for human exposure  to air pollution. Environ. Int. 8:  305-309.

30     Duan, N.; Mage, D. T. (1995) Combination of direct and indirect approaches for exposure assessment. J.
31             Exposure Anal. Environ. Epidemiol.:  accepted.

32     Federal Register. (1992) Guidelines for exposure assessment. F. R. (May 29) 57:  22887-22938.

33     Fletcher, R. A.; Bright, D. S. (1983) NBS portable ambient paniculate sampler.  Washington, DC: U.S.
34             Department of Commerce, National Bureau of Standards; report no. NBSIR 82-2561. Available from:
35             NTIS, Springfield, VA;  PB83-165019.

36     Fletcher, B.; Johnson, A. E. (1988) Comparison of personal and area concentration measurements and the use of
37             a manikin in sampling.  In: Vincent, J. H., ed. Ventilation '88. New York, NY: Pergamon Press; pp.
38             161-166.
         April  1995                                    7-130       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Fugas , M.; Wilder, B.; Paukovic, R.; Hrsak, J.; Steiner-Skreb, D. (1973) Concentration levels and paniculate
  2            size distribution of lead in the air of an urban and an industrial area as a basis for the calculation of
  3            population exposure. In: Barth,  D.; Berlin, A.; Engel, R.; Recht, P.;  Smeets, J., eds. Environmental
  4            health aspects of lead:  proceedings [of an] international symposium; October 1972;  Amsterdam, The
  5            Netherlands. Luxembourg: Commission of the European Communities; pp. 961-968; report no. EUR
  6            5004 d-e-f.

  7     Gray, M. I.; Unwin, J.; Walsh, P. T.; Worsell, N.  (1992) Factors influencing personal exposure to gas and dust
  8            in workplace air: application of a visualization technique. Safety Sci. 15: 273-282.

  9     Groves, W. A.; Hahne, R. M. A.; Levine, S. P.; Schork, M. A. (1994) A field comparison of respirable dust
 10            samplers. Am. Ind. Hyg. Assoc. J. 55: 748-755.

 11     Guerin, M. R.; Jenkins, R. A.; Tomkins, B.  A. (1992) The chemistry of environmental tobacco smoke:
 12            composition and measurement. Chelsea, MI:  Lewis Publishers.

 13     Hammond, S. K.; Leaderer, B. P.; Roche, A. C.; Schenker, M. (1986) A method to measure exposure to
 14            passive smoking. In: Proceedings of the 1986 EPA/APCA symposium on measurement of toxic air
 15            pollutants; April; Raleigh, NC. Research Triangle Park, NC: U.S. Environmental Protection Agency,
 16            Environmental Monitoring Systems Laboratory; pp. 16-24; EPA report no. EPA-600/9-86-013. Available
 17            from:  NTIS, Springfield, VA; PB87-182713.

 18     Highsmith, V. R.; Rodes, C. E.; Hardy, R. J.  (1988) Indoor particle concentrations associated with use of tap
 19            water  in portable humidifiers. Environ. Sci.  Technol. 22: 1109-1112.

 20     Highsmith, V. R.; Zweidinger, R. B.; Merrill, R. G. (1988) Characterization of indoor and outdoor air
 21            associated with residences using woodstoves: a pilot study. Environ. Int. 14: 213-219.

 22     Highsmith, V. R.; Hoffman, A. J.; Zweidinger, R. B.; Cupitt, L.  T.; Walsh, D.  (1991) The IACP: overview  of
 23            the Boise, Idaho and the Roanoke, Virginia field studies. Presented at: the 84th annual meeting of the Air
 24            & Waste  Management Association; June; Vancouver, BC, Canada. Pittsburgh, PA:  Air & Waste
 25            Management Association; paper no. 91-131.1.

 26     Hodges, S. D.; Moore, P. G. (1977)  Data uncertainties and least square regression. Appl. Stat.  44: 185-195.

 27     Holcomb, L. C. (1993) Indoor air quality and environmental tobacco smoke: concentration and exposure.
 28            Environ. Int. 19: 9-40.

 29     Ingebrethsen, B. J.; Heavner, D. L.;  Angel, A. L.; Conner, J. M.; Steichen, T. J.; Green,  C. R.  (1988) A
 30            comparative study  of environmental tobacco smoke paniculate mass  measurements in an environmental
 31             chamber.  JAPCA 38: 413-417.

 32     Ishizu, Y. (1980) General equation for the estimation of indoor pollution. Environ. Sci. Technol. 14: 1254-1257.

 33      Jenkins, P. L.; Phillips, T. J.; Mulberg, E. J.; Hui,  S. P. (1992) Activity patterns of Californians: use of and
 34             proximity to indoor pollutant sources. Atmos. Environ. Part A 26: 2141-2148.

 35      Jones, R. M.;  Pagan, R. (1974) Application of mathematical model for the buildup of carbon monoxide from
36             cigarette smoking in rooms and houses. ASHRAE J. 16: 49-53.

37      Kamens, R.; Lee, C.-T.; Wiener, R.; Leith, D. (1991) A study to characterize indoor particles in three non-
38             smoking homes. Atmos. Environ. Part A 25:  929-948.
         April 1995                                    7_131        DRAFT-DO NOT QUOTE OR CITE

-------
 1     Kim, Y. S.; Stock, T. H. (1986) House-specific characterization of indoor and outdoor aerosols. In: Berglund,
 2             B.; Berglund, U.; Lindvall, T.; Spengler, J.; Sundell, J., eds. Indoor air quality: papers from the third
 3             international conference on indoor air quality and climate; August 1984; Stockholm, Sweden. Environ.
 4             Int.  12: 75-92.

 5     Klepeis, N.  E.; Ott, W. R.; Switzer, P. (1994) A total human exposure model (THEM) for respirable suspended
 6             particles (RSP). Presented  at: the 87th annual meeting of the  Air & Waste Management Association;
 7             June; Cincinnati, OH. Pittsburgh, PA: Air & Waste Management Association; paper no. 94-WA75A.03.

 8     Kleipis, N. E.; Ott, W. R.; Switzer, P. (1995) Modeling the time series of respirable suspended particles and
 9             carbon monoxide from multiple smokers: validation in two public smoking lounges. Presented at: the
10             88th annual meeting and exhibition of the Air & Waste Management Association; June. Pittsburgh, PA:
11             Air  & Wsate Management Association; paper no. A-1233.

12     Koutrakis, P.; Briggs, S. L. K.; Leaderer,  B. P. (1992)  Source apportionment of indoor aerosols in Suffolk and
13             Onondaga Counties, New York. Environ. Sci. Technol.  26: 521-527.

14     Langan, L.  (1992) Portability in measuring exposure to carbon monoxide. J. Exposure Anal. Environ. Epidemiol.
15             2(suppl. 1): 232-239.

16     Last, J. M., ed. (1983) A dictionary of epidemiology. New York, NY: Oxford University Press; p. 43

17     Last, J. M., ed. (1988) A dictionary of epidemiology. 2nd ed. New York, NY: Oxford University Pressjp. 56.

18     Leaderer, B. P.; Hammond, S. K. (1991) Evaluation of  vapor-phase  nicotine and respirable suspended particle
19             mass as markers for environmental tobacco smoke. Environ.  Sci. Technol. 25: 770-777.

20     Leaderer, B. P.; Cain,  W. S.; Isseroff, R.; Berglund, L. G.  (1984) Ventilation requirements in buildings—II.
21             paniculate matter and carbon monoxide from cigarette smoking. Atmos. Environ.  18: 99-106.

22     Leaderer, B.; Koutrakis, P.; Briggs, S.; Rizzuto, J.  (1990) Impact of indoor sources on residential aerosol
23             concentrations. In: Indoor air '90: precedings of the 5th international conference on indoor air quality and
24             climate, volume 2, characteristics of indoor  air;  July-August; Toronto, ON, Canada. Ottawa, ON,
25             Canada: International  Conference on Indoor Air  Quality and  Climate, Inc.; pp. 269-274.

26     Lebret, E. (1985) Air pollution in Dutch homes: an  exploratory  study in environmental epidemiology.
27             Wageningen, The Netherlands: Department of Air Pollution, Department of Environmental and Tropical
28             Health; report R-138; report 1985-221.

29     Lebret, E.;  McCarthy,  J.;  Spengler, J.; Chang, B.-H. (1987) Elemental composition of indoor fine particles. In:
30             Seifert, B.; Esdorn, H.; Fischer, M.; Rueden, H.; Wegner, J., eds. Indoor air '87: proceedings of the
31             4th  international conference on indoor air quality and climate, v.  1, volatile organic compounds,
32             combustion gases,  particles and fibres, microbiological agents; August; Berlin, Federal  Republic of
33             Germany. Berlin, Federal  Republic of Germany: Institute for Water,  Soil and Air Hygiene; pp. 569-574.

34     Lebret, E.;  Boleij, J.; Brunekreef, B. (1990) Environmental tobacco  smoke in Dutch homes. In: Indoor air '90:
35             precedings of the 5th  international conference on indoor air quality and climate, volume 2, characteristics
36             of indoor air; July-August; Toronto, ON, Canada.  Ottawa, ON, Canada: International Conference on
37             Indoor Air Quality and Climate, Inc.; pp. 263-268.

38     Lehmann, E.; Rentel, K.-H.;  Allescher, W.; Hohmann, R.  (1990) Measurement of diesel exhaust at the
39             workplace. Zentralbl. Arbeitsmed. Arbeitsschutz Ergon. 40:  2-11.
         April 1995                                    7-132       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Letz, R.; Ryan, P. B.; Spengler, J. D. (1984) Estimated distributions of personal exposure to respirable particles.
  2            Environ. Monit. Assess. 4: 351-359.

  3     Lewis, C. W. (1991) Sources of air pollutants indoors: VOC and fine particulate species. J. Exposure Anal.
  4            Environ. Epidemiol. 1: 31-44.

  5     Lewtas, J.;  Mumford, J.; Everson, R. B.; Hulka, B.; Wilcosky, T.; Kozumbo, W.; Thompson, C.; George, M.;
  6            Dobias, L.; Sram, R.; Li, X.; Gallagher, J. (1993) Comparison of DMA adducts from exposure to
  7            complex mixtures in various human tissues and experimental systems. Environ. Health Perspect. 99:
  8            89-97.

  9     Ligocki, M. P.; Liu, H. I. H.; Cass,  G. R.; John, W. (1990) Measurements of particle deposition rates inside
 10            Southern California museums. Aerosol Sci. Technol. 13: 85-101.

 11     Lioy, P. J.  (1990) Assessing total human exposure to contaminants. Environ. Sci. Technol. 24: 938-945.

 12     Lioy, P. J.; Waldman, J. M.;  Buckley, T.;  Butler, J.; Pietarinen, C. (1990) The personal, indoor and outdoor
 13            concentrations of PM-10 measured in an industrial community during the winter. Atmos. Environ. Part B
 14            24:  57-66.

 15     Litzistorf, G.; Guillemin, M. P.; Buffat, Ph.; Iselin,  F. (1985) Influence of human activity on the airborne fiber
 16            level in paraoccupational environments. J. Air Pollut. Control Assoc. 35: 836-837.

 17     Lofroth, G.; Burton, R. M.; Forehand, L.;  Hammond, S,  K.; Seila, R. L.; Zweidinger, R. B.; Lewtas, J.
 18            (1989) Characterization of environmental tobacco smoke.  Environ. Sci.  Technol. 23: 610-614.

 19     Lofroth, G.; Stensman, C.; Brandhorst-Satzkorn, M. (1991) Indoor sources of mutagenic aerosol particulate
20            matter: smoking, cooking and incense burning. Mutat. Res. 261: 21-28.

21     Lowrey, A. H.; Kantor, S.;  Repace,  J. L. (1993) Outdoor and indoor respirable suspended particulates in
22            Budapest, Hungary. In: Jantunen, M.; Kalliokoski, P.; Kukkonen, E.; Saarela, K.; Seppanen, O.;
23            Vuorelma, H., eds. Indoor air '93: proceedings of the 6th international conference on indoor air quality
24            and climate: v. 3, combustion products, risk  assessment, policies; July;  Helsinki, Finland. Helsinki,
25            Finland: Indoor Air '93; pp. 93-96.

26     Mage, D. T. (1980) The statistical form for the ambient particulate standard annual arithmetic mean vs annual
27            geometric mean. J. Air Pollut. Control Assoc. 30: 706-708.

28     Mage, D. T. (1983) Public health aspects of air quality surveillance. Public Health Rev. 11: 5-54.

29     Mage, D. T. (1985) Concepts of human  exposure assessment for airborne particulate matter. Environ. Int.
30            11:407-412.

31      Mage, D. T. (1991) A comparison of the direct and indirect methods of human  exposure. In: Gledhill; Mauro,
32            F., eds. New horizons in biological  dosimetry. New York, NY: Wiley-Liss, Inc.; pp. 443-454. (Progress
33            in clinical and biological research: v. 372).

34     Mage, D. T.;  Buckley, T. J. (1995) The relationship between personal exposures and ambient concentrations of
35             particulate matter. Presented at: the  88th annual meeting of the Air & Waste Management Association;
36            June; San Antonio, TX. Pittsburgh,  PA:  Air & Waste Management Association; paper no. 95-MP18.01.

37      Mamane, Y. (1992) Characterization  of PTEAM indoor aerosol samples (electron microscopy analysis). Research
38            Triangle Park, NC: U.S. Environmental  Protection Agency, Atmospheric Research and Exposure
39            Assessment Laboratory.


         April  1995                                     7-133       DRAFT-DO NOT QUOTE OR CITE

-------
 1     Marple, V. A.; Rubow, K. L.; Turner, W.; Spengler, J. D. (1987) Low flow rate sharp cut impactors for indoor
 2            air sampling: design and calibration. JAPCA 37:  1303-1307.

 3     Martinelli, C. A.; Harley, N.  M.; Lippmann, M.; Cohen, B. S. (1983) Monitoring real-time aerosol distribution
 4            in the breathing zone.  Am. Ind. Hyg.  Assoc. J. 44: 280-285.

 5     Matthews, T. G.; Thompson,  C. V.; Wilson,  D. L.; Hawthorne, A. R.; Mage, D. T. (1989) Air velocities
 6            inside domestic environments: an important parameter in the study of indoor air quality and climate.
 7            Environ. Int. 15: 545-550.

 8     Mayo, E. (1960) Hawthorne and the Western  Electric Company. In: Merrill, H. F., ed. Classics in management.
 9            New York, NY: American Management Association; p. 237.

10     McKenzie, R.  L.; Bright, D. S.; Fletcher, R. A.; Hodgeson, J. A. (1982) Development of a personal exposure
11            monitor for two sizes of inhalable particulates. Environ. Int. 8: 229-233.

12     Miesner, E. A.; Rudnick, S. N.; Hu, F-C.; Spengler, J.  D.; Ozkaynak, H.; Preller, L.; Nelson, W. (1989)
13            Paniculate and nicotine sampling in public facilities and offices. JAPCA 39: 1577-1582.

14     Millette,  J. R.; Hays,  S. M. (1994) Resuspension of settled dust. In: Settled asbestos dust sampling and analysis.
15            Boca Raton, FL: CRC Press, Inc.

16     Morandi, M. T.; Stock, T. H.; Contant, C. F. (1986) Characterization of indoor microenvironmental exposures
17            to respirable paniculate matter. Presented at: 79th annual meeting of the Air Pollution Control
18            Association; June; Minneapolis, MN.  Pittsburgh, PA: Air Pollution Control Association; paper
19            no. 86-67.2.

20     Morandi, M. T.; Stock, T. H.; Contant, C. F. (1988) A comparative  study of respirable paniculate
21            microenvironmental concentrations and personal exposures. Environ. Monit. Assess. 10: 105-122.

22     Mumford, J. L.; Williams, R. W.; Walsh, D. B.; Burton, R. M.; Svendsgaard, D. J.; Chuang, J. C.;  Houk, V.
23            S.; Lewtas, J. (1991) Indoor air pollutants from unvented kerosene  heater emissions in mobile homes:
24            studies on particles, semivolatile organics, carbon monoxide, and mutagenicity. Environ. Sci. Technol.
25            25: 1732-1738.

26     Nagda, N.; Fortmann, R.; Koontz, M.; Konheim, A. (1990) Investigation of cabin air quality aboard commercial
27            airliners. In: Indoor air '90: precedings of the 5th international conference on indoor air quality and
28            climate, volume 2, characteristics of indoor air; July-August; Toronto, ON, Canada. Ottawa, ON,
29            Canada: International Conference on Indoor Air Quality and Climate, Inc.; pp. 245-250.

30     National  Research Council. (1986) Environmental tobacco smoke:  measuring exposures and  assessing health
31            effects. Washington, DC: National Academy Press.

32     Nazaroff, W. W.; Cass, G. R. (1989) Mathematical modeling of indoor aerosol dynamics. Environ. Sci.
33            Technol. 23:  157-166.

34     Nazaroff, W. W.; Salmon, L. G.; Cass, G. R. (1990a) Concentration and fate of airborne particles in museums.
35            Environ. Sci. Technol. 24: 66-77.

36     Nazaroff, W. W.; Ligocki, M. P.; Ma, T.; Cass, G. R.  (1990b) Particle deposition in museums: comparison of
37            modeling and measurement results.  Aerosol Sci. Technol.  13: 332-348.
         April 1995                                    7-134       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Nazaroff, W. W.; Gadgil, A. J.; Weschler, C. J. (1993) Critique of the use of deposition velocity in modeling
  2            indoor air quality. In: Nagda, N. L., ed. Modeling of indoor air quality and exposure. Philadelphia, PA:
  3            American Society for Testing and Materials; ASTM STP no.  1205; pp. 81-104.

  4     Neas, L. M.; Dockery, D.  W.; Ware, J. H.; Spengler, J. D.; Ferris, B. G.,  Jr.; Speizer, F. E. (1994)
  5            Concentration of indoor particulate matter as a determinant of respiratory health  in children. Am. J.
  6            Epidemiol. 139:  1088-1099.

  7     O'Brien, D. M.; Fischbach, T. J.; Cooper, T.  C.; Todd, W. F.; Gressel, M. F.; Martinez, K. F. (1989)
  8            Acquisition and spreadsheet analysis of real time dust exposure data:  a case study. Appl. Ind. Hyg. 4:
  9            238-243.

 10     Offermann, F. J.; Sextro, R. G.; Fisk, W. J.; Grimsrud, D. T.; Nazaroff, W.  W.; Nero, A.  V.; Revzan, K. L.;
 11            Yater, J. (1985) Control of respirable particles in indoor air with portable air cleaners. Atmos. Environ.
 12            19: 1761-1771.

 13     Ogden, T. L.; Bartlett, I. W.; Puraell, C. J.; Wells, C.  J.; Armitage, F.; Wolfson, H. (1993) Dust for cotton
 14            manufacture:  changing from static to personal sampling. Ann. Occup. Hyg. 37: 271-285.

 15     Oldaker, G. B., Ill; Ogden, M. W.;  Maiolo, K. C.; Conner, J. M.; Conrad,  F. W., Jr.; DeLuca, P. O. (1990)
 16            Results from surveys of environmental tobacco smoke in restaurants in Winston-Salem, North Carolina.
 17            In:  Indoor air '90: precedings of the 5th international conference on indoor air quality and climate,
 18            volume 2, characteristics of indoor air; July-August; Toronto, ON, Canada. Ottawa, ON, Canada:
 19            International Conference on Indoor Air Quality and Climate, Inc.; pp. 281-285.

 20     Ott, W.  R.  (1982) Concepts of human exposure to air pollution. Environ.  Int. 7: 179-196.

 21     Ott, W.  R.  (1990) A physical explanation of the lognormality of pollutant concentrations. J. Air Waste Manage.
 22            Assoc. 40: 1378-1383.

 23     Ott, W.  R.; Langan, L.; Switzer,  P.  (1992) A time series model for cigarette smoking activity patterns: model
 24            validation for carbon monoxide and respirable particles in a chamber  and an automobile. J. Exposure
 25            Anal. Environ. Epidemiol. 2(suppl. 2): 175-200.

 26     Ott, W. R.; Switzer, P.; Robinson, J. (1995a) Particle concentrations inside a tavern before and after prohibition
 27            of smoking: evaluating the performance of an indoor air quality model.  Submitted.

 28     Ott, W. R.; Klepeis, N. E.; Switzer, P. (1995b) Modeling environmental tobacco smoke in the home using
 29            transfer functions. Presented at: the 88th annual meeting and exhibition of the Air & Waste Management
 30            Association; June. Pittsburgh, PA: Air & Waste Management Association;  paper no. A-1043.

 31     Owen, M. K.; Ensor, D. S.; Hovis, L. S.; Tucker, W. G.; Sparks, L. E. (1990) Particle size distributions for an
 32            office aerosol. Aerosol Sci. Technol. 13: 486-492.

 33     Owen, M. K.; Ensor, D. S.; Sparks, L. E. (1992) Airborne particle sizes  and sources  found in indoor air.
 34            Atmos. Environ. Part A 26: 2149-2162.

 35     Ozkaynak, H.; Spengler, J.  D.; Ludwig, J. F.; Butler, D.  A.; Clayton,  C. A.; Pellizzari, E.;  Wiener, R. W.
 36            (1990) Personal exposure to particulate matter: findings from the Particle Total Exposure Assessment
 37            Methodology (PTEAM) prepilot study. In:  Indoor air '90: precedings  of the 5th international conference
 38           on indoor air quality and climate, volume 2, characteristics of indoor  air; July-August; Toronto, ON,
 39           Canada. Ottawa,  ON, Canada: International Conference on Indoor Air Quality and Climate, Inc.;
40           pp.  571-576.
         April 1995                                    7_135       DRAFT-DO NOT QUOTE OR CITE

-------
  1      Ozkaynak, H.; Xue, J.; Weker, R.; Butler, D.; Spengler, J. (1993a) The particle TEAM (PTEAM) study:
 2             analysis of the data, volume III [draft final report].  Research Triangle Park, NC: U.S. Environmental
 3             Protection Agency; EPA contract no. 68-02-4544.

 4      Ozkaynak, H.; Spengler, J. D.; Xue, J.; Koutrakis, P.; Pellizzari, E. D.;  Wallace, L. (1993b) Sources and
 5             factors influencing personal and indoor exposures to particles, elements and nicotine: findings from the
 6             particle TEAM pilot study. In: Jantunen,  M.; Kalliokoski, P.; Kukkonen, E.; Saarela, K.; Seppanen, O.;
 7             Vuorelma, H., eds. Indoor air '93: proceedings of the 6th international conference on indoor air quality
 8             and climate, v. 3, combustion products, risk assessment, policies;  July; Helsinki, Finland. Helsinki,
 9             Finland: Indoor Air '93; pp. 457-462.

10      Parker, R. C.; Bull, R. K.;  Stevens,  D. C.;  and Marshall,  M. (1990) "Studies of aerosol distributions in a small
11             laboratory containing a heated phantom", Annals of Occupational Hygiene, 34 (1), 34-44.

12      Pellizzari, E. D.; Thomas, K. W.; Clayton,  C. A.; Whitmore, R. W.; Shores,  R. C.; Zelon, H. S.; Perritt, R.
13             L.  (1992) Particle total exposure assessment methodology (PTEAM): Riverside,  California pilot study,
14             volume I  [final report]. Research  Triangle Park, NC: U.S. Environmental Protection Agency,
15             Atmospheric Research and Exposure Assessment Laboratory; EPA report no. EPA/600/R-93/050.
16             Available from: NTIS, Springfield, VA; PB93-166957/XAB.

17      Pellizzari, E. D.; Thomas, K. W.; Clayton,  C. A.; Whitmore, R. C.; Shores, H.; Zelon, S.; Peritt, R. L.
18             (1993) Particle total exposure assessment  methodology (PTEAM):  Riverside, California pilot
19             study—volume I [project summary].  Research Triangle Park, NC: U.S. Environmental Protection
20             Agency, Atmospheric Research and Exposure Assessment Laboratory; EPA report no. EPA/600/SR-
21             93/050.

22      Pengelly, L. D.; Goldsmith, C. H.; Kerigan, A. T.; Furlong, W.; Toplack, S.  (1987) The Hamilton study:
23             estimating exposure to ambient suspended particles. JAPCA 37: 1421-1428.

24      Penkala, S. J.; de Oliveira, G. (1975) The simultaneous analysis of carbon monoxide and suspended paniculate
25             matter produced by cigarette smoking. Environ. Res. 9: 99-114.

26      Perritt, R.  L.; Clayton, C. A.; Pellizzari,  E. D.; Thomas, K. W.; Wallace, L.  A.; Spengler, J.  D.; Ozkaynak,
27             H. (1991) Particle Total Exposure Assessment Methodology (PTEAM)  Pilot Study: personal, indoor, and
28             outdoor paniculate concentration distributions for southern California fall 1990—preliminary results. In:
29             Measurement of toxic and  related air pollutants: proceedings of the 1991 U.S. EPA/A&WMA
30             international symposium, v. 2;  May; Durham, NC. Pittsburgh, PA: Air & Waste Management
31             Association; pp. 665-671.  (A&WMA publication VIP-21).

32      Pickles, J.  H. (1982) Air pollution estimation error and what it does to epidemiological analysis. Atmos.  Environ.
33             16: 2241-2245.

34      Quackenboss, J. J.; Krzyzanowski, M.; Lebowitz, M. D. (1991) Exposure assessment approaches to evaluate
35             respiratory health effects of particulate matter and nitrogen dioxide. J. Exposure Anal. Environ.
36             Epidemiol. 1:  83-107.

37      Raunemaa, T.; Kulmala, M.; Saari, H.; Olin, M.; Kulmala, M. H. (1989) Indoor air aerosol model: transport
38             indoors and deposition of fine and coarse particles.  Aerosol Sci. Technol. 11:  11-25.

39      Repace, J.  L. (1987a) Indoor concentrations  of environmental tobacco smoke: field surveys. In: O'Neill,  I. K.;
40             Brunnemann, K. D.; Dodet, B.; Hoffmann, D., eds. Environmental carcinogens—methods of analysis
41             and exposure measurement: v. 9, passive smoking.  Lyon, France: World Health Organization;
42             pp. 141-162. (IARC scientific publications no. 81).
         April 1995                                    7-136       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Repace, J. L. (1987b) Indoor concentrations of environmental tobacco smoke: models dealing with effects of
  2            ventilation and room size. In: O'Neill, I. K.; Brunnemann, K. D.; Dodet, B.; Hoffmann, D., eds.
  3            Environmental carcinogens—methods of analysis and exposure measurement: v. 9, passive smoking.
  4            Lyon, France:  World Health Organization; pp. 25-41. (IARC scientific publications no. 81).

  5     Repace, J. L.; Lowrey, A. H. (1980) Indoor air pollution, tobacco smoke, and public health. Science 208: 464-
  6            472.

  7     Repace, J. L.; Lowrey, A. H. (1982) Tobacco  smoke, ventilation, and indoor air quality. ASHRAE Trans.
  8            88: 895-914.

  9     Repace, J. L.; Ott, W. R.; Wallace, L. A. (1980) Total human exposure to air pollution. Presented at:  73rd
 10            annual meeting of the Air Pollution Control Association; June;  Montreal, PQ, Canada. Pittsburgh, PA:
 11            Air Pollution Control Association; paper no. 80-61.6.

 12     Revsbech, P.; Korsgaard, J.;  Lundqvist, G. R.  (1987) Suspended particulate matter in dwellings—the impact of
 13            tobacco smoking. Environ. Int.  13: 147-150.

 14     Rickert, W. S.; Robinson, J. C.; Collishaw, N. (1984) Yields of tar, nicotine, and carbon monoxide in the
 15            sidestream smoke from 15 brands of Canadian cigarettes. Am. J. Public Health 74: 228-231.

 16     Roberts, J. W.; Camann, D. E.;  Spittler, T. M. (1990) Monitoring and controlling lead in house dust in older
 17            homes. In: Indoor air '90: precedings of the 5th international conference  on indoor air quality and
 18            climate, volume 2, characteristics of indoor air; July-August; Toronto, ON, Canada. Ottawa, ON,
 19            Canada:  International Conference on Indoor Air Quality and Climate, Inc.; pp. 435-440.

 20     Robinson, J.; Nelson, W. C. (1995) National human activity pattern survey data base. Research Triangle Park,
 21            NC:  U.S. Environmental Protection Agency.

 22     Rodes, C. E.; Kamens, R.  M.; Wiener,  R. W.  (1991) The significance and characteristics of the personal
 23            activity cloud on exposure assessment measurements for indoor contaminants. Indoor Air 2: 123-145.

 24     Roemmelt, H.; Hoeppe, P.; Prami, G.; Schierl, R.; Zielinsky, M. (1993) Measurement of exposure of drivers
 25            and passengers to dust and exhaust  gases in public transportation in Munich. Berlin, Germany: GSF;
 26            pp. 83-90; technical report 31/93.

 27     Rudolf, W.  (1994) Concentration of air pollutants inside cars driving on highways and and in downtown areas.
 28            Sci. Total Environ. 146/147: 433-444.

 29     SAS Institute. (1990) SAS language and procedures: SYNTAX, version 6. 1st ed. Gary, NC: SAS Institute.

 30     Santanam, S.; Spengler, J.  D.; Ryan, P. B. (1990) Particulate matter exposures estimated from an indoor-outdoor
 31            source apportionment study. In: Indoor  air  '90: precedings of the 5th international conference on indoor
 32            air quality and climate, volume 2, characteristics of indoor air; July-August; Toronto, ON, Canada.
 33            Ottawa, ON, Canada: International Conference on Indoor Air Quality and Climate,  Inc.; pp. 583-588.

34     Sexton, K.;  Spengler, J. D.; Treitman, R. D. (1984) Personal exposure to respirable particles:  a case study in
35            Waterbury,  Vermont.  Atmos. Environ.  18: 1385-1398.

36     Shair,  F. H.; Heitner, K. L. (1974) Theoretical model for relating indoor pollutant concentrations to those
37            outside. Environ. Sci. Technol.  8: 444-451.
         April 1995                                    7-137       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Sheldon, L. S.; Handy, R. W.; Hartwell, T. D.; Whitmore, R. W.; Zelon, H. S.; Pellizzari, E.  D. (1988a)
  2            Indoor air quality in public buildings: v. I. Washington, DC: U.S. Environmental Protection Agency,
  3            Office of Acid Deposition, Environmental Monitoring, and Quality Assurance; report no. EPA
  4            600/6-88/009a. Available from: NTIS, Springfield, VA; PB89-102503.

  5     Sheldon, L. S.; Zelon, H. S.; Sickles, J.; Eaton, C.; Hartwell, J. (1988b) Indoor air quality in public buildings,
  6            v. II. Research Triangle Park, NC: U.S. Environmental Protection Agency, Environmental Monitoring
  7            Systems Laboratory; report no. EPA 600/6-88/009b. Available from: NTIS, Springfield,  VA;
  8            PB89-102511.

  9     Sheldon, L. S.; Hartwell, T. D.;  Cox, B. G.; Sickles, J. E., II; Pellizzari, E. D.; Smith, M. L.; Perritt, R. L.;
 10            Jones, S. M. (1989) An investigation of infiltration and indoor air quality: final report. Albany, NY:
 11            New York State Energy Research and Development Authority; New York State ERDA contract no.
 12            736-CON-BCS-85.

 13     Sheldon, L.; Clayton, A.; Keever, J.  Perritt, R. L.;  Whitaker, D. (1992) PTEAM: monitoring of phthalates and
 14            PAHs in indoor and outdoor air samples in Riverside, California, volume II [final report]. Sacramento,
 15            CA: California State Air Resources Board. Available from: NTIS, Springfield, VA; PB93-205649/XAB.

 16     Sinclair, J. D.; Psota-Kelty, L. A.; Weschler, C. J.  (1988) Indoor/outdoor ratios and indoor surface
 17            accumulations of ionic substances at Newark, New Jersey. Atmos. Environ. 22: 461-469.

 18     Sinclair, J. D.; Psota-Kelty, L. A.; Weschler, C. J.; Shields, H.  C. (1990) Measurement and modeling of
 19            airborne concentrations and indoor surface accumulation rates of ionic substances at Neenah, Wisconsin.
20            Atmos. Environ. Part A 24: 627-638.

21     Sinclair, J. D.; Psota-Kelly, L. A.; Peins, G. A.; Ibidunni, A. O. (1992) Indoor/outdoor relationships of
22            airborne ionic substances: comparison of electronic equipment room and factory environments.  Atmos.
23            Environ. Part A 26: 871-882.

24     Smith, K. R.;  Apte, M. G.; Yuqing, M.; Wongsekiarttirat, W.; Kulkarni, A. (1994) Air pollution and the energy
25            ladder in asian cities. Energy (Oxford) 19: 587-600.

26     Smithard, E. H. R.  (1954) The 1952 fog in a metropolitan borough. Mon. Bull. Minist. Health Public
27            Health Serv. (G. B.) 13: 26-35.

28     Sparks, L. E.; Tichenor, B. A.; White, J. B.; Chang, J.; Jackson, M. D. (1991) Verification and uses of the
29            Environmental Protection Agency (EPA) indoor air quality model. Presented at: 84th annual meeting of
30            the Air and  Waste Management Association; June; Vancouver, BC, Canada. Pittsburgh, PA: Air and
31            Waste Management Association; paper no. 91-62.12.

32     Spengler, J. D.; Soczek, M. L. (1984) Evidence for  improved ambient air quality and the need for personal
33            exposure research.  Environ. Sci.  Technol. 18: 268A-280A.

34     Spengler, J. D.; Thurston, G. D.  (1983)  Mass and elemental composition of fine and coarse particles in six U.S.
35            cities. J. Air Pollut. Control Assoc. 33: 1162-1171.

36     Spengler, J. D.; Dockery, D. W.; Turner, W. A.; Wolfson, J. M.; Ferris, B. G., Jr. (1981) Long-term
37            measurements of respirable sulfates  and particles inside and outside homes. Atmos. Environ. 15: 23-30.

38     Spengler, J. D.; Treitman, R. D.; Tosteson, T. D.; Mage, D. T.; Soczek, M. L. (1985) Personal exposures to
39            respirable particulates and implications for air pollution epidemiology. Environ. Sci. Technol.
40            19: 700-707.
         April 1995                                    7-138       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Spengler, J. D.; Ware, J.; Speizer, F.; Ferris, B.; Dockery, D.; Lebret, E.; Brunekreef, B. (1987) Harvard's
  2            indoor air quality respiratory health study. In: Seifert, B.; Esdorn, H.; Fischer, M.; Rueden, H.;
  3            Wegner, J., eds. Indoor air '87: proceedings of the 4th international conference on indoor air quality and
  4            climate, v. 2, environmental tobacco smoke, multicomponent studies, radon, sick buildings,  odours and
  5            irritants, hyperreactivities and allergies; August; Berlin, Federal Republic of Germany. Berlin, Federal
  6            Republic of Germany: Institute for Water, Soil and Air Hygiene; pp. 218-223.

  7     Sterling, T. D.; Dimich, H.; Kobayashi, D. (1982) Indoor byproduct levels of tobacco smoke: a critical review
  8            of the literature. J. Air Pollut. Control Assoc. 32: 250-259.

  9     Stevens, D. C. (1969) "The particle size and mean concentration of radioactive aerosols measured by personal
 10            and static  air samples", Annals of Occupational Hygiene, 12, 33-40.

 11     Suh, H. H.; Koutrakis, P.; Spengler, J. D. (1993) Validation of personal  exposure models for sulfate and aerosol
 12            strong acidity. J. Air Waste Manage. Assoc. 43: 845-850.

 13     Switzer, P.; Ott, W. R. (1992) Derivation of an indoor air averaging time model from the mass balance equation
 14            for the case of independent source inputs and fixed air exchange rates. J. Exposure Anal. Environ.
 15            Epidemiol. 2(suppl. 2): 113-135.

 16     Szalai, A., ed.  (1972)  The use of time: daily activities of urban and suburban populations in 12  countries. The
 17            Hague, The Netherlands: Mouton and Co.

 18     Teschke, K.; Hertzman, C.; Van Netten, C.; Lee, E.; Morrison, B.;  Cornista, A.; Lau, G.; Hundal, A.(1989)
 19            Potential exposure of cooks to airborne mutagens and carcinogens. Environ. Res.  50: 296-308.

 20     Thatcher, T. L.; Layton, D. W. (1994) Deposition, resuspension, and penetration of particles within a residence.
 21            Riverside, CA: University of California, Lawrence Livermore National Laboratory; report no.
 22            UCRL-JC-116597.

 23     Thomas, K. W.; Pellizzari, E. D.; Clayton, C. A.;  Whitaker, D. A.; Shores, R. C.; Spengler, J. D.; Ozkaynak,
 24            H.; Wallace, L. A. (1993) Particle total exposure assessment  methodology (PTEAM) study:  method
 25            performance and data quality for personal, indoor, and outdoor aerosol monitoring at 178 homes in
 26            southern California. J. Exposure Anal.  Environ. Epidemiol. 3: 203-226.

 27     Traynor, G. W.; Aceti, J. C.; Apte, M. G.; Smith, B. V.; Green,  L. L.;  Smith-Reiser, A.;  Novak,  K. M.;
 28            Moses, D. O. (1989) Macromodel for assessing residential  concentrations of combustion-generated
 29            pollutants: model development and preliminary predictions for CO, NO2, and respirable  suspended
 30            particles.  Berkeley, CA: U.S. Department of Energy, Lawrence Berkeley Laboratory; report no.
 31            LBL-25211. Available from: NTIS, Springfield, VA; DE89013396/XAB.

 32     Turk, A. (1963) Measurements of odorous vapors in test chambers: theoretical. ASHRAE J. 5: 55-58.

 33     Turk, B. H.; Brown, J. T.; Geisling-Sobotka, K.; Froehlich, D. A.; Grimsrud, D. T.; Harrison, J.;
 34            Koonce, J. F.; Prill, R. J.; Revzan, K. L. (1987) Indoor air quality and  ventilation measurements  in
 35            38 Pacific  Northwest commercial buildings. Volume I: Measurement results  and interpretation [final
 36            report]. Berkeley, CA: Lawrence Berkeley Laboratory; report no.  LBL-22315.

 37     Turk, B. H.; Grimsrud, D. T.; Brown, J. T.;  Geisling-Sobotka, K. L.; Harrison, J.; Prill, R. J. (1989)
 38            Commercial building ventilation rates and particle concentrations. ASHRAE  Trans. 95: 422-433.

39     Turner, S.; Cyr, L.; Gross, A. J. (1992) The measurement of environmental tobacco smoke  in 585 office
40            environments. Environ. Int. 18: 19-28.
         April 1995                                    7^39       DRAFT-DO NOT QUOTE OR CITE

-------
  1      U.S. Centers for Disease Control. (1988) Assessing exposures of health care personnel to aerosols of
  2             ribavirin—California. Morb. Mortal. Wkly. Rep. 37: 560-563.

  3      U.S. Environmental Protection Agency. (1992) Respiratory health effects of passive smoking: lung cancer and
  4             other disorders [review draft]. Washington, DC: Office of Research and Development, Office of Health
  5             and Environmental Assessment; EPA report no. EPA/600/6-90/006B. Available from: NTIS, Springfield,
  6             VA; PB92-182344.

  7      U.S. House of Representatives. (1994) Environmental tobacco smoke investigation: staff report. Washington,
  8             DC: Subcommittee on Health and the Environment.

  9      Vaughan, W. M.; Hammond, S. K.  (1990) Impact of "designated smoking area" policy on nicotine vapor and
10             particle concentrations in a modern office building. J. Air Waste Manage. Assoc. 40:  1012-1017.

11      Wallace, L.; Thomas, J.; Mage, D.; Ott, W. (1988) Comparison of breath CO, CO exposure, and Coburn model
12             predictions in the U.S. EPA Washington-Denver (CO) study. Atmos. Environ. 22: 2183-2193.

13      Wallace, L.; Clayton, A.; Whitmore, R.; Shores,  R.; Thomas, K.; Whitaker, D.; Reading, P.; Pellizzari, E.;
14             Spengler, J.; Ozkaynak, H.; Froehlich, S.; Jenkins,  P.; Ota, L.; Westerdahl, D. (1991a) Initial results
15             from the PTEAM study: survey design, population response rates, monitor performance and quality
16             control. In: Measurement of toxic and related air pollutants: proceedings of the 1991 U.S.
17             EPA/A&WMA international symposium, v. 2; May; Durham, NC.  Pittsburgh, PA: Air & Waste
18             Management Association; pp. 659-664. (A&WMA publication VIP-21).

19      Wallace, L. A.; Pellizzari, E.; Sheldon, L.; Whitmore, R.;  Zelon, H.; Clayton, A.; Shores, R.; Thomas, K.;
20             Whitaker, D.; Reading, P.; Spengler, J.; Ozkaynak, H.; Froehlich,  S.; Jenkins, P.; Ota, L.; Westerdahl,
21             D. (1991b) The TEAM study of inhalable  particles (PM10): study design, sampler performance, and
22             preliminary results. Presented at: 84th annual meeting of the Air &  Waste Management Association;
23             June; Vancouver, British Columbia, Canada. Pittsburgh, PA: Air & Waste Management Association;
24             paper no. 91-171.3.

25      Wallace, L. A.; Ozkaynak, H.; Spengler, J. D.; Pellizzari, E. D.; Jenkins,  P. (1993)  Indoor, outdoor, and
26             personal air exposures to particles, elements, and nicotine for 178 southern California residents. In:
27             Jantunen, M.; Kalliokoski, P.;  Kukkonen,  E.; Saarela, K.; Seppanen, O.; Vuorelma, H., eds. Indoor air
28             '93: proceedings of the 6th international conference  on indoor air quality and climate, v.  3, combustion
29             products, risk assessment,  policies; July; Helsinki, Finland. Helsinki, Finland: Indoor Air '93;
30             pp.  445-450.

31      Weschler, C. J.; Shields, H. C.; Kelty, S. P.; Psota-Kelly, L. A.; Sinclair,  J. D. (1989) Comparison of effects
32             of ventilation, filtration and outdoor airat telephone office buildings. In: Nagda, N. L.; Harper, J. P.,
33             eds. Design and protocol for monitoring indoor air quality. Philadelphia, PA: American Society for
34             Testing and Materials; pp. 9-34. (ASTM special technical publication 1002).

35      Wiener, R.  W. (1988) Measurement and evaluation of personal exposure to  aerosols. In: Measurement of toxic
36             and related air pollutants: proceedings of the 1988 EPA/APCA international symposium; May; Research
37             Triangle Park, NC. Pittsburgh, PA: Air Pollution Control Association;  pp. 84-88. (APCA publication no.
38             VIP-10).

39      Wiener, R.  W. (1989) Particle total exposure assessment methodology—an overview of planning and
40             accomplishments. In: Measurement of toxic and related air pollutants: proceedings of the 1989 U.S.
41             EPA/A&WMA international symposium. Pittsburgh, PA: Air & Waste  Management Association;
42             pp.  442-448. (A&WMA publication VIP-13).
         April 1995                                    7-140       DRAFT-DO NOT QUOTE OR CITE

-------
  1     Wiener, R. W.; Wallace, L.; Pahl, D.; Pellizzari, E.; Whittaker, D.; Spengler, J.; Ozkaynak, H. (1990) Review
  2            of the particle TEAM 9 home field study. In: Measurement of toxic and related air pollutants:
  3            proceedings of the 1990 EPA/A&WMA international symposium, v. 2;  May; Raleigh, NC. Pittsburgh,
  4            PA: Air and Waste Management Association; pp. 452-460. (A&WMA publication VIP-17).

  5     World Health Organization.  (1982a) Human exposure to carbon monoxide and suspended paniculate matter in
  6            Zagreb, Yugoslavia. Geneva, Switzerland: World Health Organization; GEMS (Global Environmental
  7            Monitoring System)  report no. EFP/82.33.

  8     World Health Organization.  (1982b) Human exposure to SO2, NO2 and suspended paniculate matter in Toronto,
  9            Canada. Geneva, Switzerland: World Health Organization; GEMS (Global Environmental Monitoring
10            System) report no. EFP/82.38.

11     World Health Organization.  (1984) Human exposure to suspended paniculate matter and sulfate in Bombay,
12            India. Geneva, Switzerland: World Health Organization; GEMS (Global Environment Monitoring
13            System) report no. EFP/84.86.

14     World Health Organization.  (1985) Human exposure to carbon monoxide and suspended paniculate matter in
15            Beijing, People's Republic of China. Geneva, Switzerland: World Health Organization; GEMS (Global
16            Environmental Monitoring System) report no. EFP/84.66.

17     World Health Organization.  (1992) Urban air pollution in the megacities of the world. Oxford, United Kingdom:
18            Blackwell Publishers.

19     Xu, X.; Gao, J.; Dockery, D. W.; Chen, Y. (1994) Air pollution and daily mortality in residential areas of
20            Beijing, China. Arch. Environ. Health 49: 216-222.

21     Yost, M. G.; Gadgil, A. J.;  Drescher, A. C.; Zhou, Y.; Simonds, M. A.;  Levine, S. P.; Nazaroff, W. W.;
22            Saisan, P. A. (1994) Imaging indoor tracer-gas concentrations with computed tomography: experimental
23            results with a remote sensing FTIR system. Am. Ind. Hyg. Assoc. J. 55: 395-402.
        April  1995  *u.s.  G.p.o.:i995-652-486       7-141       DRAFT-DO NOT QUOTE OR CITE

-------