EPA
          United States
          Environmental Protection
          Agency
             Environmental Criteria and
             Assessment Office
             Washington DC 20460
EPA/600/8-84/020dP
August 1986
           Research and Development
Air Quality Criteria for
Ozone and Other
Photochemical
Oxidants
           Volume IV of V

-------

-------
                       EPA-600/8-84-020dF
                               August 1986
    Air Quality Criteria
  for Ozone and Other
Photochemical  Oxidants

      Volume IV  of V
    Environmental Criteria and Assessment Office
   Office of Health and Environmental Assessment
      Office of Research and Development
      U.S. Environmental Protection Agency
      Research Triangle Park, N.C. 27711

-------
                                DISCLAIMER
     This document has been reviewed in accordance with U.S.  Environmental
Protection Agency policy and approved for publication.  Mention of trade
names or commercial products does not constitute.endorsement or
recommendation.

-------
                                   ABSTRACT


     Scientific information is presented and evaluated relative to the health
and welfare effects associated with exposure to ozone and other photochemical
oxidants.   Although it is not intended as a complete and detailed literature
review, the document covers pertinent literature through early 1985.

     Data on health and welfare effects are emphasized, but additional infor-
mation is provided for understanding the nature of the oxidant pollution pro-
blem and for evaluating the reliability of effects data as well as their
relevance to potential exposures to ozone and other oxidants at concentrations
occurring in ambient air.  Information is presented on the following exposure-
related topics:  nature, source, measurement, and concentrations of precursors
to ozone and other photochemical oxidants; the formation of ozone and other
photochemical oxidants and their transport once formed; the properties, chem-
istry, and measurement of ozone and other photochemical oxidants; and the
concentrations of ozone and other photochemical oxidants that are typically
found in ambient air.

     The specific areas addressed by chapters on health and welfare effects
are the toxicological appraisal of effects of ozone and other oxidants; effects
observed in controlled human exposures; effects observed in field and epidemic-
logical studies; effects on vegetation seen in field and controlled exposures;
effects on natural and agroecosystems; and effects on nonbiological materials
observed in field and chamber studies.
                                       111

-------
                        AIR QUALITY CRITERIA FOR OZONE
                       AND OTHER PHOTOCHEMICAL OXIDANTS
                                                                           Page
VOLUME I
   Chapter 1.    Summary and Conclusions		     1-1

VOLUME II
   Chapter 2.    Introduction		..	.     2-1
   Chapter 3.    Properties, Chemistry, and Transport of Ozone and
                 Other Photochemical Qxidants and Their Precursors 	     3-1
   Chapter 4.    Sampling and Measurement of Ozone and Other
                 Photochemical Oxidants and Their Precursors .»	     4-1
   Chapter 5. '   Concentrations of Ozone and Other Photochemical
                ^Oxidants in Ambient Air	,.	     5-1

VOLUME III
   Chapter 6.    Effects of Ozone and Other Photochemical Oxidants
                 on Vegetation	     6-1
   Chapter 7.    Effects of Ozone on Natural Ecosystems and Their
                 Components		     7-1
   Chapter 8.    Effects of Ozone and Other Photochemical Oxidants
                 on Nonbiological Materials	     8-1

VOLUME IV
   Chapter 9.    Toxicological Effects of Ozone and Other
                 Photochemical Oxidants	     9-1

VOLUME V
   Chapter 10.   Controlled Human Studies of the Effects of Ozone
                 and Other Photochemical Oxidants	     10-1
   Chapter 11.   Field and Epidemiological Studies of the Effects
                 of Ozone and Other Photochemical Oxidants 	     11-1
   Chapter 12.   Evaluation of Health Effects Data for Ozone and
                 Other Photochemical Oxidants 	     12-1
                                      IV

-------
                               TABLE OF CONTENTS


                                                                           Page

LIST OF TABLES	    viii
LIST OF FIGURES 	    ix
LIST OF ABBREVIATIONS				    xiii
AUTHORS, CONTRIBUTORS, AND REVIEWERS	    xiv

9.  TOXICOLOGICAL EFFECTS OF OZONE AND OTHER PHOTOCHEMICAL OKIDANTS ...    9-1

     9.1 INTRODUCTION	    9-1

     9.2   REGIONAL DOSIMETRY IN THE RESPIRATORY TRACT	    9-3
           9.2.1   Absorption in Experimental Animals	    9-4
                   9.2.1.1  Nasopharyngeal Absorption .....	.,    9-4
                   9.2.1.2  Lower Respiratory Tract Absorption 	    9-5
           9.2.2   Ozone Dosimetry Models 	J    9-6
                   9.2.2.1  Modeling Nasal Uptake ..	............/;.;    9-6
                   9.2.2.2  Lower Respiratory Tract Dosimetry
                             Models	    9-6
           9.2.3   Predictions of Lower Respiratory Tract Ozone
                   Dosimetry Modeling	    9-10
                   9.2.3.1  Illustration of Dosimetry Simulations  	    9-11
                   9.2.3.2  Comparison of Simulations to Experimental
                             Data		..............	    9-14
                   9.2.3.3  Uses of Predicted Dose .	    9-15

     9.3   EFFECTS OF OZONE ON THE RESPIRATORY TRACT	    9-16
           9.3.1   Morphological Effects	    9-16
                   9.3.1.1   Sites Affected	    9-16
                   9.3.1.2   Sequence in which Sites are Affected
                             as a Function of Concentration and
                             Duration of Exposure		    9-41
                   9.3.1.3   Structural Elements Affected  ...	......    9-42
                   9.3.1.4   Considerations of Degree of Suscepti-
                             bility to Morphological Changes	    9-46
           9.3.2   Pulmonary Function Effects ........		    9-52
                   9.3.2.1   Short-Term Exposure 	    9-52
                   9.3.2.2   Long-Term Exposure		    9-57
                   9.3.2.3   Airway Reactivity	    9-62
           9.3.3   Biochemically Detected Effects	    9-73
                   9.3.3.1   Introduction 	    9-73
                   9.3.3.2   Antioxidant Metabol ism	    9-73
                   9.3.3.3   Oxidative and Energy Metabolism ...	...    9-88
                   9.3.3.4   Monooxygenases 	,	    9-92
                   9.3.3.5   Lactate Dehydrogenase and  Lysosomal
                             Enzymes	    9-95
                   9.3.3.6   Protein Synthesis	    9-98
                   9.3.3.7   Lipid Metabolism and Content  of the
                             Lung	    9-108

-------
                      TABLE OF CONTENTS (continued)
9.5
9.6
      9.3.4
      9.3.5
9.3.3.8   Lung Permeability 	
9.3.3.9   Proposed Molecular Mechanisms of
          Effects 	
Effects on Host Defense Mechanisms	
          Mucociliary Clearance 	
          Alveolar Macrophages 	
          Interaction with Infectious Agents
          Immunology•	
9.3.4.1
9.3.4.2
9.3.4.3
9.3.4.4
Tolerance
9.4   EXTRAPULMONARY EFFECTS OF OZONE
      9.
      9.
      9.
      9.4.4
      9.4.5
      9.4.6
Central Nervous System and Behavioral Effects
Cardiovascular Effects		
Hematological and Serum Chemistry Effects ....
9.4.3.1   Animal Studies - Iri Vivo Exposures .
9.4.3.2   In Vitro Studies 	
9.4.3.3   Changes in Serum 	
9.4.3.4   Interspecies Variations 	
Reproductive and Teratogenic Effects 	
Chromosomal and Mutational Effects 	
9.4.5.1   Chromosomal Effects of Ozone 	
9.4.5.2   Mutational Effects of Ozone 	
Other Extrapulmonary Effects	
9.4.6.1   Liver 	
9.4.6.2   The Endocrine System		i.,
9.4.6.3   Other Effects	
EFFECTS OF OTHER PHOTOCHEMICAL OXIDANTS ......
9.5.1 Peroxyacetyl Nitrate 	 	 	
9.5.2 Hydrogen Peroxide 	 	
9.5.3 Formic Acid 	
9.5.4 Complex Pollutant Mixtures 	
SUMMARY 	 	 	 	 	
9.6.1 Introduction 	 	 	
9.6.2 Regional Dosimetry in the Respiratory
9.6.3 Effects of Ozone on the Respiratory Tr
9.6.3.1 Morphological Effects ......
9.6.3.2 Pulmonary Function 	
9.6.3.3 Biochemical Effects 	
9.6.3.4 Host Defense Mechanisms ....
9. 6. 3. 5 To! erance 	 	 	 	
9.6.4 Extraoulmonarv Effects of Ozone 	







Tract 	
"act 	 	






              9.6.4.1   Central Nervous System and Behavioral
                        Effects	
              9.6.4.2   Cardiovascular Effects 	
              9.6.4.3   Hematological and Serum Chemistry
                        Effects 	
Page

9-110

9-113
9-119
9-120
9-125
9-132
9-141
9-144
                                                        9-153
                                                        9-153
                                                        9-158
                                                        9-159
                                                        9-159
                                                        9-166
                                                        9-170
                                                        9-172
                                                        9-173
                                                        9-176
                                                        9-176
                                                        9-189
                                                        9-190
                                                        9-190
                                                        9-197
                                                        9-204

                                                        9-205
                                                        9-205
                                                        9-206
                                                        9-209
                                                        9-209

                                                        9-217
                                                        9-217
                                                        9-218
                                                        9-221
                                                        9-221
                                                        9-224
                                                        9-228
                                                        9-235
                                                        9-243
                                                        9-244

                                                        9-245
                                                        9-245

                                                        9-246
                                 VI

-------
                      TABLE OF CONTENTS (continued)
                                                                      Page
              9.6.4.4   Cytogenetic and Teratogenic Effects	     9-247
              9.6.4.5   Other Extrapulmonary Effects	     9-248
      9.6.5   Interaction of Ozone with Other Pollutants	     9-249
      9.6.6   Effects of Other Photochemical Oxidants		     9-252
9.7   REFERENCES			     9-257

      APPENDIX A	      A-l
                                 vn

-------
                                LIST OF TABLES
Table
Page
9-1   Morphological effects of ozone 	    9-17
9-2   Effects of ozone on pulmonary function:  short-term exposures ...    9-53
9-3   Effects of ozone on pulmonary function:  long-term exposures ....    9-58
9-4   Effects of ozone on pulmonary function:  airway reactivity ......    9-63
9-5   Changes in the lung antioxidant metabolism and oxygen
      consumpti on by ozone	    9-75
9-6   Monooxygenases	    9-93
9-7   Lactate dehydrogenase and lysosomal enzymes	,    9-96
9-8   Effects of ozone on lung protein synthesis	    9-99
9-9   Effects of ozone exposure on lipid metabolism and content of
      the lung	.	    9-109
9-10  Effects of ozone on lung permeability	    9-111
9-11  Effects of ozone on host defense mechanisms:  deposition and
      clearance		    9-122
9-12  Effects of ozone on host defense mechanisms:  macrophage
      alterations	.'	    9-126
9-13  Effects of ozone on host defense mechanisms:  interactions
      with infectious agents .,	    9-135
9-14  Effects of ozone on host defense mechanisms:  mixtures 	    9-138
9-15  Effects of ozone on host defense mechanisms: immunology 	    9-142
9-16  Tolerance to ozone					    9-147
9-17  Central nervous system and behavioral effects of ozone 	    9-154
9-18  Hematology;  animal—in vivo exposure 	    9-160
9-19  Hematology:  animal—Tn vitro exposure 	    9-167
9-20  Hematology:  human—in vitro exposure 	    9-168
9-21  Reproductive and teratogenic effects of ozone 	    9-175
9-22  Chromosomal effects from iji vitro exposure to high ozone
      concentrations 	    9-177
9-23  Chromosomal effects from ozone concentrations at or below
      1960 ug/m3 (1 ppm)	    9-179
9-24  Hutational effects of ozone	    9-184
9-25  Effects of ozone on the 1 i ver	    9-191
9-26  Effects of ozone on the endocrine system, gastrointestinal
      tract, and urine	    9-198
9-27  Effects of complex pollutant mixtures 	    9-211
9-28  Summary Table: morphological effects of ozone in experimental
      animal s 	    9-226
9-29  Summary Table: effects on pulmonary function of short-term
      exposures to ozone in experimental animals 	    9-230
9-30  Summary Table: effects on pulmonary function of long-term
      exposures to ozone in experimental animals	    9-232
9-31  Summary Table: biochemical changes in experimental animals
      exposed to ozone	    9-237
9-32  Summary Table: effects of ozone on host defense mechanisms in
      experimental animals 		........    9-242
9-33  Summary Table: extrapulmonary effects of ozone in experimental
      animals 	'	    9-251
9-34  Summary Table: interaction of ozone with other pollutants in
      experimental animal s		    9-254

                                      viii

-------
                                LIST OF FIGURES


Figure                                                                     Page

9-1   Predicted tissue dose for several trachea!  03 concentrations
      for rabbit and guinea pig	      9-12
9-2   Tissue dose versus ozone for rabbit and guinea pig and tissue
      dose versus airway generation for human.  Trachea! 03 concen-
      tration is 500 ug/ms (0.26 ppm) 		      9-13
9-3   Intracellular compounds active in antioxidant metabolism of
      the 1 ung	.-	      9-73
9-4   Summary of morphological effects in experimental animals
      exposed to ozone	      9-225
9-5   Summary of effects of short-term ozone exposures on pulmonary
      function in experimental animals 	    _  9-229
9-6   Summary of effects of long-term ozone exposures on pulmonary
      function in experimental animals	      9-231
9-7   Summary of biochemical changes in experimental animals
      exposed to ozone	      9-236
9-8   Summary of effects of ozone on host defense mechanisms in
      experimental animals 	 	      9-241
9-9   Summary of extrapulmonary effects of ozone in experimental
      animal s 	      9-250
9-10  Summary of effects in experimental animals exposed to ozone
      combined with other pollutants	      9-253

-------
                             LIST OF ABBREVIATIONS

A-V                              Atrioventrieular
ACh                              Acetylcholine
AChE                             Acetylcholinesterase
AM                               Alveo.lar macrophage
AMP                              Adenosine monophosphate
ATP                              Adenosine triphosphate
ATPS                             ATPS condition (ambient temperature and pressure,
                                 saturated with water vapor)
BTPS                             BTPS conditions (body temperature, barometric
                                 pressure, and saturated with water vapor)
CC                               Closing capacity
C.                               Dynamic lung compliance
CHEM                             Gas-phase chemiluminescence
C.                                Lung compliance
C.  i                             Static lung compliance
CMP                              Cytidine monophosphate
CNS                              Central nervous system
CO                               Carbon monoxide
COHb                             Carboxyhemoglobin
COLD                             Chronic obstructive lung disease
COMT                             Catechol-£-methy1-transferase
COg                              Carbon dioxide
CPK                              Creatine phosphokinase
CV                               Closing volume
D.                                Diffusing capacity of the lungs
D.CO                     '        Carbon monoxide diffusing capacity of the lungs
DNA                              Deoxyribonucleic acid
E                                Elastance
EGG, EKG                         Electrocardiogram
EEG                              Electroencephalogram
ERV                              Expiratory reserve volume
   m=,v                           The maximal forced expiratory flow achieved
   max
                                 during an FVC test

-------
                      LIST OF ABBREVIATIONS (continued)

FEF                              Forced expiratory flow
FEF2QO-12QQ                      Mean forced expiratory flow between 200 ml and
                                 1200 ml of the FVC [formerly called the maximum
                                 expiratory flow rate (MEFR)].
FEF25-75%                        Mean f°rced expiratory flow during the middle
                                 half of the FVC [formerly called the maximum
                                 mid-expiratory flow rate (MMFR)].
FEF75%                           Instanteous forced expiratory flow after 75% of
                                 the FVC has been exhaled.

FEV                              Forced expiratory volume
FIVC                             Forced inspiratory vital capacity
fg                               Respiratory frequency.
FRC                              Functional residual capacity
FVC                              Forced vital capacity
G                                Conductance
G-6-PD                           Glucose-6-phosphate dehydrogenase
G,..                              Airway conductance
 dw
GMP                              Guanosine monophosphate
GS-CHEM                          Gas-solid chemiluminescence
GSH                              Glutathione
GSSG                             Glutathione disulfide
Hb                               Hemoglobin
Hct                              Hematocrit
HO*                              Hydroxy radical
H£0                              Water
1C                               Inspiratory capacity
IRV                              Inspiratory reserve volume
IVC                              Inspiratory vital capacity
j/
 o                               Average mucous production rate per unit area
LDH                              Lactate deyhydrogenase
LDgQ                             Lethal dose (50 percent)
LM                               Light microscopy
LPS                              Lipopolysaccharide

-------
                      LIST OF ABBREVIATIONS (continued)

MAO                              Monamine oxidase
MAST                             Kl-coulometric (Mast meter)
max ^g                           Maximum ventilation
max VOg                          Maximal aerobic capacity
MBC                              Maximum breathing capacity
MEFR                             Maximum expiratory flow rate
MEFV                             Maximum expiratory flow-volume curve
MetHb                            Methemoglobin
MMFR or MMEF                     Maximum mid-expiratory flow rate
MNNG                             N-methyl-N'-nitrosoguanidine
MPO                              Myeloperoxidase
MVV                              Maximum voluntary ventilation
NBKI                             Neutral buffered potassium iodide
(NH^gSO,                        Ammonium sulfate
NQg                              Nitrogen dioxide
NPSH                             Non-protein sulfhydryls
Og                               Oxygen •
Og-                              Oxygen radical
Og                               Ozone
P(A-a)Og                         Alveolar-arterial oxygen pressure difference
PABA                             Para-amiriobenzoic acid
P/\COg                   ,         Alveolar partial pressure of carbon dioxide
PaCQg                            Arterial partial pressure of carbon dioxide
PAN                              Peroxyacetyl nitrate
P*Qg                             Alveolar partial pressure of oxygen
PaOg                             Arterial partial pressure of oxygen
PEF       •                       Peak expiratory flow
PEFV                             Partial expiratory flow-volume curve
PG                               Prostag!andin
pHa                              Arterial pH
PHA                              Phytohemagglutinin
P.                               Transpulmonary pressure
PMN                              Polymorphonuclear leukocyte
PPD                              Purified protein derivative

                                      xii

-------
                      LIST OF ABBREVIATIONS (continued)

PS£                              Static transpulmonary pressure
PUFA                             Polyunsaturated fatty acid
R                                Resistance to flow
Raw                              Airway resistance
RBCs                             Red blood cells
RC011                            Collateral resistance
RL                               Total pulmonary resistance
RQ, R                        -    Respiratory quotient
RJ..J                              Tissue resistance
RV                               Residual volume
Sa02                             Arterial oxygen saturation
SCE                              Sister chromatid exchange
Se                               Selenium
SEM                              Scanning electron microscopy
SGaw                             Specific airway conductance
SH                               Sulfhydryls
SOD                              Superoxide dismutase
S02                              Sulfur dioxide
SPF                              Specific pathogen-free
SRaw                             Specific airway resistance
STPD                             STPD conditions (standard temperature and
                                 pressure, dry)
TEM                              transmission electron microscopy
TGV                              Thoracic gas volume
TIC                              Trypsin inhibitor capacity
TLC                              total lung capacity
TRH                              Thyrotropin-releasing hormone
TSH                              Thyroids-stimulating hormone
TV                               Tidal volume
UFA                              Unsaturated fatty acid
UMP                              Uridine monophosphate
UV                               Ultraviolet photometry
*
V^                               Alveolar ventilation
V^/Q                             Ventilation/perfusion ratio

                                      xi i i

-------
                      LIST OF ABBREVIATIONS (continued)

VC                               Vital capacity
*
VCOp                             Carbon dioxide production
Vp                               Physiological dead space
Vp                               Dead-space ventilation
^D anat                          Anatomical dead space
Ve                               Minute ventilation; expired volume per minute
Vj                               Inspired volume per minute
V^                               Lung volume
V*                                Maximum expiratory flow
VOg                              Oxygen uptake
V0«j ^Og                         Oxygen consumption
125£                             Radioactive iodine
5-HT                             5-hydroxytryptamine
6-P-GD                           6-phosphogluconate dehydrogenase

                           MEASUREMENT ABBREVIATIONS

g                                gram
hr/day                           hours per day
kg               '                ki1ogram
kg-m/min                         kilogram-meter/min
L/min                            liters/min
ppm                              parts per million
mg/kg                            milligrams per kilogram
    2
mg/m                             milligrams per cubic meter
min                              minute
ml                               milliliter
mm                               millimeter    -        .
}jg/m                             micrograms per cubic meter
urn                               micrometers
uM                               micromolar
sec                              second
                                      xiv

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


Chapter 9:  Toxicological Effects of Ozone and Other Photochemical Oxidants

Principal Authors
Dr.  Donald E, Gardner
Northrop Services, Inc.
Environmental Sciences
P.O. Box 12313
Research Triangle Park, NC  27709
Dr. Judith A. Graham
Health Effects Reearch Laboratory
MD-82
U.S. Environmental Protection Aency
Research Triangle Park, NC  27711

Dr. Susan M. Loscutoff
16768 154th Ave., S.E.
Renton, WA  98055

Dr. Daniel B. Menzel
Laboratory of Environmental Toxicology
  and Pharmacology
Duke University Medical Center
P.O. Box 3813
Durham, NC  27710

Dr. Daniel L. Morgan
Laboratory of Environmental Toxicology
  and Pharmacology
Duke University Medical Center
P.O. Box 3813
Durham, NC  27710

Dr. John H. Overton, Jr.
Health Effects Research Laboratory
MD-82
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
Mr. James A. Raub
Environmental Criteria and
  Assessment Office
MD-52
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr. Stephen C. Strom
Department of Radiology
Duke University Medical Center
P.O. Box 3808
Durham, NC  27710

Dr. Walter S. Tyler
Department of Anatomy
School of Veterinary Medicine
University of California,
Davis, CA  95616
                                      xv

-------
Contributing Authors

Mr. James R. Kawecki
TRC Environmental Consultants, Inc.
701 W. Broad Street
Falls Church, VA  22046

Dr. Frederick J. Miller
Health Effects Research Laboratory
MD-82
U.S. Environmental Protection Agency
Research Triangle Park, NC.  27711

Ms. Elaine D. Smolko
Laboratory of Environmental Toxicology
 and Pharmacology
Duke University Medical Center
P.O. Box 3813
Durham, NC  27710

Dr. Jeffrey L. Tepper
Northrop Services, Inc.
Environmental Sciences
P.O. Box 12313
Research Triangle Park, NC  27709
                                      xvi

-------
Authors also reviewed individual sections of the chapter.  The following addi-
tional persons reviewed this chapter at the request of the U.S. Environmental
Protection Agency.  The evaluations and conclusions contained herein, however,
are not necessarily those of the reviewers.
Dr. Karim Ahmed
Natural Resources Defense Council
122 East 42nd Street
New York, NY  10168
Dr. Ann P. Autor
Department of Pathology
St. Paul's Hospital
University of British Columbia
Vancouver, British Columbia
Canada  V6Z1Y6
Dr. David V. Bates
Department of Medicine
St. Paul's Hospital
University of British Columbia
Vancouver, British Columbia
Canada  V6Z1Y6

Dr. Philip A. Bromberg
Department of Medicine
School of Medicine
University of North Carolina
Chapel Hill, NC  27514

Dr. George L. Carlo
Dow Chemical, U.S.A.
1803 Building, U.S. Medical
Midland, MI  48640

Dr. Larry D. Claxton
Health Effects Research Laboratory
MD-68
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
Dr. Donald L. Dungworth
Department of Veterinary Pathology
School of Veterinary Medicine
University of California
Davis, CA  95616
Dr.  Richard Ehrlich
Life Sciences Division
Illinois Institute of Technology
  Research Institute
Chicago, IL  60616

Dr.  Robert Frank
Department of Environmental
  Health Sciences
Johns Hopkins School of Hygiene
  arid Public Health
615 N. Wolfe Street
Baltimore, MD  21205

Dr.  Milan J. Hazucha
School of Medicine
Center for Environmental Health
  and Medical Sciences
University of North Carolina
Chapel Hill, NC  27514

Dr.  Donald H. Horstman
Health Effects Research Laboratory
MD-58
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr.  Steven M. Horvath
Institute of Environmental Stress
University of California
Santa Barbara, CA  93106

Dr.  George Jf Jakab
Department of Envi ronmental
  Health Sciences
Johns Hopkins School of Hygiene
  and Publ-i-c-Health
615 N. Wolfe St.
Baltimore, MD  21205

Dr.  Robert J.,Kavlock
Health Effects Research Laboratory
MD-67
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
                                      xvn

-------
                              Reviewers (cont'd)
Dr. Thomas J, Kulle
Department of Medicine
School of Medicine
University of Maryland
Baltimore, MD  21201

Dr, Michael D. Lebowitz
Department of Internal Medicine
College of Medicine
University of Arizona
Tucson, AZ  85724
Dr. Robert C. MacPhail
Health Effects Research Laboratory
MD-74B
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
               McDonnell
               Research Laboratory
Dr. William F.
Health Effects
MD-58
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711
Dr. Myron A. Mehlman
Environmental Affairs and
  Toxicology Department
Mobil Oil Corporation
P.O. Box 1026
Princeton, NJ  08540
Dr. Harold A. Menkes
Department of Environmental
  Health Sciences
Johns Hopkins School of Hygiene
  and Public Health
615 N. Wolfe Street
Baltimore, MD  21205
                                             Dr. Phyllis J. Mullenix
                                             Forsyth Dental Center
                                             140 The Fenway
                                             Boston, MA  02115
                                             Dr. Mohammad G. Mustafa
                                             Division of Environmental and
                                               Nutritional Sciences
                                             School of Public Health
                                             University of California
                                             Los Angeles, CA  90024

                                             Dr. Russell P. Sherwin
                                             Department of Pathology
                                             University of Southern California
                                             Los Angeles, CA  90033
Dr. Robert J. Stephens
Division of Life Sciences
SRI International
333 Ravenwood Avenue
Menlo Park, CA  94025

Dr. David L. Swift
Department of Environmental
  Health Sciences
Johns Hopkins School of Hygiene
  and Public Health
615 N. Wolfe Street     i
Baltimore, MD  21205

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

Dr.- Jaroslav J. Vostal
Executive Department
General Motors Research Laboratories
Warren, MI  48090
                                     xvm

-------
                            SCIENCE ADVISORY BOARD
                    CLEAN AIR SCIENTIFIC ADVISORY COMMITTEE
     The substance of this document was reviewed by the Clean Air Scientific
Advisory Committee of the Science Advisory Board in public sessions.
                             SUBCOMMITTEE ON OZONE
                                   Chairman

                              Dr. Morton Lippmann
                                   Professor
                     Department of Environmental Medicine
                      New York University Medical Center
                            Tuxedo, New York  10987
                                    Members
Dr. Mary 0. Amdur
Senior Research Scientist
Energy Laboratory
Massachusetts Institute of Technology
Cambridge, Massachusetts  02139

Dr. Eileen G. Brennan
Professor
Department of Plant Pathology
Martin Hall, Room 213, Lipman Drive
Cook College-NJAES
Rutgers University
New Brunswick, New Jersey  08903

Dr. Edward D. Crandall
Professor of Medicine
School of Medicine
Cornell University
New York, New York  10021

Dr. James D. Crapo
Associate Professor of Medicine
Chief, Division of Allergy, Critical
   Care and Respiratory Medicine
Duke University Medical Center
Durham, North Carolina  27710
Dr. Robert Frank
Professor of Environmental Health
  Sciences
Johns Hopkins School of Hygiene
  and Public Health
615 N. Wolfe Street
Baltimore, Maryland  21205

Professor A. Myrick Freeman II
Department of Economics
Bowdoin College
Brunswick, Maine  04011

Dr. Ronald J. Hall
Senior Research Scientist and Leader
Aquatic and Terrestrial Ecosystems
  Section
Ontario Ministry of the Environment
Dorset Research Center
Dorset, Ontario  POA1EO

Dr. Jay S. Jacobson
Plant Physiologist
Boyce Thompson Institute
Tower Road
Ithaca, New York  14853
                                      xix

-------
Dr. Warren B. Johnson
Director, Atmospheric Science Center
SRI International
333 Ravenswood Avenue
Menlo Park, California  94025

Dr. Jane Q. Koenig
Research Associate Professor
Department of Environmental Health
University of Washington
Seattle, Washington  98195

Dr. Paul Kotin
Adjunct Professor of Pathology
University of Colorado Medical School
4505 S. Yosemite, #339
Denver, Colorado, 80237

Dr. Timothy Larson
Associate Professor
Environmental Engineering and
  Science Program
Department of Civil Engineering
University of Washington
Seattle, Washington  98195

Professor M. Granger Morgan
Head, Department of Engineering
  and Public Policy
Carnegie-Mellon University
Pittsburgh, Pennsylvania  15253

Dr. D. Warner North
Principal
Decision Focus Inc., Los Altos
  Office Center, Suite 200
4984 El Camino Real
Los Altos, California 94022

Dr. Robert D. Rowe
Vice President, Environmental and
  Resource Economics
Energy and Resources Consultants, Inc.
207 Canyon Boulevard
Boulder, Colorado  80302
 Dr.  George  Taylor
 Environmental  Sciences  Division
 P.O.  Box X
 Oak Ridge National  Laboratory
 Oak Ridge,  Tennessee 37831

 Dr.  Michael  Treshow
,Professor
 Department  of  Biology
 University  of  Utah
 Salt Lake City,  Utah 84112

 Dr.  Mark J.  Utell
 Co-Director, Pulmonary  Disease Unit
 Associate Professor of  Medicine  and
   Toxicology in  Radiation Biology
   and Biophysics
 University  of  Rochester Medical
   Center
 Rochester,  New York  14642

 Dr.  James H. Ware
 Associate Professor
 Harvard School of  Public Health
 Department  of  Biostatisties
 677 Huntington Avenue
 Boston Massachusetts  02115

 Dr.  Jerry Wesolowski
 Air and Industrial  Hygiene Laboratory
 California  Department of Health
 2151 Berkeley  Way
 Berkeley, California  94704

 Dr.  James L. Whittenberger
 Director, University of California
   Southern  Occupational Health Center
 Professor and  Chair, Department  of
   Community and  Environmental Medicine
 California  College of Medicine
 University  of California - Irvine
 19772 MacArthur  Boulevard
 Irvine, California  92717

 Dr.  George  T.  Wolff
 Senior Staff Research Scientist
 General Motors Research Labs
 Environmental  Science Department
 Warren, Michigan  48090
                                       xx

-------
                         PROJECT TEAM FOR DEVELOPMENT
                                      OF
        Air Quality Criteria for Ozone and Other Photochemical Oxldants
Ms.  Beverly E.  Tilton, Project Manager
  and Coordinator for Chapters 1 through 5, Volumes I and II
Environmental Criteria and Assessment Office (MD-52)
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Mr.  Norman E. Chi Ids
Environmental Criteria and Assessment Office (MD-52)
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr.  J.H.B. Garner
Coordinator for Chapters 7 and 8, Volume III
Environmental Criteria and Assessment Office (MD-52)
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

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

Mr.  James A. Raub
Coordinator for Chapters 10 through 13, Volumes IV and V
Environmental Criteria and Assessment Office (MD-52)
U.S. Environmental Protection Agency
Research Triangle Park, NC  27711

Dr.  David T. Tingey
Coordinator for Chapter 7, Volume III
Environmental Research Laboratory
U.S. Environmental Protection Agency
200 S.W. 35th Street
Corvallis, OR  97330
                                       xxi

-------

-------
     9.   TOXICOLOGICAL  EFFECTS OF  OZONE  AND OTHER  PHOTOCHEMICAL OXIDANTS
9,1  INTRODUCTION
     This chapter  discusses the effects  of  ozone on experimental animals.
Carefully controlled studies  of  the effects  of ozone on animals are particu-
larly important  in elucidating subtle effects not easily found  in man through
epidemiological  studies and in identifying chronic toxicity not apparent from
short-term controlled human exposures.   Animal  studies allow investigations
into the  effects of ozone  exposure over a  lifetime,  uncomplicated  by the
presence of other  pollutants.   In  the animal experiments  presented  here,  a
broad range of ozone  concentrations has been studied  but  emphasis  has been
                                       3
placed  on  recent studies at  1960  ug/m  (1 ppm) of  ozone  or  less.   Higher
concentrations have been cited when the data add to an understanding of mechan-
isms. Concentrations of 1 ppm or greater cannot  be  studied ethically  in man
because of the toxicity of even short-term exposures.
     A majority of the literature describes the effects of ozone on the respira-
tory tract, but  extrarespiratory system effects have  now  been  noted and are
documented in this chapter.   Most of the studies utilize invasive methods that
require sacrifice  of  the  animals on completion of the  experiment; thus, the
studies would be impossible to perform in human subjects.   Noninvasive methods
of examining most of these endpoints are not readily available.
     Emphasis  has  been  placed on the more recent  literature published  after
the  prior  criteria document  (U.S.  Environmental  Protection Agency, 1978);
however, older literature has been reviewed .again in  this chapter.  As more
information on the toxicity of ozone becomes available, a better understanding
of earlier  studies  is  possible and a more detailed and comprehensive picture
of ozone toxicity is emerging.  The literature used in developing this  chapter
is, set  out  in a series of  tables,  Not all  of..the. literature cited in the
                     \
tables  appears in the  detailed discussion of the text,  but  citations are
provided to give the reader more details on the background from which the text
is drawn.
     In  selecting  studies for consideration, a detailed review of each paper
has  been completed.  This review included an evaluation of the  exposure methods;
the  analytical  method  used  to determine the chamber ozone concentration; the
                                    9-1

-------
calibration of the ozone monitoring equipment and the analytical methods used
(wherever possible);  the species,  strain, age and physical characteristics of
the animals; the technique used for obtaining samples;  and the appropriateness
of the technique used to measure the effect.  In interpreting the results,  the
number of animals  used,  the appropriateness and results  of  the statistical
analysis, the degree  to  which the results conform with past studies, and the
appropriateness of  the interpretation  of the  results  are  considered.   No
additional statistical analysis  beyond  that reported by  the author  has been
undertaken.   Unless otherwise  stated,  all statements of  effects in  the text
are statistically significant  at p < 0.05.  Many reports, especially in the
older literature,  do not present sufficient information to permit the assessment
described above.  However,  should a particular study not meet  all  of these
criteria, but  provide reasonable  data  for consideration, a disclaimer is
provided in the text and/or tables.
     In this chapter,  a  discussion of the  regional  respiratory dosimetry of
ozone in  common laboratory animal species is presented and compared to human
dosimetry.  Morphological  alterations of  the lungs  of  animals exposed to 03
are described, followed  by the effects of ozone on the pulmonary function of
animals.   The biochemical  alterations  observed in the  ozone-exposed animals
are then related to morphological changes and to potential mechanisms of toxi-
city and  biochemical  defense mechanisms.   The  influence  of  dietary  factors,
such as vitamins E and C, in animals is discussed with consideration of poten-
tial roles  in  humans.   It should be stressed,  however, that no evidence for
complete  protection  against ozone toxicity has been found  for any  factor,
dietary or. therapeutic.  The effects of ozone on the defense mechanisms of the
lung against respiratory infectious agents are discussed using the infectivity
model system and  effects  on alveolar macrophages as examples of experimental
evidence. This  section is  followed  by a discussion of ozone  tolerance in
animals.   Last, the  effects of ozone on  a  number  of extrarespiratory organ
systems  are  discussed to  provide insight  into potential effects  of ozone
inhalation in the respiratory system beyond those now well documented.
     A brief discussion  of the available literature on the  effects  of other
oxidants likely to occur in polluted air as a result of photochemical reactions
or other  sources  of  pollution-is presented.  Peroxyacetyl nitrate,  hydrogen
peroxide, and automobile exhaust are the principal  pollutants studied in these
experiments. This section  is short because  of the general lack of information
                                    9-2

-------
in this area,  but  its brevity does not necessarily reflect a general  lack of
importance.
     A summary is  provided  for all of the sections of the chapter to  set the
tone for a  clearer understanding of the effects  of  ozone on animals.  The
major emphasis of  this  chapter is to provide  evidence  for the toxicity of
ozone which can  not,  ethically or practically, be obtained from the study of
human subjects.  The  overall  health effects of ozone can  be judged from three
types of studies:  animal exposures, controlled human exposures, and epidemio-
logical studies  of adventitious  human exposures.   No single method alone is
adequate for  an  informed judgment, but  together  they provide a  reasonable
estimate of the human health effects of ozone on man.
9.2  REGIONAL DOSIMETRY IN THE RESPIRATORY TRACT
     A major  goal  of environmental  toxicological studies on  animals  is the
eventual quantitative extrapolation of results to man.  One type of information
necessary to  obtain  this  goal  is  dosimetry, which is  the  specification  of the
quantity of inhaled material, -in this instance ozone (0~), absorbed by specific
sites in  animals  or  man.   This information is  needed because the  local dose
(quantity of  03 absorbed per  unit  area),  along with cellular  sensitivity,
determines the  type  and extent of  injury.  At  this time, only dosimetry is
sufficiently  advanced  for discussion  here.  Until both  elements are advanced,
quantitative extrapolation cannot be conducted.
     At present,  there are  two approaches to  dosimetry, experimental  and
deterministic mathematical modeling.  Animal experiments have been carried out
to obtain direct measurements of 0- absorption; however, experimentally obtaining
local lower respiratory tract  (tracheobronchial  and pulmonary regions)  uptake
data  is currently extremely difficult.   Nevertheless,  experimentation is
important in assessing concepts and hypotheses, and in validating mathematical
models that can be used to predict  local doses.
     Because  the  factors affecting the  transport  and absorption of 03 are
general to  all  mammals, a model that  uses appropriate species and/or disease-
specific  anatomical  and ventilatory parameters  can  be used  to describe  0, ab-
sorption  in the species and in different-si zed, aged, or diseased members of
the  same  species.   Models may also be  used to  explore processes or factors
which cannot  be studied experimentally, to identify areas needing additional
                                    9-3

-------
research,  and to test our understanding  of 03 absorption in the respiratory
tract.

9.2.1  Absorption in Experimental Animals
     There have  been very few experiments  in which measurements  of the regional
uptake of  0, or other reactive  gases  have been determined.  Of the several
results published,  only  one is  concerned  with the uptake of 03 in  the lower
respiratory  tract; the others deal with nasopharyngeal uptake.
9.2.1.1  Nasopharyngeal  Absorption.   Nasopharyngeal  removal  of 0~  lessens
the  quantity of 03  delivered to the  lung and must be accounted for  when
estimating the  03  dose responsible for observed  pulmonary effects.  Vaughan
et al. (1969) exposed the isolated upper  airways  of beagle dogs to 03 at  a
continuous  flow of  3.0  L/min  and collected the gas  below the larynx  in a
plastic (mylar)  bag.  One-hundred percent  uptake by the nasopharynx was reported
for concentrations of 0.2 to 0.4 ppm.  Using a different  procedure, Yokoyama and
Frank  (1972) observed 72 percent  uptake  at 0.26 to  0.34 ppm  (3.5  L/min  to
6.5 L/min flow rate).  They also replicated the procedure of Vaughan et al. and
found that 03 was absorbed on the mylar bag wall.  This may account for the dif-
ference between  the  observations of Yokoyama and Frank (1972) and of Vaughan et al.
(1969).
     Yokoyama and  Frank  (1972)  also observed a decrease  in the percent uptake
due to increased flow rate, as well as to  increased 0, concentration.  For  example,
with  nose  breathing and an 03  concentration  of 0.26 to  0.34 ppm,  the  uptake
decreased  from 72 percent to 37  percent for a  flow rate increase from  3.5 to 6.5
to 35  to 45 L/min.   An increase in concentration from 0.26 to 0.34  to  0,78 to
0.80 ppm decreased nose  breathing uptake  (3.5  to 6.5  L/min flow rate)  from  72 per-
cent to 60 percent.  Thejr data, however,  indicate that the trachea! concentra-
tion  increases  with increased nose or mouth concentrations.   They also demon-
strated that the concentration of 03 reaching  the trachea depends heavily on the  ....
route  of breathing.   Nasal  uptake significantly exceeded oral  uptake  at flow
rates  of both 3.5 to .6.5 and  35 to 45 L/min.  For  a given flow rate, nose
breathing  removed 50 to  68 percent more 03 than did mouth breathing.
     Moorman et  al.  (1973) compared the loss of 0, in the nasopharynx  of acutely
e                                             *    *•*
and chronically  exposed  dogs.  Beagles chronically exposed (18  months) to 1 to
3 ppm of 0~  under various daily  exposure  regimes had  significantly higher tra-
chea! concentrations of  03 than  animals tested after  1 day of exposure to cor-
responding regimes.  Moorman et  al. (1973) suggested  that the differences were
                                    9-4

-------
due to physiochemical alterations of the mucosal lining in the chronically ex-
posed beagles.  When dogs were exposed  for 18 months to 1 ppm for 8 hr a day,
they had significantly lower trachea! values than those continuously exposed.
The average trachea! concentration  (0.01 ppm)  for the acutely exposed group,
however, was not significantly different from that (0.023 ppm) of the 8 hr/day
chronic exposure group,  when  the relative insensitivity of the Mast 03 meter
(unmodified) used to measure  the responses is  taken into account.  Thus, at
levels  of  1.0 ppm  or  less,  there is no  significant evidence  that chronic
exposure would result in trachea! 0- concentrations significantly greater than
those observed with acute exposure.
     Nasopharyngea!  removal of  0- in rabbits and guinea pigs was studied by
Miller et al.  (1979) over a concentration range of 0.1 to 2.0 ppm.   The trachea!
0., concentration in  these two species was markedly  similar  at a  given  inhaled
concentration and was  linearly  related to the chamber concentration that was
drawn unidirectionally through  the  isolated upper airways.   Ozone removal in
the nasopharyngeal  region was  approximately 50 percent in  both  species over
the concentration range  of  0.1 to 2.0 ppm.   The positive correlation between
the trachea!  and chamber concentrations  is  in  agreement with Yokoyama and
Frank (1972).   Caution needs to  be  exercised in applying the above  results to
relate  ambient  and  trachea!  concentrations of  0- since the effects.of naso-
pharyngeal  volume  and  the cyclical  nature  of  breathing  are not taken into
account.  For  example,  if the tidal volume was less than the nasopharyngeal
volume  and  convection  was the only  process of  axial  transport,  then  no  0~
would be  delivered  to  the trachea!  opening, regardless of  the percent uptake
measured for unidirectional flow.
9.2.1.2  Lower  Respiratory Tract  Absorption.  Morphological studies on animals
suggest that  03 is  absorbed  along the  entire respiratory tract;  it  penetrates
further into  the  peripheral  nonciliated  airways as inhaled 0- concentrations
increase  (Dungworth  et al.,  1975b).  Lesions were  found consistently  in  the
trachea and proximal bronchi and  between  the junction of the conducting airways
and the gaseous exchange area; in both  regions, the severity of  damage decreases
distally.   In addition,  several  studies have  reported  the most severe or
prominent lesions to be  in the centriacinar region  (see section  9.3).
     No experiments  determining  0-  tissue dose  at the generational or  regional
level  have  been reported;  however,  there is one experiment  concerned with the
uptake  of 0-  by the lower respiratory tract.   Removal  of 0- from inspired air

                                     9-5

-------
by the  lower  airways was measured by  Yokoyama  and  Frank (1972)  in  dogs  that
were mechanically ventilated through a tracheal cannula.  In the two ranges of
Og concentrations  studied,  0.7 to 0.85 ppm and 0.2 to 0.4 ppm,  the rate of
uptake was  found  to vary between 80 and 87 percent when the tidal volume was
kept constant and the respiratory pump was operated at either 20 or 30 cycles/
min.   This  estimate of uptake applies to  the  lower respiratory tract as a
whole; it does not describe uptake of 03 by individual regions or generations.

9.2.2  Ozone Dosimetry Models
9.2.2.1  Modeling Nasal Uptake.   LaBelle et al. (1955)  considered the  absorp-
tion of  gases  in  the nasal passages to  be similar to  absorption on wetted
surfaces of distillation equipment and scrubbing towers and applied the theory
and models  of  these devices to the  nasal  passages  of rats.  By associating
biological parameters  of  rats  with the chemical engineering device  parameters
of the  model,  they calculated the percent of penetration of several gases to
the lung.   They  concluded that Henry's law constant is the major variable in
determining penetration.  Based on these calculations, The National  Academy of
Sciences (National  Research  Council, 1977) concluded that the model predicts
99 percent penetration for Og.  This is much more than that measured by Yokoyama
and Frank (1972) or by Miller et al.  (1979).   Several possible reasons for the
differences were  discussed  (National Research Council,  1977), but  the major
factor was  considered  to  be  that  the model does not account for  the reactions
of 03 in the mucus and epithelial tissue.
     Aharonson et al.  (1974)  developed a model  for  use in analyzing data from
experiments on the  uptake of vapors by the nose.   The model was based on the
assumptions of quasi-steady-state flow,  mass balance, and proportionality of
flux  of a trace  gas at the  air-mucus interface  to the  gas-phase  partial
pressure of the trace  gas and a  "local uptake  coefficient"  (Aharonson  et al.,
1974).  The model was applied to data from their own experiments on the removal
of acetone  and ether in dog  noses.  They  also applied the model to the 03
uptake  data of Yokoyama and Frank (1972) and concluded that the uptake coef-
ficient  (average  mass  transfer coefficient)  for Og» as well as for the other
gases considered, increases with increasing air flow rate.
9.2.2.2   Lower Respiratory Tract Dosimetry Models.   There  are three  models
for which  published results  are available.   The model  of  McJilton et al.
(1972)  has  been  discussed and simulation  results  for 0^ absorption in each
generation of the human lower respiratory tract are available (National Research
                                    9-6

-------
Council 1977;  Morgan and Frank, 1977).  Two  models  have been developed by
Miller and co-workers.  A detailed description of the formulation of the first
and earliest mathematical  model  of Miller and co-workers  is  found  in Miller
(1977) and major features are given in Miller et al.  (1978b) and Miller (1979).
Results,  using this model,  of simulations of the  lower respiratory tract
absorption of  Q,  in humans, rabbits, and  guinea  pigs  are in Miller (1977,
1979)  and Miller et  al.  (1978b).   The formulation of the  second model of
Miller  and  co-workers, as well as  results of simulating lower  respiratory
tract absorption of 0- in humans, appears in Miller et al. (1985).
     Because all  of the above models were  developed to simulate the local
absorption of  03,  they have much in  common.  This  is  especially true  with
respect to the following  areas:   formulation  of 0^ transport  in  the airspaces
or  lumen  of  the airways,  use of morphometric models of the lower respiratory
tract,  and inclusion of a liquid lining that coats  the tissue walls of the
airspaces or lumen of the airways.
     In each model  the descriptions of 03  transport and absorption in the
lumen  are based on  a one-dimensional  differential  equation  relating axial
convection, axial dispersion or diffusion, and the loss of 0- by absorption at
the gas-liquid  interface.  The use of a one-dimensional approximation has been
very common  in modeling the transport of  gases such as  0^, N^,  etc., in the
lower  respiratory  tract (see Scherer et al., 1972;  Paiva, 1973; Chang and
Farhi,  1973; Yu, 1975; Pack et al., 1977; Bowes et al., 1982). .The approxima-
tion is appropriate  for 0^ as well.
     The  models of Miller and co-workers took  into  account effective axial
dispersion in  the  airways by using an effective dispersion coefficient based
on  the results of Scherer et al. (1975).  McJilton's model did not take this
factor  into  account (Morgan and Frank, 1977).  However,  this may  not be an
important difference since Miller et al. (1985) report that results are little
affected  by different values of the coefficient.  Also, Pack  et al. (1977) and
Engel  and Macklem (1977) reported results  that indicate an insensitivity  of
airway  concentrations to  the effective dispersion coefficient.
     Airway or morphometric zone models such as  those of Weibel (1963) and
Kliment (1973) were used to define the lengths, radii, surface areas, cross-
sectional  areas,  and volumes of the airways and air  spaces of each  generation
or  zone.   The  breathing pattern was  assumed  sinusoidal;  however,  dimensions
were held constant  throughout the breathing cycle.  The physical properties of
                                    9-7

-------
the liquid  lining  were assumed to be those  of water.   The lining thickness
depended on generation or zone, being thicker in the upper airways than in the
lower.
     The flux  of Og from the  lumen  or  airspaces to the  liquid  lining was
defined in  each  model  in terms of a mass-transfer coefficient.  McJilton et
al. (1972) made the assumption that radial mass transfer was controlled by the
absorbing medium and estimated the transfer coefficient from empirical data on
the physical  properties (not  chemical)  of the medium and  of  03 (National
Research  Council,  1977).   In  the first  model of Miller and co-workers
(Miller, 1977; Miller  et al.,  1978b) the radial dependence of the luminal 03
concentration was  assumed to vary quadratically with the radius.  From this
formulation, the gas-phase  mass  transfer was  determined.  For their  later
model,  the  gas  phase  mass  transfer  coefficient was defined in  terms of  a
Sherwood number.   In  both models the gas phase coefficient was combined with
the mass transfer  coefficient  of the absorbing medium (which depended on the
chemical and physical  properties  of the absorbing medium  and of 03) to obtain
the overall  transfer  coefficient.  However, Miller et al.  (1985)  conclude,
based on the data available for the absorbing medium,  that radial mass transfer
is controlled by the medium, making specification of the gas phase mass trans-
fer coefficient  unnecessary.
     The main  differences in the models are the mechanisms of absorption and
their formulation.  In the  model of McJilton  et al. (1972) and  in the early
model of Miller  and co-workers (Miller, 1977;  Miller et al., 1978b) there is
only  one compartment,  the  liquid lining, which can absorb  unreacted  Qg.  In
the later model  of Miller et al.  (1985) there are three absorbing compartments,
liquid  lining, tissue, and  capillary or blood (in the pulmonary region where
the air-blood  tissue  barrier is  very thin).   Further, (k is known to react
chemically with  constituents of the absorbing medium(s).   This aspect, included
by Miller and co-workers, was not included in McJilton's model.  The inclusion
resulted in  significant differences between  the tissue dose pattern curves  in
the tracheobronchial  region predicted by  the models.  In  addition, McJilton's
model predicts a dose curve (equivalent to a tissue dose curve because of no
mucous  reactions)  in  the tracheobronchial region that has its maximum at the
trachea and decreases  distally to the thirteenth or fourteenth generation (see
Figure  7-5  in  National Research  Council,  1977).  By contrast, the models  of
Miller  and  co-workers  predict  the tissue dose to be a minimum at the trachea
and to  increase  distally to the pulmonary region.
                                    9-8

-------
     The concentration of  ozone in tissue and at the liquid-tissue interface
was assumed to  be zero by McJilton et al. (1972) (National Research Council,
1977; Morgan  and  Frank, 1977) and  in the  first model  of Miller  and  co-workers
(Miller, 1977; Miller et al., 1978b).   The interpretation was that this boundary
condition means that  03  reacts  (chemically) instantaneously when it reaches
the tissue.   Miller et al.  (1978b)  define the tissue  dose  as that quantity of
0~ per  unit area  reacting  with  or  absorbed by the tissue at the liquid-tissue
interface.
     The first  model  of  Miller  and co-workers took into account the reaction
of 03 with  the  unsaturated fatty acids (UFA) and ami no acids in the mucous-
serous  lining.  Reactions  of 03 with other components  (such as  carbohydrates)
were not  included in  the model  because of insufficient information (Miller,
1977; Miller  et al.,  1978b).   The  CL-UFA and  Ov-amino acid reactions were
assumed fast  enough  so that an instantaneous reaction  scheme  based on that
outlined in Astarita (1967) could be used.  The scheme required the specifica-
tion of the  production rate of the UFA and  ami no acids in each mucous-lined
generation.   These rates were  estimated  by  using trachea! mucous flow data,
the surface area  of  the tracheobronchial  region, the  concentrations  of the
specific reactants known to react with Q,, and the assumption that the produc-
tion rate decreased distally (Miller, 1977; Miller et al., 1978b).
     Although the instantaneous reaction  scheme is a good preliminary approach
to treating 0, reactions in the mucous-serous lining, its use is not completely
justifiable.  Second-order rate constants of 03 with some of the UFA present
in mucus indicate that although they are  large (Razumovskii and Zaikov, 1972),
they are less than the diffusion-limited  rates necessary for the instantaneous
reaction  scheme.   Experimental  evidence  (Mudd et al.,  1969) suggests that the
reactions of  0- with amino acids  are  very rapid.   Rate constants for these
              »3        .         ...
reactions and others are not known; thus,  the information available is scanty,
which makes  the  specification  of a reaction mechanism  or reaction scheme
difficult and assumptions  necessary.
     The  approach to  chemical   reactions  used in the  later model (Miller et
al., 1985)  goes a long,way in  addressing the above criticisms.  The reaction
of 0~ with biochemical constituents is assumed to be bimolecular; however, the
concentration of  the  constituents is considered to be large enough so as not
to be depleted  by the reactions.  Hence,  the model uses a pseudo first order
reaction  scheme in which the pseudo first order rate constant  is the product
                                    9-9

-------
of  the  bimolecular rate constant  and  the concentration of the biochemical
constituents that react with Q~.
     For modeling purposes, Miller et al. (1985) consider that only the reaction
of  ozone with  the UFA is important, using the 03~oleic acid rate constant of
Razumovskii and  Zaikov  (1972)  for the 03-UFA  reaction.  They point out that
although 03 reacts with ami no acids and other constituents, rate constants are
not known  and  that Bailey (1978) estimates the reaction of 0~ with UFA to be
   3
~10  times  faster than  the 03-amino acid reaction, justifying omitting amino
acids from consideration.

9.2.3  Predictions of Lower Respiratory Tract Ozone Dosimetry Modeling
     The predictions  of  lower  respiratory tract dosimetry  models are  reviewed
by illustrating the results of simulations, by comparing predictions to experi-
mental observations, and by describing uses for dosimetry models.
     The following  discussion  of modeling results of lower respiratory tract
absorption  is  based mainly on  simulations using the  first  model of Miller and
co-workers  (Miller, 1977;  Miller et al., 1978b).  This  is because the model
includes the important effects of 03 reactions  in the mucous-serous lining and
because  simulations of 03 absorption  in laboratory animals are available.
     Simulations  of 03  absorption in different animals can be carried out by
modifying input parameters of the computer program that solves the mathematical
equations.  These input  parameters,  which characterize  an  animal,  include the
number and  dimensions of the airways, tidal  volume,  length of time  of one
breath,  etc.   The airway and alveolar dimensions  of Weibel  (1963) were used
for the  simulation of 03 uptake  in  humans.   For the rabbit and guinea pig,
Miller and  co-workers used the morphometric  zone  models  of Kliment  (1973).
The zone model  is a  less detailed model  than the generationally based airway
model of Weibel  (1963) since more than one generation corresponds to  a zone in
an  animal; they were  used because they were the only complete (tracheobronchial
and pulmonary regions) "airway" models available at the time.  However, Schreider
and Hutchens  (1980) criticize  the guinea pig  model  of  Kliment  (1972,  1973)  as
having a lung volume  that  is too  low, suggesting the possibility of incomplete
casts.   Since  the same method also  was  used  by Kliment (1972,  1973)  for  the
rabbit,  this criticism may also apply to  this model.
     To  illustrate simulation  results,  two  aspects  of the simulations by
Mi77er  and co-workers are considered:   (1)  the effect of various  trachea!
                                    9-10

-------
concentrations on the  tissue  dose pattern (tissue dose as a function of zone
or generation) in guinea pigs and rabbits; and (2) the similarity between the
dose patterns of guinea pigs,  rabbits, and humans.

9.2.3.1   I]1ustration of Dosimetry Simulations.    Figure  9-1  is a  set of
plots of the  tissue  dose- for one breath  versus zone for  various trachea! 0,,
concentrations for the  rabbit and the guinea pig.  All curves  have  the same
general characteristics.  Independent of  the inhaled concentration,  the model
predicts that the first surfactant-lined zone (first non-mucous-lined or first
zone  in  the pulmonary  region),  zone  6,  receives the  maximal  dose  of  Q~.
Although the model predicts  uptake of 0Q by respiratory  tissue  (zones  6, 7,
                                                                    3
and  8)  for all  trachea!  concentrations studied  (62.5 to 4000  ug/m ),  the
penetration of 0^ to the tissue  in the airways lined by mucus depended  on the
trachea! concentration  and the  specific  animal   species.   For  example, as
                                                                        2
illustrated in Figure  9-1 for the tracheal 03 concentration  of  1000  jjg/m  ,  no
0- reaches the airway tissue of the rabbit until   zone 3, whereas 0-  is predicted
to penetrate  to  the  tissue in guinea pig airways in all  zones.  However, at
                                                                 3
the two lowest tracheal concentrations plotted, 250 and 62.5 ug/m ,  no penetra-
tion occurs until zones 4 and 6, respectively,  for both animals.  The dependence
of penetration on  tracheal  0, concentration is a result  of the instantaneous
reaction scheme  used  to describe chemical reactions.   Penetration does occur
in the  simulation  of uptake in  humans using the  newer model of Miller and
co-workers  (1985), as  depicted  in  Figure  9-2, and can  be  expected to occur  in
future animal simulations.
     The similarity  of the predicted  dose patterns  in  rabbits and guinea  pigs
extends to  the  simulation of 0- uptake in  humans.   Figure 9-2 compares the
                                                                            3
tissue  dose  for  the  three species for a  tracheal concentration of  500  ug/m
(0.26 ppm)  using results of the first model for  the three  species  and the
newer model  results  for man.   In Figure 9-2, the-guinea pig and rabbit  tissue
doses are  plotted  in the form of  a  histogram to  allow a  comparison  to  human
dosimetry data that are expressed as  a function of airway generation.
     For the earlier model, the  dose  patterns of  the three- species peak at  the
first  surfactant-lined  zone  (6) or in generation 17  (which is in zone 6).
Also, 03  penetrates  to the  tissue  everywhere in the pulmonary region (zones >
5  and generations  >  16); however, 03 is  not predicted to  penetrate to the
tissue before zone 3 for the  rabbit and guinea pig or before zone 5  (generations
12-16)  for man.
                                    9-11

-------
   104
              i    r~i    i    i   i   r~i    r
   105
m
Q)
O)
a.

O  106
u.
O
ui
CO
O
Q
Ul

CO
g  10-7
   10-8
I    I    I    I   I    I    I
                                                    TRACHEAL
                                                    O3 CONG.
                              • 4000
                              A1000
                              •  250
                              •   62.5
                                                          ppm
                                    2.041
                                    0.510
                                    0.128
                                    0.032
               0

         TRACHEA
 2   3

BRONCHI-
               5678 MODEL ZONE
               BR°N» A.D. A.S. MORPHOMETRIC ZOIME
                                 ' CHIOLES
          Figure 9-1. Predicted tissue dose for several trachea! O3
          concentrations for rabbit (	) and guinea pig (	).
          See text for details. (A.D. = alveolar duct; A.S. = alveolar
          sac).

          Source: Adapted from Miller et al. (1978b).
                               9-12

-------
    105
«J
l_
&
  i
O
LL
O
LU
C/3
O
d
UJ

C/5
    106  	
~   107
    1.0-8
                 TRACHEOBRONCHIAL
                                                                       GENERATION

                                                                       —ZONE
       Figure 9-2. Tissue dose versus zone for rabbit (o) and guinea pig (•);
       and tissue dose versus airway generation for human (	, earlier
       model;	, newer model). Tracheal O3 concentration is 500 ^g/m3
       (0.26 ppm). See text for details.

       Source: Guinea pig, rabbit and earlier human simulations adapted from
       Miller et al. (1978b); newer model results for human adapted from
       Miller etal. (1985).
                                      9-13

-------
     All of the earlier simulations presented by Miller and co-workers (Miller,
1977; Miller et al., 1978b; Miller, 1979) share the following characteristics:
(1)  the  maximal  tissue  dose  occurs in the  first  surfactant-lined  zone or
generation; (2) 0,  penetrates  to tissue everywhere in  the  pulmonary region
(zones >5), decreasing  distally from the maximum; and  (3)  the onset of 0,
penetration to mucous-lined tissue,  as well as dose  in general,  depends on
trachea! 0, concentration, animal  species,  and the breathing pattern.  These
general  characteristics  of tissue dose pattern are  independent of the two
airway models used.
     Results using  the  newer model to  simulate tissue dose  in  humans  are also
illustrated in  Figure 9-2.   Comparison  of  dose  values for the two  human
simulations show.the most notable differences in the conducting airways (genera-
tions 0 to 16).  The new and old results are similar in that both models predict
relatively low doses  in the  upper airways,  a maximum in the first pulmonary
generation, and then  a  rapid decline in dose distally.  There  is no  reason to
assume that these features would be missing in simulation results for laboratory
animals using the newer model.
9.2.3.2   Comparison of Simulations to Experimental Data.   There are no quanti-
tative experimental  observations with which to compare the results of modeling
the  local  uptake  of 03 in the  lower respiratory  tract.  Yokoyama and Frank
(1972) observed 80 to 87 percent uptake of 03 by the isolated lower respiratory
tract of dogs.  With  the earlier model, Miller et al.  (1978b) predict a 47
percent uptake for humans; the newer model  predicts 89 percent (Miller et al.,
1985).  Because of  differences  between the  two species, comparing the experi-
mental and simulated results is most likely inappropriate.
     Morphological  studies  in animals  report  damage throughout  the lower
respiratory tract.  Major damage, and  in some cases the most severe damage, is
observed to occur at  the junction between the conducting airways and the gas
exchange region,  and to decrease  distally  (see  Section 9.3.1.1).   For the
animals simulated by  Miller  and co-workers, the maximal tissue dose of 0, is
predicted  to occur  at this junction,  with  the dose curve decreasing  rapidly
for more distal regions.  Thus, in the pulmonary region, the model results are
in qualitative agreement with experimental  observations.
     Damage is also observed in the trachea and bronchi of animals (see Sections
9.3.1.1.1.2 and  9.3.1.1.1.3).   In the animals modeled, the early model of
Miller and co-workers either  predicts  significantly less tissue dose  in the
                                    9-14

-------
upper airways compared  to  the dose in the first zone of the pulmonary region
or it predicts no penetration to the tissue in the upper portion of the conduc-
ting airways  (see  Figure  9-2).   Based on  simulation results using the newer
model for  humans  (Miller  et al., 1985), one  can  infer that this model will
predict non-zero tissue doses in the  upper airways of  animals, but that these
doses also will  be significantly less  than  predicted  for the ceritriacinar
airways.  The observations of damaged upper and lower respiratory tract airways
in the same animal and the predictions of significantly different tissue doses.
in the  two regions appear  inconsistent.  However, much of the reported damage
in the  trachea and bronchi is associated  with  the  cilia of ciliated cells,
which in current  model  formulations are not  part of the tissue.  The cilia
extend into the hypophase (perciliary) portion of the mucous-serous layer, and
the dosimetry models  do not distinguish the cilia of the ciliated cells as a
separate component of this  layer.   Thus, relatively low  predicted tissue  dose
should  not be interpreted  as predicting no damage  to  cilia.   Likewise, the
frequent reporting of cilia being damaged  following 03 exposure  should not be
interpreted necessarily as  an indication of  03 tissue dose  since the model
definition  of tissue  does  not currently include cilia.   The  inclusion of a
"cilia compartment" in  future dosimetry models may be helpful.   There are also
other factors that complicate our understanding of ozone toxicity, such as the
possibility of 0- reaction products being toxic or differences in cell sensiti-
vity that  may prevent  explanations of  observed effects  based on dosimetry
modeling alone.
9.2.3.3    Uses of Predicted Dose.  Model-predicted doses can be used to estimate
comparable  exposure  levels  that  produce  the same dose  in different species  or
different  members of  the same species for  use in comparing toxicological data.
One can simulate tissue dose for several species for the same time for a range
of tracheal 0-  concentrations.   The doses for a specific zone or generation,
for  each  species,  can be plotted versus  tracheal or ambient Q~ concentration.
By using  such plots  and information  on  nasopharyngeal  removal,  the ambient
concentration necessary to produce the same  dose in different species can be
estimated.  Also, the relative quantity of 0~ delivered  to a zone or generation
in a given species for the  same time span and ambient  concentration can be
predicted  from the same graph.   If the  same  biological  parameters have not
been measured in these  species at dose-equivalent exposure levels, the procedure
can  be  used to scale  data and to design new studies to fill gaps in the current
data base.
                                    9-15

-------
     To  illustrate the above  procedure,  Miller (1979) calculated  exposure
levels of  Og,  giving the  same  respiratory bronchiolar  dose  in  rabbits,  guinea
pigs, and  man.   Considering the  discussion on  nasopharyngeal removal  (Section
9.2.1.1) and  the question  concerning  the guinea pig and rabbit  anatomical
models  (Section  9.2.3),  these calculations are  mainly useful  for illustra-
tive purposes.
9.3  EFFECTS OF OZONE ON THE RESPIRATORY TRACT
9.3.1  Morphological Effects
     The many  similarities  and differences in the  structure  of  the  lungs  of
man and  experimental  animals were the subject of a recent workshop  entitled
"Comparative Biology of the Lung:  Morphology", which was sponsored by the Lung
Division of the National Institute of Blood, Heart,  and Lung Diseases (National
Institutes of  Health,  1983).   These anatomical differences complicate but do
not necessarily prevent qualitative extrapolation of risk to man.  Moreover,
because  the  lesions due to 0, exposure  are similar in many of  the  species
studied  (see Table 9-1),  it appears likely that many  of the postexposure
biological processes of animals could also occur in man.
9.3.1.1   Sites Affected.   The  pattern  and  distribution  of  morphological
lesions  are  similar in the species  studied.   Their precise characteristics
depend on  the  location (distribution) of  sensitive  cells and on the  type  of
junction between the conducting airways and the gaseous exchange area.
     The upper  or  extrathoracic  airways  consist of  the  nasal  cavity,  pharynx,
larynx,  and  cervical  trachea.   Except for a few sites, the lining epithelium
is ciliated, pseudostratified  columnar,  with mucous (goblet) cells; it rests
on a  lamina  propria or s.ubmucosa that contains  numerous  mucous, serous, or
mixed glands and  vascular plexi.  Sites with differing structure include the
vestibule  of the  nasal  cavity and portions  of the  pharynx and larynx, which
tend to  have stratified squamous epithelium, and those portions of the nasal
cavity lined by olfactory epithelium, which contain special  bipolar  neurons
and glands associated with the sense of smell.   Significant morphological dif-
ferences exist  among  the  various animal  species used for 0, exposures as well
as between most of them and man (Schreider and Raabe, 1981;  Gross et al., 1982).
With the exception of the cervical  trachea, these  structures have received
little attention with respect to 0-  sensitivity,  but 0., removal  through "scrub-
bing"  has been studied (Yokoyama and Frank, 1972;  and Miller et al., 1979).
                                     9-16

-------
                                                                TABLE 9-1.   MORPHOLOGICAL EFFECTS OF OZONE
f
Ozone
concentration
ug/m3
196

392


196



392



392
686
980
1568
392
686




392
980
1568







ppm
0.1

0.2


0.1



0.2



0.2
0.35
tf. 5
0.8
0.2
0.35




0.2
..0.5
0.8








Measurement, '
method
UV,
NBKI



UV,
NBKI

,
MAST,
NBKI


UV,
NBKI

j
UV,
NBKI
i



UV, :
NBKI








Exposure
duration
and
protocol
7 'days,
continuous



7 days,
continuous


30 days,
continuous


7 days,
8 hr/day


7 days.,
8 hr/day




7 days,
8 hr/day
or
24 hr/day








Observed effect(s) Species
Two of six fed with "basal" vitamin E diet had increased cen- Rat
triacinar AMs (SEM, LM).
Centriacinar accumulation of AMs, commonly in clumps of
3-5. Occasionally cilia were reduced in number, nonciliated
cells, some reduction in height.
Five of six fed E-deficient "basal" diet had centriacinar AMs Rat
and bronehiolar epithelial lesions (SEM). Four of six fed "basal"
diet +11 ppm E had lesser but similar lesions. One of six fed
"basal" diet +110 ppm E had lesser lesions.
Increased lung volume, mean chord length, and alveolar Rat
surface area. Lung weight and alveolar number did not
change. Decrease in lung tissue elasticity. Parenchyma
appeared "normal" by LM,
Respiratory bronchi olitis at all concentrations. Increased Monkey
AMs. Bronchiolar -epithelium both hyperplastic and hypertrophic. (Rhesus
Increased alveolar type 2 cells. Random foci of short, blunt and
cilia or absence of cilia (LM, SEM, TEM). Bonnet)
All exposed monkeys had LM & EM lesions. Trachea and bronchi Monkey
had areas of shortened or less dense cilia. RBs had AM (Bonnet)
accumulation and cuboidal cell hyperplasia. Alveoli off RBs
had AM accumulations and increased type 2 cells. RB walls of
the 0.35-ppra group were often thickened due to mild edema and
cellular infiltration.
Exposed groups gained less weight. -Focal areas of missing or Rat
damaged cilia in trachea and bronchi. TB nonciliated (Clara)
cells were shorter arid had increased surface granularity and less
smooth endoplasmic reticulum. Ciliated cells of TB had fewer
cilia and focal blebs.. Centriacinus had clusters of AMs and
PMNs. Type 1 cells swollen and fragmented and type 2 cells fre-
quently in pairs or clusters. Proximal IAS were minimally thick-
ened. Lesions in 0.2 rats were, mild (LM, SEM, TEM). Only slight
differences between rats exposed continuously 24 hr/day compared
to those exposed only 8 hr/day.


Reference
Plopper et al., 1979




Chow et al., 1981



Bartlett et al., 1974



Dungworth et al . , 1975b



Castleman et al.,-1977





Schwartz et al., 1976










-------
                                                      TABLE 9-1.   MORPHOLOGICAL EFFECTS OF OZONE   (continued)
Ozone
concentration
figTi3 ppi
392
980
1568
W>
03 392
980
I960
392
1960
0,2
0.5
0.8
0,2
0.5
1.0
0,2
1.0
Exposure
k duration
Beasurement ' and
•ethod protocol
UV, 20, 50, or
NBKI 90 days;
8 hr/day
ND, ' 4 days,
NBKI 3 hr/day,
exercised in
a rotating
case alter-
nate 15 min
CHEM 7, 14, 30, 60,
90 days; con-
tinuous
10 days, con-
tinuous
Observed effect(s) Species Reference
Epithelial changes and PAH accumulations at 90 days were - Rat Boornan et al., 1980
similar to 7-day exposures, but less severe. 0,5- and 0.8-pptn
groups had increased centriacinar PAMs at all tines. 0.2 ppm
and controls could not be separated by "blind" LH examination,
nor were there distinguishing EH changes. 90-day 0,8-ppm group
had changed the terminal bronchiole/alveolar duct junction to
terminal bronchiole/respiratory bronchiole/alveolar duct junctions.
TBs had loss or shortened cilia. Nonciliated cells were flattened
lumenal surfaces that occasionally occurred in clusters. Proximal
alveoli of 20- and 90-day 0.8-ppn groups had thicker blood/air
barriers.
Exercised control mice have significantly smaller body weights. Mouse Fukase et al., 1978
Both unexercised and exercised mice exposed to 0.5 or 1.0 ppm (male,
had smaller body weights and larger lung weight. Exercised 5 weeks
mice exposed to 0,2 ppm also had larger lung weights. Other old,
pathology not studied. ICR-JCL)
Short-term exposures produced a slight degree of tonsil epithelial Rabbit Ikematsu, 1978
detachment. Cell infiltration below the epithelium was slight.
Long-term exposures caused slight edema of the lacunar epithelium
which was destroyed or detached in places. Lymphocyte infiltra-
tion also occurred.
Tonsil epithelium had a high degree of detachment. Cell satura-
tion occurred below the epithelium. Some protrusion of the tonsil
9800    5.0
                                      3 hr
into the oral cavity.

Strong detachment of the tonsil  epithelium.   High degree of cell
saturation below the epithelium, including lymphocyte infiltra-
tration around the blood vessels and swelling of the endothelial
cells.  Large amount of lymphocytes, viscous liquid, and detached
epithelial cells in the tonsilar cavity.

-------
TABLE 9-1.   MORPHOLOGICAL EFFECTS  OF  OZONE  (continued)
Ozone
concentration
Jig/m"
490
510
980
1960
588
ppm
0.25
0.26
0.50
1.0
0.3
Exposure
• . duration
Measurement ' and
method protocol
CHEM ; 6 weeks,
| 12 hr/day
i
i
HAST, 4.7-6.6 hr,
NBKI ; endotracheal
tube
NBKI : '16 days,
3 hr/day
Observed effect(s)c Species Reference
Centriacinar or proximal alveoli had thicker interalveolar septa Rat Barry e£ al., 1983;
with significant increases in epithelium, cellular interstitiutn, Crapo et al., 1984
and endothelium. Type 1 and 2 alveolar epithelial cells and macro-
phages were increased in numbers. Type 1 cells had smaller volumes
and surface areas and were thicker.
Desquamation of ciliated epithelium. Focal swelling or Cat Boatman et al., 1974
sloughing of type 1 cells.
SEN, but not LM, showed swollen cilia with hemispheric Rat Sato et al., 1976a
extrusions and surface roughness. Some adhesion of
  severely injured cilia otcurred.   Small,  round bodies were
  frequently noted, mainly in  the  large  airways  and proximal
  bronchioles.   Luminal  surfaces of  the  epithelium  were often
  covered with  a pseudomembrane.   The  surfaces of Clara cells
  showed swellings and round bodies.   The  surfaces  of alveolar
  ducts and walls showed scattered areas of cytoplasmic swelling
1
H

588



588


588



686
980
1372
1470
1960
686
980





0.3



0.3


0.3



0.35
0.50 •
0.70
0.75
1.00
0,35
0.50





NBKI ! 28 weeks,
: 5 days/week,
i 3 hr/day
i
UV ; 6 weeks,
1 5 days/week,
! 7 hr/day
ND 1, 5, 11, and
16 days,
r 3 hr/day

ND ; 1, 2, 4, 5, 6
or 8 days, con-
tinuous


i 4 days, contin-
'•. uous, followed
i by 0.50, 0.70,
; 0.75 or 1.00
: for 1-4 days
and attachment of round bodies. All responses were pronounced •
in vitamin E-deficient rats. Some rats had chronic respiratory
disease.
No morphological differences noted between vitamin E-defi- • Rat
cient and vitamin E-supplemented groups with the use of SEM
and TEM. Exposed and control rats had chronic respiratory
disease.
Increased LDH positive cells stated to be type 2 cells. Mouse


Rats were fed a basal diet with or without vitamin E supplement. Rat
Volume density of lamellar bodies in type 2 alveolar epithelial
cells were increased. Giant lamellar bodies were seen after 11
days exposure.
Dividing cells were labeled with tritiated thymidine and Rat
studied with autoradiographic techniques by using LM. All
labeled cells increased and then decreased to near control
levels within 4 days. Type 2 cells showed largest change
in labeling index.
Type 2 cells from groups showing adaptation to 0, were
exposed to higher concentrations. Groups exposed to low
initial concentration of Oa (0.35 ppn) did not maintain
tolerance. Groups exposed to higher initial concentration
(0.50 ppm) demonstrated tolerance.


Sato et al . , 1980



Sherwin et al. , 1983


Shimura et al. , 1984



Evans et al. , 1976b










-------
                                                         TABLE 3-1.  MORPHOLOGICAL EFFECTS OF OZOME  (continued)
to
o
Ozone
concentration Mtasureaent
pg/a3 ppi method
784 0.4 HAST
784 0.4 NBKI
980 0.5 NO
980 0.5 UV,
or or NBKI
1568 0.8
980 0.5 UV,
NBKI
Exposure
k duration
>b and
protocol
10 months,
5 days/week,
6 hr/day
7 hr/day,
5 days/week,
6 weeks
2 to 6 hr
7 continuous
days;
2, 4, 6, 8,
or E4 hr/
day
7, 21, and
35 days,
continuous
Observed effect(s)c
All (exposed and control) lungs showed some degree of inflam-
matory infiltrate possibly due to Intel-current disease.
A "moderate" degree of "emphysema" was present in 5 of the 6
exposed rabbits. Lungs of the 6th were so congested that
visualization of the mural frwework of the alveoli was
difficult. Small pulmonary arteries had thickened tunica iiedias,
sometimes due to edema, other times to muscular hyperplasia.
Lung growth which follows pneumonectomy also occurred
following both pneumonectomy and 03 exposure.
Centriacinar type 1 cells were swollen then sloughed.
Type 2 cells were not damaged and spread over the denuded
basement membrane. In some areas of severe type 1 cell
damage, endothelial swelling occurred. Damaged decreased
rapidly with distance from TB. Damage was most severe only
in the most central 2-3 alveoli. Interstitial edema occasionally
observed.
Centriacinar inflammatory cells (mostly AMs) were counted in
SEMs. Dose- related increase in inflammatory cell numbers except
in the continuously (24-hr/day) 0,8-ppm exposed rats. Rats
exposed 0.5 and 0.8 ppm 24 hr/day had the same intensity of
effect.
Most severe damage at terminal bronchiole/alveolar duct
junction. TB had focal hyperplastic nodules of non-
ciliated cells. Proximal alveoli had accumulations of
macrophages and thickening of IAS by mononuclear cells
at 7 days. At 35 days, changes much less evident, but
increased type 2 cells.
Species
Rabbit
(New
Zealand)
Rabbit
Rat
(young
males)
Rat
House
(Swiss-
Webster;
60 days
old; 35-40
Reference
P'an et al., 1972
Boatman et al., 1983
Stephens et al . , 1974b
Brutnmer et al., 1977
Zitnik et al., 1978
9)
    980     0.5
                            NO
                                          2  days
                                                Tolerance was induced by exposure to 0.5 ppm 03 for 2 days.
                                                Challenge was by exposure to 6.0 ppm Q3  for 24 hours.   Toler-
                                                ance was present at 3 days and declined  at 7 and 15 days after
                                                the initial  exposure.   When the animals  were tolerant,  the type 1
                                                alveolar epithelium was thicker, had a smaller surface area and a
                                                smaller surface-to-volume ratio.
                                                                                                                                 Rat
                                                                                                                    Evans  et al.,  1985
    980
    1568
0.5
0.8
UV,           7, 28,  or         Principal  lesion was a "low-grade  respiratory bronchiolitis"
NBKI          90 days,          characterized by "intraluminal  accumulations  of macrophages
              continuous,       and hypertrophy and hyperplasia of cuboidal bronchiolar
              8 hr/day          epithelial  cells."  Conducting  airway lesions not apparent
                                by LM, but parallel linear arrays  of uniform  shortening and
                                reduction  of density of cilia by SEN.   Kulschitzky-type cells
                                appeared more numerous in exposed.
Monkey        Eustis et al., 1981
(Bonnet)

-------
                                             TABLE 9-1.  MORPHOLOGICAL EFFECTS OF OZONE  (continued)
Ozone
concentration
fig/in3
980
1568
980
1960
f 980
& to
H 3920
ppm
0.5
0.8
0,5
1.0
0.5
to
2.0
Measurement3 '
method
UV,
NBKI
MAST
NBKI
UV,
NBKI
Exposure
duration
and
protocol
7 days ,
8 hr/day
60 days
6 hr/day
30 days
3 hr/day
7 days,
24 hr/day
Observed effect(s) , Species Reference
All exposed, monkeys had lesions. Lesions similar in 0.5- Monkey - Mellick et al., 1975,
and 0.8-, less severe in 0.5-ppm exposure groups. Patchy (Rhesus, 1977
areas of epithelium devoid of cilia in trachea and bronchi. adult)
Luminal surfaces of RB and proximal alveoli coated with
macrophages, a few neutrophils and eosinophils and debris.
Nonciliated cuboidal bronchiolar cells were larger, more
numerous, and sometimes stratified. Proximal alveolar
epitheliuui thickened by increased numbers of type 2 cells.
Progressive decrease in intensity of lesions from proximal to
distal orders of RBs..
Both immersion and infusion fixed lungs were studied by LH. Rats Yokoyama et al., 1984
Immersion fixed large and middle size bronchi had deeper than
normal infolding of the mucosa with increased secretions. The
low concentration rats had less severe mucosal infolding, but
a greater accumulation of secretions.
Elevated collagen synthesis rates and histologically Rat • Last et al., 1979
discernible fibrosis was present at all levels of 03.
O.i
to
1.5
14 days
and
21 days,
24 hr/day
0.5 ppm        Minimal  or no thickening of walls or evidence of
               fibrosis.   Increased number of cuboidal  cells and
               •acrophages present.

0.8-2.0 ppm:    Moderate-thickening of AD walls and associated
               IAS by fibroblasts, reticulin and collagen with
               narrowing of the ducts and alveoli.   Thickening
               decreased with increased length of exposure.

0.5 ppm        Sometimes minimal thickening of alveolar duct
               walls with mildly increased reticulin and
               collagen.
980    0.5

       0.5 03

       10 mg/m3
                CHEM,          6 months,          Only  03  caused  pulmonary  lesions.  Only  LM  histopathology,
                NBKI          5 days/week,       no SEM nor  TEM.   Rats  did not have exposure-related pulmonary
                              6 hr/day          lesions,  except 2 of 70 rats in  the 03 group, which had
                                                type  2 hyperplasia and focal alveolitis;  2  of 70  rats from the
                H2S04                            03 +  H2S04  group, had slight hypertrophy  and hyperplasia of
                                                bronchiolar epithelium.   Guinea  pigs exposed to 03 or 03 +
                                                H2S04 had lesions "near"  the TB.  Epithelium was  hypertrophied
                                                and hyperplastic.   Macrophages were in centriacinar alveoli.
                                                Occasionally proliferation of type 2 cells.  Trachea and bronchi
                                                had slight  loss of cilia, reduction of goblet cells, and mild
                                                basal cell  hyperplasia.   Ozone alone had no effect on body weight
                                                gain; lung/body weight ratio; RBCs, hemoglobin, or hematocrit.
                                                                                        Rat
                                                                                        and
                                                                                        Guinea
                                                                                        pig
                                                                                    Cavender et al.,  1978

-------
                                                         TABLE 9-1.  HORPHOLOGICAL EFFECTS OF OZONE  (continued)
f
to
to
Ozone
concentration
ug/raa pp"5
980
1058
1725
0.5
0.5 03
1 nig/13
0.54
0.88
Exposure
k duration
Measurement ' and
method protocol
UV, 3, 50, 90,
NBKI or 180 days;
continuous,
plus 62 days
. H2S04 postexposure,
24 hr/day
NO 2, 4, 8, 12,
or 48 hr
Observed effect(s)c Species * Reference
H2S04 did not potentiate effects of Og alone. Fixed lung Rat Moore and Schwartz, 1981
wolmes were increased at 180 days, but decreased at 62 days
postexposure/ After SO, 90, 94 180 days all 03 exposure
rats had "bronchiolization of alveoli" or formation of an RB
between the TB and ADs. Centriacinar inflammatory cells were
significantly increased at all exposure times and after 62 days
postexposure. TB lesions were qualitatively similar at 3, 50,
90, and 180 days. Cilia were irregular in number and length.
Nonciliated secretory (Clara) cells had flattened apical pro-
trusions and a blebbed granular surface. At 90, but not 180 days,
small clusters of nonciliated cells were present in the TB. At
180 days, 2 of 12 rats had larger nodular aggregates of noncili-
ated cells which bulged into the lumen. Most rats had a very nild
interstitial thickening of alveolar septa in the centriacinar
region (LH, SEM, TEH),
Severe loss of cilia from T8 after 2 hr. TB surface more Rat Stephens et al., 1974a
uniform in height than controls, Necrotic ciliate cells in
    1058    0.54
6 months,
24 hr/day
TB epithelium and free in lumen after 6-12 hr of a 0.88-ppm
exposure.  Ciliated cell necrosis continued until 24 hr, when
little evidence of further cell damage or loss was seen.  Only
mi nilgai loss of ciliated cells in 0.5-ppm rat group.  Non-
ciliated cells were "resistant" to injury from 03 and hyper-
trophic at 72 hr.  Damage to the first 2 or 3 alveoli after
0.54-ppm for 2 hr.  Type 1 cell "fraying" and vesiculation.
Damage was greater after 0.88 for 2 hr.   "Basement lamina"
denuded.  Type 2 and 3 cells resistant.   Hacrophages
accumulated in proximal alveoli.  Endothelium appeared
relatively normal.

Repair started at 20 hr.  Type 2 cells divide, cuboidal
epithelium lines proximal alveoli where type 1 cells were
destroyed.  Continued exposure resulted in thickened alveolar
walls and tissue surrounding TBs,  Exposure for 8-10 hr followed
by clean air until 48 hr resulted in a proliferative response
(at 48 hr) about equal to that observed after continuous exposure
(LM, SEH, TEH).

No mention in either the results or discussion of the 6 months
at 0.54-ppra group.

-------
TABLE 9-1.   MORPHOLOGICAL EFFECTS OF OZONE  (continued)
Ozone
concentration
ug/m3 ppm
1058 0.54
1725 0.88
1764 0.9 Oa
0.9 N02 -
490 0.25 03
2.5 N02
1176 0.6
vo 2548 1.3
1
NJ
US
1254 0.64
1254 0.64'
1882 0.96
5000 (NH4)2S04
1254 0.64
Exposure
, duration
Measurement ' and,
method protocol
ND 4 hr
to
3 weeks
60 days
6 months
NO 1 or 2 days,
6 hr/day or
7 hr/day
UV 1 year
8 hr/day
UV 3, 7 or 14 days
UV 23 hr/day
3 or 7 days
Observed effect(s)c
Ozone-exposed lungs heavier and larger than controls. Increased
centriacinar macrophages. Hyperplasia of distal airway epithelium
Increased connective tissue elements. Collagen-like strands
formed bridges across alveolar openings. Fibrosis more pronounced
in 0.88-ppm group.
Respiratory distress during first month. Several rats died.
Gross and microscopic appearance of advanced experimental
emphysema as produced by NOZ earlier (Freeman et al., 1972).
Ozone potentiated effect of nitrogen dioxide. .
"At 6 months the pulmonary tissue seemed quite normal." Proximal
orders of ADs minimally involved.
Endothelial cells showed the most disruption. The lining mem-
branes were fragmented. Cell debris was often present in the
alveoli as well as the capillaries. Some disorganization of
the cytoplasm of the large alveolar corner or wall cells was
evident.
LM and TEM morphometry revealed increased volume density and
volumes of RBs which had thicker walls and narrower lumens.
PeHbronchiolar and perivascular connective tissue was increased
by increased inflammatory cells and amorphous extracellular
matrix rather than stainable fibers. In RBs cuboidal bronchiolar
cells were increased and type 1 cells decreased. The media and
intima of small pulmonary arteries were thicker.
Ammonium sulfate aerosol enhanced the effects of 03 alone and
accelerated the occurrence of these effects. The same numbers of
lesions were seen in lungs from rats exposed to either 03 alone or
to 03 with the aerosol, but the lesions were larger in the latter
group. Lesions in the 03 plus aerosol rat lungs had more inflam-
matory cells, fibroblasts and stainable collagen fibers.
SEM and TEH, including TEM morphoraetry, revealed necrosis of
ciliated cells, decreased numbers of ciliated cells and loss of
Species
Rat
(month
old)
Rat
(month
old)
Rat
(month
old)
Mouse
(young^
Bonnet
monkey
Rat
Bonnet
monkey
Reference
Freeman et al . , 1974
Bils, 1970
Fujinaka et al. , 1985
Last et al . , 1984a
Wilson et al. , 1984
  cilia.   Extracellular space was  increased  and focal  areas of epi-
  thelial  stratification were seen.   Small mucous  granule cells
  were increased and an intermediate  cell was  described.   Regular
  mucous  cells had decreased density  and  smaller irregularly sized
  secretory granules which contained  only filamentous  or granular
  material.   TEH morphometry indicated  the most severe lesions
  occurred at 3 days of exposure and  that the  epithelium had returned
  towards normal after 7 days of exposure.

-------
TABLE 9-1.  HQRPHOLOGICflL EFFECTS OF OZOHE  (continued)
Ozone
concentration
jjgTi3"
1372

1568


1372



ppffl
0.7
to
0,8


0,7



Exposure
b duration
Measurement ' and
method
UV,
NBKI



NO



protocol
7 days,
continuous



24 hr,
continuous



Observed effect(s)c Species
|n situ cytochenrical studies of lungs from 03 exposed and Rat
control rats. Ozone-exposed rats had increased acid phosphatase,
both in lysosoties and in the cytoplasm, in nonciliated bronchiolar
(Clara) cells, alveolar macrophages , type 1 and Z cells, and
fibroblasts.
Exposure end: General depletion of cilia from TB surface. Rat
Nonciliated cells were shorter and contained
fewer dense granules, less SER, and lore free
ribosomes.

Reference
Castleiaan et al,, 1973a




Evans et al., 1976a



Post exposure: TB returned towards normal.
(0-4 days)
1568 0.8 UV, 6, 10, 20
NBKI days
exposure,
24 hr/day
or
20 days
exposure +
10 days
postexposure
1568 0.8 ' UV, 7 days,
NBKI continuous
with samples
at 6, 24,
72 and
168 hr.
1568 0.8 UV, 4, 8, 12,
NBKI 18, 26, 36,
50, and post
48 and
188 hr,
continuous
SEH of distal trachea and primary bronchi: Mouse Ibrahim et al,, 1980
6 days: Cilia of variable length. (Swiss
10 days: Marked loss of cilia. Very few cells had Webster)
normal cilia. Some nonciliated cells were
in clusters and had wrinkled corrugated
surfaces.
20 days: Similar to 10 days.
10 days Cilia nearly normal.
postexposure: Clusters of nonciliated cells were present and
elevated above the surface.
Clusters of nonciliated cells were interpreted as proliferative
changes.
Exposure-related epithelial changes. TB cell populations Rat Lura et al., 1978
changed after Os exposure; fewer ciliated and more non-
ciliated secretory cells.
Degeneration and necrosis of RB type 1 cells predominates Monkey Castlentan et al., 1980
from 4-12 hr. Type 1 cell most sensitive of RB epithelial (Rhesus)
cells. Labeling index highest at 50 hr. Mostly cuboidal
bronchiolar cells but some type 2 cells. Bronchiolar
epithelium hyperplastic after 50 hr exposure, which
persisted following 7 days postexposure. Intraluminal
•aerophages increased during exposure, but marked clusters
of K cells at 26-36 hr.

-------
TABLE 9-1.   MORPHOLOGICAL EFFECTS OF OZONE  (continued)
Ozone
concentration
(ig/nia ppm
1568 0.8










Exposure
. duration
Measurement ' and
method protocol
UV, 3 days,
NBKI continuous
0, 2, 6, 9,
16, and 30
days post-
" exposure;
2nd 3 days
continuous
after 6, 13,
and 27 days
postexposure


Observed effect(s)c
1st Exposure; TB epithelium flattened and covered with debris.
Ciliated cells either unrecognizable or had
.shortened cilia. The type of most epithelial
cells could not be determined. Proximal
alveoli had clumps of macrophages and cell
debris. Type 2 cells lined surfaces of many
proximal alveoli. Occasionally, denuded basal
lamina or type 1 cell swelling.

Postexposure:
6 days: Most obvious lesions were not present. TB epithelium


Species Reference
Rat Plopper et al., 1978
(Sprague-
Dawley;
70 days
old)






                 had usual pattern.   Clumps of macrophages had cleared
                 from the lumen.   Host proximal alveoli  lined by normal
                 type 1 and 2 cells.
NJ
UT
1666 0.8S ND
Similar
exposure
regimen
for 14 ppm
N02) but not
mixtures.
1960 1.0 NBKI
|
' i 1, 2, 3 .
I days ,
. continuous
"50 weeks,
~5 days/week,
6 hr/day
(268 expo-
posnres)
30 days: Lungs indistinguishable from controls.
2nd 3-Day 6 or 27 days after the end of the 1st. Lesions
Exposure: same as 1st exposure.
Birth to weaning at 20 days: "Very little indication of
response" or "tissue nodules" with dissecting microscope.
12 days: N02 lesions but no 03 lesions.
22 days old: 03, loss of cilia, hypertrophy of TB cells,
tendency towards flattening of luminal
epithelial surface..
32 days old: 03, loss of cilia, and significant hypertrophy
of TB epithelial cells.
21 days old Alveolar injury, including sloughing to type 1
and older: cells resulting in bare basal lamina.
Response plateau is reached at 35 days of age.
Chronic injury occurred in the lungs of each species of small
animal. The principal site of injury was In the terminal air-
way, as manifested by chronic bronchiolitis and bronchiolar
wall fibrosis resulting in tortuosity and stenosis of the
passages.

Rat Stephens et al . , 1978
(Sprague-
Dawl ey ;
1, 5, 10,
15, 20, 25,
30, 35, and
40 days old)
Mouse, Stokinger et al., 1957
Hamster,
Rat

-------
                                                     TABLE 9-1.  MORPHOLOGICAL EFFECTS OF OZONf  (continued)
Ozone
concentration
ug/«3 ppi
1960
to
5880
1960
1960
I960
3920
5880
1.0
to
3.0
1.0
1.0
1.0
2.0
3.0
Exposure
, duration
Measurement ' and
method protocol
NO 18 months,
8-24 hr/
day
A = 8 hr/day
8 = 16 hr/day
C = 24 hr/day
0 = 8 hr/day
; E = 8 hr/day
Observed effect(s)c Species Reference
Result: Dog Freeman et al., 1973
A 1 ppm, 8 hr/day: Minimal fibrosis occasionally
and randomly in the periphery of an alveolar duct.
A few "extra" macrophages in central alveoli.
B 1 ppm, 16 hr/day: Occasional fibrous strands
in some alveolar openings of RBs and ADs. A few more
"extra" macrophages.
                                                                  1 ppm, 24 hr/day:  More extensive fibrosis of
                                                                  centriacinus.  Thickened AD walls.  More "extra"
                                                                  macrophages.  Sporadic hyperplasia of epithelium of
                                                                  RB  and AD.
NJ
1960 1.0
and and
3920 2.0

Also
mixtures
with
H2S04




' D & More fibrosis. Epithelial hyperplasia and squamous
: E metaplasia.
CHEH, ! 2 or 7 days, Results: Rat Cavender et al., 1977
NBKI 6 hr/day 03 alone: Lesions limited to centriacinus. and
; 1 ppm: Hypertrophy and hyperplasia of TB epithelium. Guinea
[ Centriacinar alveoli had increased type 2 pig
i cells, increased macrophages, and thickened
i walls. Some edema in all animals.
; Lesions less severe at 7 days than at 2 days.
i This adaptation was more rapid in rats than
; guinea pigs.
2 ppm: Same plus loss of cilia in bronchi.
} 03 plus No additive or synergistic morphological
i H2S04 changes.
Measurement method:   MAST = KI-coulonetric (Mast meter);  CHEM =  gas  phase  chemiluminescence); NBKI = neutral buffered potassium iodide;
 UV = UV photometry;  NO = not described.
 Calibration method:   NBKI = neutral  buffered potassium  iodide.
 Abbreviations used:   LM = light (nicroscope;  EM = electron microscope;  SEN  =  scanning electron microscope; TEM = transmission electron microscope; RAM = pulmonary
 alveolar macrophage; RB = respiratory bronchiole; TB =  terminal  bronchiole;  AD =  alveolar duct; IAS = interalveolar septa; LDH = lactic dehydrogenase;

 SER = smooth endoplasnric reticulum;  RER = rough  endoplasmic  reticulum.

-------
     The lower, or  intrathoracic  conducting airways include the thoracic tra-
chea, bronchi, and  bronchioles.   Species variation of lower airway structure
is large, as  recorded  at the NIH workshop on comparative biology of the lung
(National Institutes of  Health,  1983).   The thoracic trachea and bronchi have
epithelial  and subepithelial tissues  similar to those of the upper conducting
airways.  In  bronchioles,  the epithelium  does  not contain mucous (goblet)
cells,  but in  their place  are specialized nonciliated bronchiolar epithelial
cells,  which  in  some  species  can appropriately  be called "Clara" cells.
Subepithelial tissues are sparse and do not contain glands.
     The ciliated  cell  is  the cell in the upper and lower conducting airways
in which morphological evidence  of damage  is most  readily  seen.  This cell is
primarily responsible  for  physical clearance or removal  of inhaled foreign
material from conducting airways of the respiratory system (see Section 9.3.4).
The  effects  of ozone on this cell type, which  is  distributed  throughout the
length  of conducting airways, are detected through various physiological tests
and  several  types' of morphological examination  (Kenoyer  et al.,  1981; Oomichi
and  Kita, 1974;  Phalen et al.,  1980;  Frager  et al.,  1979; Abraham et  al.,    ||
1980).
     The other portion of  the respiratory system directly damaged by inhala-
tion of 0, is the junction of the conducting airways with the gaseous exchange
area.   The structure and cell makeup of this junction varies with the species.
In  man, the most  distal conducting  airways,  the terminal  bronchioles, are
followed by  several generations of transitional  airways,  the respiratory
bronchioles, which  have  gas exchange areas as a part of their walls.  In most of
the  species  used for experimental exposures to 0,, (i.e., mouse, rat, guinea
pig,  and rabbit),  the terminal  bronchioles  are followed by alveolar ducts
rather  than  respiratory bronchioles.  The only common experimental animals
with  respiratory bronchioles are  the dog and  monkey,  and they have fewer
generations of nonrespiratory bronchioles  than  does man as well as differences
in  the  cells of the respiratory bronchioles  (National  Institutes  of  Health,
1983).
9.3.1.1.1  Airways
      9.3.1.1.1.1    Upper  airways (nasal cavity,  pharynx,  and  larynx).  The
effects of 03  on the upper extrathoracic airways have received  little attention.
The  effect  of upper airway scrubbing  on the level of 03  reaching  the  more
distal  conducting  airways  has been studied in rabbits  and guinea pigs (Miller
et  al., 1979).   They  demonstrated  removal  of  approximately 50 percent of
                                     9-27

-------
                                                3
ambient concentrations between 196 and 3920 (jg/m  (0.1 and 2.0 ppm).   Earlier,
Vokoyama and  Frank  (1972)  studied  nasal  uptake  in dogs.  They found uptake to
                                                                           3
vary with  flow  rate as well as with  03  concentration.   At 510 to 666 ug/m
(0.26 to 0.34 ppm) of 03, the uptake at low flows of 3.5 to 6.5 L/min was 71.7
±1.7 percent,  and  at  high flow rates of 35 to  45 L/min the uptake was 36.9 ±
                                   q
2.7 percent.  At  1529  to 1568 pg/m (.0.78 to 0.8 ppm), the uptakes at low and
high flows were 59.2 ±1.3 percent and 26.7 ±2.1 percent, respectively.  The
scrubbing effect of the oral cavity was significantly less at all concentrations
and flow  rates  studied.   Species  variations in uptake  by the nasal cavity
probably relate to  species differences in the complexity and surface areas of
the nasal  conchae and meatuses (Schreider and Raabe, 1981).
     No studies of  the effects  of 0~ on the nasal  cavity were found, but two
references to articles in  the  Japanese  literature were  cited  by Ikematsu
(1978).   At least one  study of the morphological effects of ambient levels of
0~ on the nasal  cavity of nonhuman primates is  in progress,  but not published.
                                            o
     The effects of 392, 1960, and 9800 pg/m  (0.2,  1.0, and 5.0 ppm) of Q3 on
the tonsils, the primary lymphoreticular structures of the upper airways, were
                                                               o
studied.   In  palatine  tonsils  from rabbits exposed to 392 ug/m  (0.2 ppm) of
0~ continuously for 1 and 2 weeks, Ikematsu (1978)  reported epithelial detach-
ment and  disarrangement and a slight  cellular infiltration.  The  significance
of these  observations  to the function of immune mechanisms in host defense is
unknown.
     9.3.1.1.1.2  Trachea.   Trachea!  epithelial lesions have been described
                                                             O
in several species  following exposure to less  than 1960 ug/m  (1 ppm) of Og.
Boatman et al.  (1974)  exposed anesthetized, paralyzed cats  to  510,  980, or
1960 ug/m3 (0.26, 0.50,  or 1 ppm) of 03 via an endotracheal tube for 4.7 to
6.6 hr.   This  exposure  technique bypassed the nasal  cavity,  resulting in
higher trachea!  concentrations than in usual exposures.  They reported desquama-
                                          3
tion of ciliated  epithelium-at 1960 ug/m  (1 ppm)  of 03, but none,at 510 or
980 ug/m3 (0.26 or 0.5 ppm).
     In rats  exposed  to  960 or 1568  ug/m  (0.5 or 0.8 ppm) of  03, 8 or 24
hr/day for 7  days, Schwartz et al. (1976) described focal areas of the trachea
in which  the  cilia  were reduced in density and were of variable diameter and
length.   Mucous cells appeared to  have been fixed in the process of discharging
mucigen droplets.   These  changes  were  more easily seen  with  the scanning
                                                                           2
electron microscope (SEM)  and were not  obvious in  rats exposed  to 392 ug/m
(0.2 ppm)  of  0- for the same times.   Cavender  et al.  (1977), when using only
              O          *
                                    9-28

-------
light microscopy  (LM), studied tracheas from rats and guinea pigs exposed for
7 days to  1960  or 3920 ug/m3 (1 or 2 ppm) of 03 or sulfuric add (HgSO^) or
both.  The article  does  not state the hours of exposure per day.   Trachea!
lesions,  which  consisted  of reduced numbers of cilia and goblet cells, were
reported only for guinea  pigs exposed to 03.   Animals exposed to both pollu-
tants had lesions similar to those exposed to 0, alone.
     By using SEM and transmission electron microscopy (TEM), Castleman et al.
(1977) described  shortened  and  less dense cilia in tracheas from bonnet mon-
keys exposed to 392 or 686 pg/m   (0.2 or 0.35 ppm),  8  hr/day  for  7 days.
Lesions occurred  as random  patches or longitudinal tracts.  In these  areas,
the  nonciliated  cells  appeared  to be more numerous.  The TEM study  revealed
that  cells with long cilia had the. most  severe cytoplasmic changes,  which
included dilated  endoplasmic  reticulum,  swollen mitochondria,  and condensed
nuclei, some of which  were  pyknotic.  In  lesion areas, evidence of ciliogene-
sis was seen in noncilated cells with a microvillar surface.  Mucous cells did
not appear to be  significantly altered,  but some had roughened apical surfaces.
                                                                o
The  changes were more  variable  and less  severe  in  the 392 ug/m  (0.2 ppm)
group.  More  extensive and  severe lesions of similar nature were  seen in
                                                         3
tracheas from rhesus monkeys  exposed to 980 or  1568 pg/m (0.5  or 0.8  ppm)  of
03  in  the  same'  exposure regimen (Dungworth et  al.,  1975b;  Mellick  et  al.s
1977).
     Wilson et  al.  (1984) evaluated the  response of  the tracheal epithelium
                                                                      3
from  bonnet monkeys exposed continuously for 3 or 7 days to 1254 ug/m  (0.64
ppm) Og using SEM,  quantitative TEM, and autoradiography.  Extracellular space
was  increased and foci of stratified epithelium were  reported.   Changes in
ciliated cells were generally similar, but more severe, than those reported by
Castleman  et al.  (1977).   These changes were more severe at 3 days  than at 7
days  when  both  the volume percentage and  population  density had returned to
control values.   They  also  reported fewer and  smaller  granules  in regular,  as
opposed to small-mucous-granule (SMG),  mucous cells.  These granules  also
lacked the biphasic appearance of  those seen in control monkeys.  At both time
periods SMG cells in exposed monkeys were more numerous and had greater numbers
of  granules than  controls.
                                           3
     After mice were exposed to 1568 yg/m   (0.8 ppm)  of 0, 24  hr/day  for 6,
10,  or 20  days  and  for 20 days followed by a 10-day postexposure period during
which  the  animals  breathed filtered air, the  suffice of .the  tracheas was
examined by  SEM by Ibrahim  et al.  (1980).   Short and  normal-length  cilia were
                                    9-29

-------
seen at 6  days,  but at 10 and 20 days, a marked loss of cilia and few normal
cilia were  seen.   Some of the nonciliated cells occurred as clusters and had
wrinkled or  corrugated surfaces.   After the mice  breathed  filtered air for
10 days, the  surface morphology of  the  cilia returned to  near  normal, but the
clusters of  nonciliated  cells were still present.   Earlier,  Penha and Wer-
thamer (1974) observed metaplasia of the trachea! epithelium from mice exposed
                                        o
to high concentrations of Q3 (4900 jjg/m  ,  2.5  ppm)  for 120 days.   After the
mice breathed clean air for  120 days,  the metaplasia disappeared, and the
epithelium had a  nearly normal frequency of ciliated and nonciliated cells.
     9.3.1.1.1.3   Bronchi.   Bronchial   lesions  were  studied  in many of the
same animals  as  those whose  tracheal  lesions  are described above and were
reported as  gene-rally  similar to the tracheal  lesions.  -At low concentrations
(Castleman et  a!.,  1977),  lesions tended  to be  more  severe in  the  trachea and
proximal bronchi  than  in  distal bronchi or  in  the  next segment of  the conduc-
tion airways,  the nonalveolarized bronchioles.   Eustis et al.  (1981) reported
lesions in  lobar,  segmental,  and  subsegmental  bronchi  from bonnet monkeys
exposed to  980 or 1568 |jg/m3 (0.5 or 0.8  ppm)  of 03  8 hr/day  for 7, 28, or 90
days.  Lesions at 7 days were similar to those previously described by Mellick
et al.  (1977)  and Castleman  et al.  (1977),  as  summarized  above.  At 28 and 90
days, lesions were  not readily apparent by  LM, but extensive damage to ciliated
cells was  seen using  SEM.  Uniform shortening  and reduced  density of cilia
were seen  in linear,  parallel arrays.   In  these areas, the numbers of cells
with a  flat surface covered  by microvilli  increased.  Wilson et al. (1984)
reported similar  changes  in  primary bronchi from  bonnet  monkeys  exposed to
         q
1254 pg/m  (0.64 ppm) 03 continuously for 3 or 7 days.
     Sato et al.  (1976a,b) studied bronchi  from vitamin E-deficient and control
                         q
rats exposed  to  588 jjg/m  (0.3 ppm) of 03  3 hr/day  for up to 16 days.   Using
LM, they did not  see bronchial lesions  with asymmetrical  swelling  and surface
roughness,  which  were obvious with SEM.   The  observations of Sato et al.
(1976a) support the concept  that lesions in conducting airways are best seen
with SEM and that LM tends to underestimate damage to these ciliated airways.
In these short-term studies,   lesions were more prominent  in vitamin E-deficient
rats.  This  is in contrast to later studies in  which Sato et  al. (1978, 1980)
                                                             2
exposed vitamin E-deficient and supplemented rats to 588 |jg/m  (0.3 ppm) of 03
3 hr/day,  5 days/week for 7  months following  which  they did  not  see clear
differences due to  vitamin E with SEM or TEM.
                                    9-30

-------
     Yokoyama et al. (1984) studied effects of 0~ on "middle-sized bronchi" of
rats fixed by  immersion  rather than infusion via the airways.  They compared
the effects of 3 hr/day  exposures  to 1960  jjg/m   (1.0 ppm)  0,  for  30 days with
                         3
those following 980 ug/m  (0.5 ppm) 03  6  hr/day for 60 days,  They reported
increased mucus and irregular  loss of  cilia,  especially on the projections of
mucosal folds typical  of this  type of fixation.   Changes were less severe in
the 0.5 ppm group.
     9.3.1.1.1.4   Bronchioles.   There are two  types  of  bronchioles with
similar basic structure:   those without alveoli  in their walls (i.e., nonalveo-
larized) and those with alveoli opening directly into their lumen (respiratory
bronchioles).   Man  and the larger experimental animals (e.g., nonhuman primates
and dogs)  have  both nonrespiratory and respiratory bronchioles, whereas most
of the  smaller  experimental  animals (e.g., mice, rats, and guinea pigs) have
only nonrespiratory bronchioles.  Because  the types, functions, and lesions of
epithelial cells are different, these two  types  of bronchioles will be discussed
separately.
     Nonrespiratory bronchioles  are conducting airways  lined  by two principal
types of  epithelial cells:   the ciliated  and nonciliated  bronchiolar cells.
The latter cell  is frequently called  the  Clara  cell.   Although man and most
animals have  several  generations  of nonrespiratory  bronchioles, some nonhuman
primates  have  only one (Cast!eman  et  al.,  1975).   The last-generation con-
ducting airway  before the gas  exchange area of the  lung is the  terminal bron-
chiole.   Terminal  bronchioles  may end by  forming respiratory bronchioles, as
in man, monkeys, and dogs, or by forming alveolar ducts, as in mice, rats, and
guinea  pigs.   The  acinus, the functional  unit of the  lung,  extends distally
from the  terminal  bronchiole and  includes  the gas  exchange  area  supplied  by
the terminal  bronchiole  and the vessels and  nerves that service  the terminal
bronchiole and its  exchange area.
     A  major  lesion due to 0~  exposure  occurs in the  central portion  of the
acinus, the centriacinar  region, which includes  the end of the terminal bronchi-
ole and the first'few  generations  of either respiratory bronchioles or alveolar
ducts,  depending  on the species.   The centriacinar region is the junction of
the conducting  airways with the gas exchange tissues.   The 0,, lesion  involves
both the  distal  portion of the  airway and the immediately adjacent  alveoli,
the proximal  alveoli.   In animals with respiratory  bronchioles,  the  lesion is
a respiratory bronchiolitis.   Regardless of species differences in structure,
the  lesion occurs  at the junction of the conducting  airways with the gas
exchange  area.
                                    9-31

-------
                                                                            o
     Terminal bronchiolar  lesions  in rats due to  inhalation of  < i960 ug/m
(< 1.0 ppm)  of  0~ for 2 hr  to  1 week have been  described by Stephens et  al.
(1974a), Evans et al. (1976a,c), Schwartz et al.  (1976), and others (Table 9-1}.
These changes were  recently  reviewed by  Evans  (1984).   Ciliated  cells are the
most damaged of the airway cells, and fewer of them are found in the bronchiolar
epithelium  of  exposed animals.   Those ciliated  cells present  tend to have
cilia with focal blebs and blunt ends.  Damaged ciliated cells are replaced by
nonciliated bronchiolar (Clara) cells (Evans et al., 1976a;  Lum et al.,  1978),
which become hyperplastic.   The typical projection of the nonciliated or Clara
cell into the lumen is reduced, and the luminal surface has increased granular-
ity.  The reduction in  projection height appears to be due to a reduction in
agranular (smooth) endoplasmic reticulum (Schwartz et al., 1976).  Many ciliated
cells contain basal bodies and precursors of basal bodies indicative of cilio-
genesis (Schwartz et al., 1976).  The few brush cells present in nonrespiratory
bronchioles  appeared  normal  (Schwartz et al., 1976).   The  lesions were more
severe in higher  generation, more distal nonrespiratory bronchioles than in
the lower generation, more proximal nonrespiratory bronchioles.
     In an  earlier  study,  Freeman et al. (1974)  exposed month-old rats con-
tinuously to 1058 or  1725  ug/m3 (0.54 or  0.88  ppm)  of Q3 for 4 hr to 3 weeks.
In addition  to  the  centriacinar accumulations of  macrophages  and the hyper-
plasia of the distal airway epithelium, they reported an increase in connective
tissue elements and collagen-like strands which formed bridges across alveolar
                           o
openings.    In the 1725-ug/m   (0.88-ppm) 0~ group,  the fibrosis was pronounced
and sometimes extended into terminal bronchioles.   In the same study,  Freeman
                                                               3
et al.  (1974)  exposed month-old rats to a mixture of 1764 ug/m  (0.9 ppm) 0,
             3                                                        3
and 1690 ug/m   (0.9 ppm) nitrogen  dioxide, or  to  a  mixture  of  490 ug/m  (0.25
                      q
ppm) 03 and 4700  ug/m  (2.5  ppm) nitrogen dioxide.  After 60 days of exposure
to the 0.9/0.9  mixture, Freeman et al. (1974)  reported  that "both grossly and
microscopically,  the  appearance of the lungs was  characteristic of advanced
experimental emphysema,"  referring  to  earlier nitrogen  dioxide exposures
(Freeman et  al.,  1972).   Freeman et  al.  (1974)  did not report emphysema in
rats exposed to 0^  alone,  only in those rats exposed to the 0.9/0.9 mixture.
The topic of emphysema is discussed later (Section 9.3,1.4.2).
     Results differ  in  four  studies  of long-term  (3-  to 6-month) exposures  of
                    o
rats to < 1960  ug/m  (<1.0 ppm) for  6 or  8 hr/day.  The differences appear to
be due at least in part to the methods used to evaluate the bronchioles.   When
                                    9-32

-------
                                                                       3
using only LM to study effects  in rats exposed for 6 months to 980 ug/m  (0.5
ppm) for 6 hr/day,  Cavender et al.  (1978) did not find exposure-related lesions.  .
Barr (1984) exposed rats to 1862 pg/m  (0.95 ppm) 03 8 hr/day for 90 days.; : He  '
examined the bronchioles with LM, TEM, and SEM.   Using SEM he reported loss of
apical projections of nonciliated cells and loss of both density and height of
cilia.  Boorman et  al.  (1980) and Moore  and Schwartz  (1981) reported  signi-
                                                                 3
ficant bronchiolar lesions following exposure to 980 or 1568 ug/m  (0.5 or 0.8
ppm) of 03 8 hr/day for 90 or 180 days. In both studies, loss or shortening of
cilia and  flattening  of the  luminal  projections  of  nonciliated bronchiolar
cells were observed in terminal bronchioles at each time period, including the
end of exposure at 90 or 180 days.   Clusters of four to six nonciliated bronchio-
lar cells, in contrast to dispersed individual  cells in controls, were seen at
90 days  in  both studies, but not at  180  days.  However,  in 2 of the 12 rats
exposed 180 days,  larger nodular aggregates of hyperplastic cells projected
into the  bronchial  lumen.  After 50,  90,  and 180  days  of  exposure, the nature
of the junction between the terminal bronchiole and the alveolar ducts changed
from the  sharp  demarcation seen in controls to a gradual transition with the
appearance of a respiratory bronchiole.  The presence of this change in distal
airway morphology was confirmed by Barr (1984).   Both ciliated and nonciliated
bronchiolar cells were  seen  on  thickened  tissue  ridges between  alveoli.  This
change could result from either alveolarization of the terminal  bronchiole or
bronchiolization of alveolar ducts.   Although this change in the airway mor-
phology persisted,  the  changed segment reduced in  length after  the 180-day-
exposed rats had breathed filtered air for 62 postexposure days.  The addition
of 1  to  10 mg/m  HgSQ*  to these concentrations  of 0., for the same exposure
times did  not  potentiate the lesions  seen  in  the CU-alone rats (Moore and
Schwartz, 1981; Cavender et al., 1978; Juhos et al., 1978).
     Ozone-induced  bronchiolar  lesions in mice are similar to  those seen  in
rats, But  the  hyperplasia of the nonciliated cells is more severe (Zitnik et
al., 1978); Ibrahim et al., 1980).  Following high concentrations (4900 ug/m ,
2.5 ppm), Penha and Werthamer (1974)  noted persistence (unchanged in frequency
or appearance) or micronodular  hyperplasia of noncilated  bronchiolar cells for
120 postexposure days following  120 days  of exposure.  At  lower  0~ levels (980
            3
or 1568 ug/m , 0.5 or 0.8 ppm),  the hyperplasia was pronounced (Zitnik et al.,
1978; Ibrahim et al., 1980).  Ibrahim et  al. (1980) noted  hyperplastic clusters
of  nonciliated cells 10 days  after  exposure  but did  not  make  observations
after longer postexposure periods.
                                    9-33

-------
      Guinea pigs were exposed by Cavender et al. (1977, 1978) continuously to
                                                                  •3
 1 or 2 ppm of Og for 2 or 7 days in acute studies and to 980 pg/m  (0.5 ppm)  6
 hr/day, 5 days/week for 6 months.  Morphological effects were studied only by
 LM.   The  acute,  higher  concentration distal-airway lesions were  similar  to
 those seen in rats  and  included loss of cilia and hyperplasia of nonciliated
 cells.   The authors reported  that  the long-term, lower concentration lesions
 were more severe in guinea pigs than those in rats exposed to a similar regimen.
 The  lesions were no more  severe in guinea pigs  exposed to a combination  of
 H2SCL aerosol  and Oj.
 9.3.1.1.2  Parenchyma.  Ozone does not affect the  parenchyma in a uniform
 manner.  The  centriacinar region (i.e.,  the junction of the conducting airways
 with the  gas exchange area)  is the focus of damage, and no changes have been
 reported in the peripheral  portions of the acinus.
      9.3.1.1.2.1  Respiratory bronchioles.   Respiratory bronchioles are the
 focus of effects, because they are  the junction of the conducting airways  with
 the  gas exchange area.   However, not all animals have respiratory bronchioles.
 They are  well  developed in man but  are absent or poorly  developed  in the
 common laboratory animals frequently used for On study, with the exception of
 dogs (Freeman et al., 1973)  and macaque monkeys (Mellick et al.,  1975, 1977;
 Dungworth et al., 1975b;  Castleman et al., 1977, 1980; Eustis et al., 1981).
                                                                3
 Short-term exposures of monkeys  to 392, 686, 980, or 1568 yg/m  (0.2, 0.35,
 0.5, or 0.8 ppm) of 0, 8 hr/day for 7 days resulted in damage to type 1 cells
 and  hyperplasia of nonciliated bronchiolar cells, which were visible by either
 light  or  electron  microscopy.   At  lower concentrations, these  lesions were
 limited to the proximal, lower generation respiratory bronchioles.  At higher
 concentrations,  the lesions extended deeper into the acinus.   The lesions  were
 focused at the  junction  of the conducting airways with the gas exchange area
 and  extended  from that Junction with increasing 0, concentration.
                                                        3
      The pathogenesis of the  lesions  due to 1568 ug/m   (0.8 ppm)  of Og for
 periods up to 50 hr of exposure was studied quantitatively by Castleman et al.
 (1980).  Damage to type  1  cells was very severe following 4, 8j and 12 hr of
 exposure.   Type 1 cell necrosis, which  resulted  in bare basal larninas reached
 a maximum at  12 hr.   The absolute and relative numbers of these cells decreased
 throughout the exposure.   Only a few type 2 cells had mild degenerative changes
* and  only  at  4 or 12 hr  of exposure.   Cuboidal  bronchiolar  cells  had mild
 degenerative  changes, swollen  mitochondria,  and endoplasmic reticulum at  all
 times except 18 hr.  Both cuboidal  bronchiolar and type 2 cells functioned as
                                     9-34

-------
stem cells in renewal epithelium, and both contributed to the hyperplasia seen
at the latter  exposure  times.   The inflammatory exudate included both fibrin
and a variety  of  leukocytes in the early phases.  In the latter phase's, the
inflammatory cells were almost entirely macrophages.  Inflammatory cells were
also seen  in the  walls  of  respiratory bronchioles and  alveoli  opening into
them.   These lesions were not completely resolved after 7 days of filtered air
breathing.
                                        3
     Monkeys exposed to 960 or 1568 ug/m  (0.5  or 0.8 ppm) of 0^ 8 hr/day  for
90 days had a low-grade respiratory bronchiolitis characterized by hypertrophy
and hyperplasia of cuboidal bronchiolar cells and intraluminal accumulation of
macrophages (Eustis  et a!.,  1981).  After the 90-day exposure, the percentage
of  cuboidal  respiratory bronchiolar epithelial  cells was  90 percent rather
than the 60 percent found in controls.   Intraluminal  cells, mostly macrophages,
reached a  maximum of a thirty-seven-fold increase after 7 days of exposure.
Their numbers  decreased  with continued exposure, but at 90 days of exposure,
they were  still  sevenfold higher than those of controls.  This study did not
include a postexposure period.
                                                                            3
     Fujinaka  et  al.  (1985) exposed adult male  bonnet  monkeys to 1254  jjg/m
(0,64 ppm)  03  8 hr/day  for one year.  They  reported  significantly increased
volume  of respiratory  bronchioles  which had  smaller  internal diameters.
Respiratory  bronchiolar  walls were thickened by epithelial  hyperplasia and
increased  peribronchiolar  connective tissue.  Several small  nodules  of  hyper-
plastic and  hypertrophied cuboidal   bronchiolar cells were reported near the
openings  of  alveoli  into the respiratory bronchiole.   The authors interpret
the morphometry of  respiratory  bronchioles  as  indicating  an extension of
bronchiolar epithelium into airways Which were  formerly  alveolar ducts  similar
to  the  formation  of respiratory bronchioles reported  in 03-exposed  rats by
Boorman et al.  (1980) and Moore and Schwartz (1981).
     In an earlier study,  Freeman  et al.  (1973) exposed female  beagle dogs to
          3
1960 |jg/m  (1  ppm) of 0, 8, 16, or 24 hr/day for 18 months.  Dogs exposed  to
          3
1960 ug/m  (1  PPm) °f 03  f°r  8  hr/day had the  mildest lesions, which  were
obvious only  in the terminal airway and  immediately  adjacent alveoli,  where
minimal fibrosis  and a  few extra macrophages were seen.  More fibrous strands
and macrophages were seen in centriacinar areas of lungs from dogs exposed 16
hr/day.  Lungs  from dogs exposed 24  hr/day  had  terminal airways  distorted by
fibrosis  and thickened by  both fibrous tissue and a mononuclear cell  infiltrate.
Relatively broad  bands  of  connective tissue were reported in distal airways
                                    9-35

-------
and proximal  alveoli.   Epithelial  hyperplasia was seen  sporadically  in the
bronchiolar-ductal zone.  Other  dogs  in that study exposed  to  3920 or 5880
fjg/m  (2 or 3 ppm) of 03 8 hr/day for the same period had more severe fibrosis,
greater accumulations of  intraluminal macrophages, and areas of both squamous
and mucous metaplasia of bronchiolar epithelia.
     9.3.1.1.2.2  Alveolar ducts and alveoli.   Alveoli  in the  centriacinar
region, but  not  those  at the periphery of the acinus, are damaged by ambient
levels of 03  (Stephens  et a!., 1974b; Schwartz et a!.,  1976; Mellick et a!.,
1977;  Crapo  et a!.,  1984).  The  lesion  is characterized by the destruction of
type 1 alveolar epithelial cells exposing the basal lamina;  an accumulation of
inflammatory  cells,  especially macrophages; hyperplasia of  type  2 alveolar
epithelial cells  that recover  the denuded basal lamina; and  thickening  of the
interalveolar septa.   In animals with  respiratory bronchioles (e.g.,  dog,
monkey) the  alveoli  involved at  low concentrations are those opening directly
into the respiratory bronchiolar lumen of low-generation respiratory bronchioles
(Dungworth et al,,  1975b).   As the concentration  is  increased,  the lesions
include alveoli  attached  to  but not,  seemingly, opening into the respiratory
bronchioles and extending distally to higher-generation respiratory bronchioles
(Mellick et  al.,  1977).   In  monkeys exposed to 1568  ug/m  (0.8 ppm) of 03,
alveoli opening  into alveolar  ducts are minimally involved  (Mellick et al.,
1977),  In animals that lack respiratory bronchioles (e.g.,  rat, mouse, guinea
pig) the alveoli involved open into or are immediately adjacent to the alveolar
ducts formed by the termination of the terminal  bronchiole.
     In the centriacinar region of animals which lack respiratory bronchioles,
damage to type 1 cells has been reported as early as 2 hours following exposure
           o
to 980 ug/m   (0.5 ppm)  03 using  LM  (Stephens et al.,  1974a).  While not an 0-
concentration studied comprehensively or illustrated in that article, the same
authors comment  in  the  results section  of their report that  TEM evaluation of
rats exposed  to  392 ug/m  (0.2 ppm) 0,, for 2 hours revealed "...considerable
damage and  loss  of  type  1  cells from  proximal alveoli..."   Recovering of
denuded basal lamina by type 2 cells  has been reported to start as  early as  4
hr (Stephens et al,, 1974b).   The type 2 cell labeling index following tritiated
thymidine reached a maximum at 2 days of continuous exposure to either  686 or
980 ug/m3 (0.35 or 0.5 ppm) of 0,, (Evans et al., 1976b).   Although the labeling
                                «5
index decreases  as  the  exposure continues (Evans et al., 1976b), clusters of
type 2 cells and cells intermediate between types 2 and 1 were reported following
90 days of exposure to 1568 ug/m   (0.8 ppm) by Boorman  et al.  (1980).  They
                                    9-36

-------
interpreted these intermediate cells as due to delay or arrest of the transfor-
mation from type 2 to type 1 epithelial cells.  Type 1 cell damage and occasional
sloughing were observed by Barry et al.  (1983)  in newborn  rats exposed to 490
    3
ug/m  (0.25 ppm) of 0- 12 hr/day for 6 weeks.   By using LM and TEM morphometric
techniques, these authors  also  found  that centriacinar alveoli also had more
type 1 and 2 epithelial cells and macrophages.  The type 1 cells were smaller
in volume, covered less surface, and were thicker.   Evans et al.  (1985) studied
this effect and  suggested that  03 tolerance exists when the surface  area of a
cell is  small  enough so that antioxidant mechanisms contained in that volume
can protect it from  damage.  Sherwin  et  al. (1983) found increased numbers of
lactate  dehydrogenase  (LDH)-positive  cells,  presumed  to be  type 2 alveolar
epithelial cells, by automated  LM morphometry of lungs from  mice exposed to
         3
588 ug/m  (0.3 ppm)  of 0- for  6 weeks.   Moore  and  Schwartz  (1981) reported
nonciliated bronchiolar cells lined some alveoli opening into the transformed
airways  located between terminal bronchioles and alveolar ducts of rats exposed
            3
to  1568  ug/m   (0.8  ppm) of 0-  8  hr/day  for 180 days.  Sulfuric  acid aerosol
did not  potentiate this lesion.
     The  inflammatory  cell response appears to  occur  immediately following or
concurrent with  the  type  1 cell damage  and  has been reported in monkeys as
early as after 4 hr  of 1568 |jg/m  (0.8 ppm) of  03 exposure (Castleman et al.s
1980).   In  rats,  the numbers of inflammatory cells per centriacinar alveolus
                                                                             3
appear to be related to 0- concentration, at  levels between 392 and 1568 ug/m
(0.2 and 0.8  ppm),  during 7-day  (Brummer et  al., 1977) and 90-day exposures
(Boorman et al. ,  1980).   Using the same  technique, Moore and Schwartz (1981)
found statistically  significant increases after 3, 50, 90,  and  180 days of
                              3
exposure 8 hr/day to 980  ug/m  (0.5  ppm) of 0~ and after 62 days of filtered
                                                           3
air following 180 days of exposure.    The  addition of 1 mg/m  H^SO* aerosol  did
not result in larger increases.   In monkeys,  the intensity of the response was
                                                                      3
greater  than  in  rats exposed to the  same 0~  concentration (1568 ug/m , 0.8
ppm)  in  the  same regimen (8 hr/day)  for 7,  28, or 90 days  (Eustis  et  al. ,
1981).   Although in  both species the  numbers  of inflammatory  cells per alveolus
decreased with increasing length of exposure, the decrease was not as rapid in
the monkey as in the rat (Eustis et al.,  1981).
     Several  investigators  (Boorman et al.,  1980; Moore and  Schwartz,  1981;
Fujinaka et al., 1985; Barr, 1984) have  presented evidence of bronchiolization
of  airways  which were previously alveolar  ducts;  i.e.,  bronchial  epithelium
replaces the  type 1  and  2 alveolar  epithelium typical  of alveolar ducts.
                                    9-37

-------
This phenomenon  is  most easily seen  and  quantitated  in  species which  normally
do not have respiratory bronchioles or have one very short generation of them.
In these  species, airways  with the characteristics of respiratory  bronchioles
are seen  between the terminal bronchioles and  the alveolar  ducts  of  exposed
but not control  animals (Boorman et a!., 1980; Moore and Schwartz,  1981; Barr,
1984).    In  species  which  normally  have  several generations of  respiratory
bronchioles, bronchiolization  is detected morphometrically by increases in the
volume fraction  or total volume of respiratory bronchioles.  Using morphometric
techniques, bronchiolization was reported in nonhuman primates exposed to 1254
    3
ug/m  (0.64 ppm) CU 8 hr/day for one year (Fujinaka et a!., 1985).
     The  interalveolar  septa  of centriacinar alveoli are thickened following
exposure  to ambient concentrations  of 03.  After 7  days of  continuous  expo-
sure, the thickening was  attributed to  eosinophilic  hyaline  material  and
mononuclear cells (Schwartz  et al.,  1976).  Loose arrangements  of cells and
extracellular materials suggested separation by edema fluid.   Castleman et al.
(1980) also reported  edema of interalveolar septa of monkeys exposed to 1568
    3
ug/m  (0.8 ppm)  of 0, for 4 to 50 hr.  Boorman et al. (1980) used morphometric
techniques on electron micrographs to quantitatively evaluate the thickness of
centriacinar interaveolar  septa.  The arithmetic mean thickness  was increased
                            3
in rats exposed  to 1568 ug/m   (0.8 ppm) of 03 8 hr/day for 20 or 90 days.   The
increased total  thickness  was due to thicker interstitium.  Although several
components could contribute  to this increased  thickness,  the subjective im-
pression  was one of a mild interstitial  fibrosis.  Crapo et  al.  (1984)  made a
more comprehensive morphometric study of centriacinar interalveolar septa from
                                     o
young adult rats exposed to  490 yg/m  (0.25 ppm) 03  12 hr/day  for 6 weeks.
They reported significant  increases in tissue thickness and suggested that the
increased thickness was due to significant increases  in all  cell types  except
type 2 cells, and to increased interstitium.
                                                               2
     Moore and Schwartz (1981), after exposing  rats  to  980 ug/m  (0.5 ppm) of
0- 24 hr/day  for 180 days,  reported very mild interstitial  thickening of
centriacinar interalveolar septa, which  they concluded was  due  to collagen.
Earlier,  Freeman et al.  (1973) morphologically demonstrated fibrosis  in beagle
                                  3
dogs exposed to  1960 to 5880  \sg/m   (1 to 3 ppm) of  03  8 to  24 hr/day for 18
months.
     In several  biochemical studies (see Section 9.3.3.6), Last and colleagues
have shown that  03  is collagenic.   In one of these  (Last  et al.,  1979), the

                                    9-38

-------
biochemical  observations  were correlated with  histological  observations of
slides stained  for  collagen  and  reticulin.   Elevated  collagen  synthesis  rates
were found  at all concentrations and  times  studied.   Mildly  increased  amounts
of collagen were seen morphologically in centriacinar alveolar duct septa from
                              3
most rats  exposed to 980 ug/m   (0.5  ppm) of 0~  24  hr/day  for  14  or 21 days.
                                                                   3
More severe  lesions were seen in rats  exposed  to 1568  to  3920 ug/m  (0.8 to
2.0 ppm) of  03  24 hr/day for 7,  14,  or 21  days.   Last et al.  (1983, 1984a)
have reported synergistic increases in morphometrically determined volume of
centriacinar lesions,  inflammatory cells,  and  "stainable"  collagen  in rats
exposed to  a mixture of 0,,  and  ammonium sulfate aerosols.  The 1984 report
                                                            3
concerns this synergism in rats  exposed to  1254 or 1882 ug/m   (0.64 or 0.96 ppm)
             3
03 and  5 mg/m   ammonium sulfate  for 3, 7,  or 14 days.  The results correlate
well with  biochemically  determined apparent collagen  synthesis  rates (see
Section 9.3.3.6).
     Two studies address the biologically important question of the morphological
effects that follow multiple sequential exposures to 0~ with  several days of
clean air interspersed between 0~ exposures  (i.e., a multiple  episode exposure
                                                         3
regime).  Plopper et al.  (1978)  exposed rats to 1568 ug/m  (0.8 ppm) 03 continu-
ously  for  3 days,   held  them in  the  chambers breathing filtered air until
                                                                           o
postexposure day 6, 13, or  27,  when  they were  again exposed  to  1568  |jg/m
(0.8 ppm) of 03 continuously for 3  days.  Rats were also examined 2, 6, 9, 16,
and 30  days after   the first 0«  exposure.   Lungs from rats breathing filtered
air for  9  days  after one 3-day  exposure  had only minimal lesions and after
30 days of  filtered air were indistinguishable from controls.  When the second
3-day 03  exposure  started 6 or  27  days after the end of the first exposure,
the lesions appeared identical  to  each other and to those seen at the end of
                                                                              3
the first exposure.  Barr (1984) compared lesions in  rats exposed to 1862 |jg/m
(0.95 ppm)  03  8 hours  every day  for  90 days with both control rats  and  with
rats  exposed to the same concentration 8  hours per  day in 5-day episodes
followed by 9  days  of filtered  air repeated 7 times  during an 89-day period.
The lesions were similar but  less  severe in the episodically exposed rats.
9.3.1.1.3   Vasculature,  blood, and  lymphatics.  Although edema is the  apparent
cause  of death due  to  inhalation of  high concentrations of Oa, there  is very
                                                                             3
little morphological evidence  of pulmonary  vascular damage due to £ 1960 ug/m
(^ 1.0 ppm)  of  0, exposure.  Bils (1970)  reported capillary endothelial damage
                                                               3
in mice  less than I month of age exposed  for 7  hr to  1960 jjg/m  (1 ppm) of 03,
but this experiment has  not been confirmed by  others.   Boatman et al. (1974)
                                    9-39

-------
did demonstrate endothelial damage in cats exposed via an endotracheal tube to
510, 980, or 1960 ug/m3 (0.26, 0.5, or 1.0 ppm) of Og for 4 to 6 hr, but it is
not clear  whether all exposure  levels  resulted in endothelial damage.  In
later studies  that  used  pneumonectomized and control rabbits, Boatman et al.
(1983) reported occasional  swelling  of capillary endotheliurn  in both  groups
                    3
exposed  to  784 ug/m   (0.4 ppm)  of 03  7 hr/day,  5 days/week for 6 weeks.
Stephens et al. (1974b) reported occasional  areas of endothelial swelling but
concluded "the endothelium remains intact and rarely shows signs of significant
injury."  Stephens  et  al.  (1974a)  reported that "endothelium  retained  a rela-
                                                               3
tively normal  appearance"  in  rats  exposed to  980  or 1764 ug/m  (0.5 or 0.9
                                     3
ppm) of  Oo for 2  to 12 hr  or  980 ug/m   (0.5  ppm) for up to 6  months.   In rats
                                                3
exposed  by the usual methods  to  980 or  1568  ug/m  (0.5 or 0.8 ppm) of  03 8 or
24 hr/day, centriacinar  interalveolar  septa  had a loose arrangement of cells
and extracellular material, indicating  separation by edema fluid (Schwartz et
al.» 1976).  These investigators did not find morphological evidence of damage
to endothelial cells.  Evidence  of intramural  edema  in centriacinar areas was
                                                                 3
found by Castleman  et  al.  (1980)  in monkeys exposed to 1568  ug/m  (0.8 ppm)
for 4 to 50  hr,  but they  did not  report morphological  evidence of vascular
endothelial damage.
     Arterial  lesions  have been only  rarely reported.   P'an et al.  (1972)
reported increased  thickness  of  the  media and  intima in  pulmonary arteries
                                 3
from  rabbits  exposed  to 784  ug/m   (0.4 ppm)  03 6  hr/day,  5 days/week for
10 months.   These rabbits  had evidence  of intercurrent disease  which was more
severe  in  exposed  animals.   The LM description indicates  "some  degree of
inflammatory infiltrate" in all  lungs,  and  in one exposed rabbit^the lesions
were so  severe that "visualization of the mural framework of the alveoli  was
difficult."  The  pulmonary artery media and intima  were  also significantly
                                                 3
thickened in bonnet monkeys  exposed  to 1254 yg/m  (0.64 ppm) 03 8 hr/day for
one year (Fujinaka et al.,  1985).
     No  references  to  morphological  damage   of  lymphatic vessels were  found.
                                                                           3
This is  not  surprising,  because following nasal inhalation  of £1960 ug/m
(<1 ppm) of 03, blood  capillary  endothelial  damage has not been reported, and
edema has  been reported only  in  centriacinar structures.   In  the more  genera-
lized edema  that follows  exposure to  higher concentrations, Scheel et al.
(1959) reported perivascular  lymphatics were greatly distended.
                                    9-40

-------
9.3.1.2   Sequence In which Sites are Affected as a Function of Concentration
and Duration of Exposure.  The sequence  in  which anatomic sites are affected
appears to  be  a function of concentration  rather  than of exposure duration.
At sites that are involved by a specific concentration, however, the stages in
pathogenesis of the lesion  relate  to the  duration  of exposure.   Multiple
anatomical  sites  in the  conducting and exchange  areas  of  the respiratory
system have  been  studied at sequential, time periods  in  only a few studies.
Stephens et al. (1974a), in a study of the short-term effects of several concen-
trations  of 0~, reported finding by  LM  significant damage to  centriacinar
                                                            3
type 1 alveolar epithelial cells in rats exposed to 980 ug/m  (0.5 ppm) 03 for
2 hours.  Using TEM,  they also reported, but  did  not document by figures,
                                                                          3
minimal damage  to centriacinar  type  1 cells  in rats  exposed to 392 ug/m
(0.2 ppm) 0- for  2 hours.   Boatman et al.  (1974)  found  lesions in both the
                                                                            3
conducting  airways  and  parenchyma of cats exposed to 510, 980, or 1960 |jg/m
(0.26, 0.5, or 1.0 ppm) via an endotracheal  tube for times as short as 4.7 and
6.6 hr.  Thus, if there is a time sequence in effect at various sites, it is a
short time.
     Increasing  concentration  not only results  in more  severe lesions, but
also appears to extend the lesions to higher generations of the same type of
respiratory  structure  (i.e., deeper into  the  lung)  (Dungworth et al.,  1975b).
Several investigators  who have described gradients  of lesions have related
them  to assumed decreases in concentration of 0-  as it progresses  through
increasing  generations  of airways and to  differences  in  protection and sensi-
tivity of cells at various anatomic sites.  For the conducting airways, Mellick
et al.  (1977)  reported more severe and  extensive  lesions in  the trachea  and
major  bronchi  than in  small  bronchi or terminal  bronchioles  of rhesus  monkeys
                            3
exposed to  980 or 1568  ug/m (0.5 or  0.8  ppm)  of 03  8 hr/day  for 7 days.   For
the acinus, they noted the most severe damage  in proximal respiratory bronchioles
and their alveoli  rather than in more distal,  higher generation ones.  Proximal
portions  of alveolar ducts  were only minimally involved.  The predominant
lesion  was  at  the junction of the  conducting airways with the exchange area.
In monkeys, as in  man,  the proximal  respiratory bronchioles,  not alveolar
                        *
ducts, are  in the  central portion of  the acinus.  Similar gradients of effects
in the conducting  airways and the centriacinar region were reported by Castleman
et al.  (1977),  who studied bonnet  monkeys  exposed to 392 and 686 ug/m  (0.2
and 0.35  ppm)  of 03 8 hr/day for 7 days.   In the  centriacinar region, this

                                    9-41

-------
gradient is most  easily demonstrated by the hyperplasia of cuboidal cells in
respiratory bronchioles  that  extended further distally in monkeys exposed to
686 rather than 392 ug/m  (0.35 than 0.2 ppm) of Og.
     The effects  of duration  of exposure are more complex.  In the  time frame
of a few hours, an early damage phase  has been observed at 2 or 4 hr of expo-
sure (Stephens et a!., 19,74a,b; Castleman et al., 1980).   Repair of the damage,
as indicated by DNA synthesis by repair cells, occurs as early as 18 hr (Castle-
man et al., 1980) or  24 hr  (Evans et al., 1976b; Lum et al., 1978).  Stephens
et al.  (1974a) reported  little change in the extent of damage after 8 to 10 hr
of exposure.   Full morphological development of  the lesion occurs at about   3
days of continuous exposure (Castleman  et al., 1980).  Damage continues while
repair is  in  progress,  but at a lower rate.  This phenomenon has been termed
adaptation (Dungworth et al., 1975b).  When the time  frame  is  shifted from
hours to days, severity  of the lesion at 7 days differs little between exposures
of 8 hr/day and  24  hr/day  (Schwartz et al., 1976).  When the time frame is
again shifted from days  to months of daily exposures, the centriacinar lesions
diminish in magnitude, but a significant lesion remains (Boorman et al., 1980;
Moore and Schwartz, 1981; Eustis et al., 1981).
9.3.1.3.   Structural Elements Affected
9.3.1.3.1   Extent ofinjurytoindividualcell types.   The extent  of  injury
to an  individual  cell is related to  the  product of the sensitivity of that
cell type  and the dose  of 03 delivered to the specific site occupied by that
cell.  Other  factors,  e.g.  maturity of the  individual cell, may  also  be  in-
volved.  The  dose to  an individual  cell  is  determined  by  the concentration  of
0« at that specific site in the respiratory system and the surface area of the
cell exposed  to  that  concentration of  03.   Thus,  sensitivity  and extent of
morphologically detected injury are not the same.   While  0- concentrations  at
any specific  site in  the respiratory  system  can not be determined  using  the
usual analytical  methods.,  the.,, concentration can _be estimated us ing., mode ling
techniques  (See  Section 9.2).   The  literature  does  contain extensive  informa-
tion on the extent  of morphologically detected injury to individual cells at
specific sites in the respiratory system.  That information is reviewed in the
following  paragraphs.   Ciliated cells  of the trachea and proximal,   lower
generation  bronchi are  subject to more damage than those located in  distal,
higher generation bronchi  or in lower generation bronchioles proximal to the
terminal bronchiole (Schwartz et al., 1976; Mellick et al., 1977).  Ciliated
                                    9-42

-------
cells in  terminal  bronchioles  of  animals without  respiratory bronchioles
(i.e., rats) are severely  damaged  by even low concentrations of 0, (Stephens
et a!., 1974a;  Schwartz  et a!., 1976), whereas those in terminal bronchioles
of animals with respiratory bronchioles (i.e., monkeys) are much less subject
to damage (Castleman et al.,  1977;  Mellick et a!., 1977).
     In a  similar  manner,  type 1 alveolar  epithelial  cells  located in the
centriacinar region are  subject to  damage by  low concentrations  of Qg, but
those in the peripheral portions of the acinus appear undamaged by the same or
higher concentrations (Stephens et al., 1974a,b;  Schwartz et al., 1976;  Castleman
et al., 1980; Crapo et al., 1984).
     Although type 2  alveolar epithelial  cells appear to be damaged less  by
Oo, some type 2 cells in centriacinar locations develop mild lesions detectable
with the TEM (Castleman  et al., 1980).   There is one report of larger than
normal lamellar bodies in type 2 cells from rats  fed a special  basal diet with
                                                             3
or without vitamin E  supplementation and exposed to 588 ug/m  (0.3 ppm) 03 3
hr/day for 11 or'16 days (Shimura et al., 1984).  Type 2 cells are  progenitor
cells that recover  basal lamina denuded by  necrosis  or  sloughing of type 1
cells and  transform  (differentiate)  into type 1 cells when repairing  the
centriacinar Oq  lesion  (Evans  et al.,  1976b).  Increased DMA synthesis by
type 2 cells, as evaluated by autoradiography, may be a very sensitive indica-
tor of 03 damage.
     Although  nonciliated  cuboidal   bronchiolar  cells appear  less  damaged
morphologically by 0~ than ciliated cells and are  the progenitor cells for
replacement  of  damaged ciliated cells  (Evans  et  al.,  1976a,c; Lum et  al.,
1978), they  are a  sensitive  indicator  of 0~ damage  (Schwartz  et al,, 1976).
                                3
Following  exposure to 1960 ug/m (<  1 ppm)  of 0~, their height is  reduced  and
their luminal surface is more granular  (Schwartz et al., 1976).  The reduction
in height appears to be  due to  a loss of smooth endoplasmic reticulum (Schwartz
et a-1., 1976).   		-	-.-.•.   			     	_		
     Several investigators report  that type 3 alveolar epithelial cells, the  ,
brush cells, are not damaged  by less than 1 ppm of 0, (Stephens et al., 1974a;
Schwartz et  al., 1976).  No reports  are available of damage to type 3 cells by
higher concentrations.
     Vascular endothelial  cells in  capillaries of the  interalveolar septa  may
be damaged much less than earlier reports indicated, because lesions are  not
described  in detailed studies using  TEM (Stephens et al., 1974a,b; Schwartz et
al., 1976; Mellick et al., 1977; Crapo  et al., 1984).  In the earlier reports,
                                     9-43

-------
damaged endothelial cells were those located immediately deep to denuded basal
lamina and  resulted  from sloughing of type 1 epithelial cells in the eentri-
acinar region  (Bils,  1970;  Boatman et al., 1974).   Stephens  et al.  (1974b)
reported occasional areas of endothelial swelling but the endothelium in these
areas appeared  relatively  normal  and the capillary  bed was intact (Stephens
et al,, 1974a»b).
     Morphological damage to  the  various types of interstitial cells in the
interalveolar septa  has  not been  reported.   During Q3 exposure, inflammatory
cells migrate  into the  centriacinar  interalveolar  septa (Schwartz et al.,
1976). Later, more collagen and connective tissue ground substance is found in
the interalveolar  septa  (Moore and Schwartz,  1981).  Boorman  et  al.  (1980)
                                                                          2
reported centriacinar  interalveolar septa  from rats exposed to 1568 pg/m
(0.8 ppm) Og 8  hr/day  for  20 or 90 days had thicker .blood-air barriers which
contained more  interstitium.   Crapo et al.  (1984) made a more  comprehensive
morphometric study of  centriacinar interalveolar septa from young adult rats
                    2
exposed  to  490 pg/m   (0.25 ppm)  Q3 12 hr/day  for  6 weeks.   They reported
significant  increases  in tissue thickness and  suggested that the increased
thickness was  due  to significant  increases in  all  cell  types except type 2
cells, and to increased interstitium.
     Mucous-secreting  cells  in conducting airways appear relatively resistant
                                                                         3
to 0~.   Boatman et al. (1974), in studies of cats exposed to £ 1960 pg/m  (S
1,0 ppm) of  03  via an endotracheal tube for short periods, did find limited
desquamation of these  cells,  but  the authors also observed that most appeared
intact and  increased  in  size. Castleman et al. (1977) noted roughened apical
surfaces of  mucous  cells,  which appeared to be associated with mucigen drop-
lets  near the  cell  surface, but  did not find  other  alterations in pulmonary
                                                     2
mucous cells from  monkeys exposed to 392 or 686 |jg/m (0.2 or  0.35 ppm) of 0,
8 hr/day for 7  days.   Mellick et  al. (1977) mention  that mitochondrial swell-
ing and  residual  bodies  seen in ciliated cells were not seen in mucous cells
                                                            3
in conducting airways of monkeys exposed to 980 or 1568 pg/m  (0.5 or 0.8 ppm)
of 03 8 hr/day  for  7 days.   Schwartz et al.  (1976), who  reported mucigen
droplets being  released  from the  apical surfaces t)f mucous cells and mucous
droplets trapped among cilia, did not find changes suggesting damage to organ-
ell es in rats  exposed  to 392,  980,  or  1568 ug/m3  (0.2, 0.5, or 0.8 ppm) of 03
8 or  24  hr/day for 7 days.   Wilson et al.  (1984) reported only minor changes
in trachea!  mucous  cells from bonnet monkeys continuously exposed for 3 or 7
                  3
days  to  1254 ng/m  (0.64 ppm) 03.   Using  TEM  they  reported more prominent
                                    9-44

-------
small-mucous-granule (SMG) cells which  had more abundant cytoplasm and more
specific granules.  They  speculated  that SMG cells may relate to the repair
process.  In regular mucous cells they reported fewer mucous granules  dispersed
in more cytoplasm.  The  mucous granules appeared smaller and  differed  from
controls in  that  they lacked the typical  biphasic  appearance  and had only
filamentous or granular secretory material.  No reports of damage to conducting
airways other than to ciliated and mucous cells were found.
9.3.1.3.2  Extracellular elements (structural proteins).   Although physiologic
and biochemical changes following 03 exposure suggest changes in the extracel-
lular structural  elements  of the lung,  no direct morphological  evidence has
been given of  changes  in  the  extracellular structural  elements themselves, in
contrast to  changes in their  location  or  quantity.   These physiologic and
biochemical  studies are  discussed elsewhere, in this  document  (See  Sections
9.3.2.2  and  9.3.3.6,   respectively).   Three studies  provide  morphological
evidence of  mild  fibrosis (i.e., local  increase of collagen) in centriacinar
interalveolar  septa following exposure  to  < 1960 pg/m  (< I ppm) of 03 (Last
et al., 1979;  Boorman et a!., 1980;  Moore and Schwartz,  1981).  Changes in
collagen location or  amounts, or both, which  occur with  the remodeling of
the  distal airways, were reported in two  of those  studies  (Boorman et al.,
1980; Moore  and Schwartz, 1981).  One study (Fujinaka et al.,  1985) reported
increased  connective  tissue  surrounding respiratory bronchioles from bonnet
                              2
monkeys exposed to 1254  pg/m  (0.64 ppm)  03 8 hr/day for  one year.   This
increase was due  to  Increased amorphous  extracellular matrix rather than
stainab!e  connective  tissue  fibers.   Evidence of more collagen or changes in
                                                                       3
collagen location  is in the report of dogs exposed to  1960 or 5880 pg/m  (1 or
3 ppm)  of  03 for  18 months (Freeman et al.,  1973).
9.3.1.3.3  Edema.   Morphologically  demonstrable alveolar edema, or alveolar
flooding—an effect of higher-than-ambient levels of  0, (Scheel  et  al., 1959;
                                                                      3
Cavender et  al.,  1977)—is not reported after exposures to <_ 1960 ug/m  (S 1.0
ppm) of 0-j for short or long  exposure periods (Schwartz et al., 1976; Cavender
et al.s 1978;  Mellick et  al., 1977; Eustis et al., 1981; Boorman et al., 1980;
Moore  and  Schwartz,  1981).   Mild interstitial  edema  of  conducting airways
(Mellick et  al.,  1977) and centriacinar parenchymal  structures  (Schwartz  et
al., 1976; Castleman  et al.,  1980;  Mellick  et al., 1977) is seen following
                                           3
exposure of  monkeys or rats  to < 1960 ug/m  (< 1  ppm) of  03  for  several hours
to  1 week.   Interstitial  edema is not  reported  following longer-term (i.e.,
                                         o
weeks  to  months)  exposure to < 1960 ug/m  (< 1 ppm) or less  (Cavender et al. ,
                                     9-45

-------
1977; Eustis et  a!.,  1981; Boorman et al.,  1980;  Moore and Schwartz, 1981;
Zitnik et al.,  1978).   Biochemical indicators of edema are described in Sec-
tion 9.3.3.
9.3.1.4   Considerations of DegreeofSusceptibility to Morphological Changes
9.3.1.4.1  Compromised experimental animals.  Compromised experimental animals
(e.g., those  with a  special  nutritional or  immunological  condition) in a
disease state or of young  or old  age may  respond to 0-  exposure with  greater,
lesser, or a different type of  response than  the normal, healthy, young  adult
animals usually  studied.   Some  of these may represent "at  risk" human popula-
tions.
     9.3.1.4.1.1  Vitamin E deficiency.   Rats maintained on vitamin E-deficient
diets tended to  -develop  more morphological lesions follbwing exposure to low
levels of 03 than did rats on the usual  rations (Plopper et al.,  1979; Chow et
al.,  1981).   Rats maintained on  a basal  vitamin E diet equivalent to the
average U. S.  human  adult intake were exposed  to  196 or  392 ug/m   (0.1 or
0.2 ppm) of 03 24 hr/day for 7  days.  According to LM studies, two  of the six
rats on the basal vitamin E had  increased  numbers  of macrophages  in their
centriacinar alveoli, a  typical response to higher levels of 0., (Schwartz et
                                                                        3
al., 1976).   Of five rats on the usual  rat chow diet exposed to 196 ug/m  (0.1
ppm) 0- for the same period, LM revealed no increased centriacinar macrophages.
LM analysis showed neither dietary group  had  lesions  in the ciliated  terminal
bronchiolar epithelium.   Rats in which lesions were observed by LM had increased
macrophages, according to  SEM analysis.   Analysis by TEM showed that  all rats
                   3
exposed to 196  ug/m   (0.1 ppm)  of 03 differed  from controls  in that  some of
the  centriacinar type 1 alveolar  epithelial  cells contained  inclusions and
were thicker.
     Chow et al.  (1981) fed month-old rats a basal  vitamin E~deficient diet or
that diet supplemented with  11 or 110 ppm vitamin E for 38 days,  after which
                                                         3
they  were  exposed either  to filtered air or to 196 ug/m   (0.1 ppm)  of 03
continuously for 7 days.   The morphology of six rats from each diet and exposure
group was studied using  SEM.   None of the  filtered-air control  animals had
lesions.   Of the  rats exposed to 0-, five of the six on the vitamin E-deficient
diet, four of six on  the deficient diet  supplemented by 11 ppm vitamin  E, and
one of the six on the deficient diet supplemented by 110 ppm vitamin E developed
the typical  03 lesion as seen with SEM (Schwartz et al., 1976).
     Sato et al.  (1976a,b, 1978, 1980) exposed vitamin E-deficient and supple-
                       3
mented rats to 588 ug/m  (0.3 ppm) of 03 3 hr daily for 16 consecutive days or
                                    9-46

-------
5 days a week for 7 months.  The short-term experiments (Sato et al., 1976a,b)
were marred by the presence  of  chronic  respiratory  disease  in the  rats, which
may explain the investigators' finding of large amounts of debris and numerous
small bodies  "so  thick that  the original  surface could  not  be seen" and their
failure to find the typical centriacinar 0, lesions reported by others (Stephens
et al., 1974a; Schwartz et al., 1976).  In the latter experiments, Sato et al.
(1978, 1980)  did not  find morphological differences  between the vitamin
E-depleted and  supplemented, filtered-air  control  rats or between vitamin
E-depleted and supplemented, (k-exposed rats.   They did find mild centriacinar
Og  lesions  in exposed rats  from both vitamin E-deficient and supplemented
groups.
     Stephens et al.  (1983) reported results of exposure of vitamin E-depleted
                                            3
and control young and old  rats  to  1764  ug/m  (0.9 ppm)  of 03 for I, 3, 6, 12,
34, 48, and 72 hr.  Vitamin E depletion was evaluated by determination of lung
tissue levels.   Lung response  to  ozone was based on characteristic  tissue
nodules previously reported by these authors when using a dissecting microscope
rather than on conventional  LM,  SEM,  or TEM.   They  concluded that  response to
injury and repair of the  lung was  independent of  the level of vitamin E in
lung tissue.   Most of these studies included concurrent biochemical evaluations
of oxidant metabolism and  are discussed in Section 9.3.3.
     9.3.1.4.1.2   Age at start of exposure.    Although  most exposures  use
young adult experimental animals, there are a  few reports of exposures of very
young animals (i.e., either before weaning or  very soon thereafter).
     Bartlett et  al.  (1974) exposed 3- to  4-week-old  rats  to 392  ug/m3 (0.2
ppm) of 0- for 30 days.  Lung volumes,  but not body weights, were significantly
greater in the exposed rats.  Light microscopy of paraffin sections of conven-
tionally  fixed  lungs did  not reveal  differences between exposed  and  control
rats in the lung  parenchyma or terminal bronchioles, with the exception of two
control animals  which  had  lesions of  "typical  murine pneumonia."   Morphometry
was  done  on thick sections cut by hand  with a razor blade  from the dorsal and
lateral surfaces  of  air-dried lungs rather than on the paraffin  sections of
conventionally fixed lungs.  Morphometry  using these nonrandom samples revealed
significantly increased mean alveolar chord lengths and alveolar surface area,
but  no difference in alveolar numbers.
     Freeman  et  al.  (1974) exposed month-old  rats  to 1058  or  1725 ug/m  (0.54
or 0.88 ppm)  of 0- for periods  of  4 hr  to 3 weeks.  In  addition to the centri-
acinar accumulations of macrophages and hyperplasia of  distal airway  epithelium
                                    9-47

-------
seen by others  following exposures of young adult animals, they reported an
increase in  connective  tissue elements and collagen-like strands that formed
                                                                            o
bridges across  alveolar openings.   Fibrosis was pronounced in the 1725
(0.88 ppm) group  and  sometimes  extended into terminal bronchioles.   Although
                                                     3
fixed lung volumes were not determined, the 1725-pg/m  (0.88 ppm) group required
greater volumes of fixing fluid, evidence of larger lung volumes.  In the same
research report,  Freeman et al.  (1974)  studied month-old rats exposed to 1764
    3                                3
pg/w   (0.9 ppm)  of Q3 and 1690  pg/m  (0.9 ppm)  nitrogen dioxide combined.
After 60 days  of  exposure, they  observed the gross and microscopic appearance
of advanced experimental emphysema of the type they earlier described following
nitrogen dioxide exposure (Freeman et al.,  1972).  Although others have reported
larger fixed  lung volumes  in  exposed young adult  rats  (Moore and Schwartz,
1981}, reports of emphysema  following 0- exposures are uncommon and are dis-
cussed in the next subsection of this document.
     Stephens  et  al.  (1978) exposed  rats ranging in age from  1  to 40 days old
             o
to 1666 jjg/m  (0.85  ppm) 03 continuously for 24, 48,  or 72 hr.   Rats exposed
to Og  before weaning  at 20 days of  age developed  little or no  evidence of
injury, as  evaluated by light and electron microscopy.   When exposure was
initiated after  weaning at 20 days  of age, centriacinar lesions increased
progressively, plateaued at 35 days  of  age, and  continued until  approximately
1 year of age.
     Barry et  al.  (1983) exposed 1-day-old male rats and their mother to 490
    3
(jg/m   (0.25 ppm)  of 03  12  hr/day for 6 weeks.   They  observed persistence of
the centriacinar  damage to type  1 epithelial cells and increased centriacinar
tnacrophages.    By  using  LM  and TEM morphometry of centriacinar  regions, they
reported an  increase  in both type 1  and 2 alveolar  epithelial   cells.  The
type 1 cells were smaller  in  volume, covered less surface,  and were thicker.
The authors  were  aware  of  the above study by Stephens et al.  (1978) and dis-
cussed the possibility that much of the damage they observed may have occurred
in the last 3 weeks of exposure (i.e., after weaning).  Changes in lung function
evaluated by Raub et al. (1983a) are discussed in Section 9.3.2.
     Bils (1970) studied the effects of 1176 to 2548 M9/m3 (0.6 to 1.3 ppm) of
On on  mice 4 days old and 1 and  2 months old.  From his study,  Bils concluded
that the endothelium  appeared to be the main target  of the 03,  a conclusion
not supported  by  more recent  studies, which deal mostly with other species.
81 Is did note  the lesions  were more severe in the 4-day-old mice than in the
1- or 2-month-old mice.
                                    9-48

-------
     9.3.1.4.1.3   Effect  of  pneumbnectomy.   Two  to four  weeks following
pneumonectomy of rabbits, the contra!ateral  lung  increases  in volume, weight,
collagen, and protein content to approximate that of both lungs from controls,
but alveolar  multiplication  appears dependent  on age  at surgery.   Boatman
                                                                      3
et al. (1983) exposed pneumonectomized and control rabbits to 784 [jg/m  (0.4 ppm)
of 03 7 hr/day, 5 days/week for 6 weeks.  They examined the lungs with standard
LM and  TEM morphometric techniques, but  not methods for alveolar  numbers.
Boatman and co-workers concluded that the lung growth that follows pneumonectomy
occurred after  03  exposure  and that no difference  existed  between  males and
females in this response.
9.3.1.4.2  Emphysema following ozone exposure.  The previous criteria document
for 03  (U.S.  Environmental  Protection  Agency,  1978) cites three published
research reports  in which emphysema was observed  in  experimental  animals
                                 3
following exposure to < 1960 ug/m  (< 1 ppm) of 03 for prolonged periods (P'an
et al.,  1972;  Freeman et al. , 1974; Stephens et  al.,  1976).  Since then,  no
similar exposures (i.e., same species, 03 concentrations, and times) have been
documented to  confirm these  observations-.   An additional  consideration  is  the
similarity of  the  centriacinar lesion following  0,  exposure to that  seen  in
young cigarette smokers (Niewoehner et al., 1974; Schwartz et al.,  1976; Cosio
et al., 1980; Wright et al., 1983; Fujinaka et al., 1985) and the relationship
between  cigarette  smoking and  emphysema in  humans (U.S.  Department  of Health,
Education, and  Welfare,  1967,  1969).  Further, animals exposed to  1960 pg/m
(1 ppm)  of 03  reportedly have more  voluminous  lungs than controls  (Bartlett
et al.,  1974;  Moore and  Schwartz,  1981).  Morphometry  was used  to demonstrate
enlarged subpleural  alveoli  in one of these reports (Bartlett  et al., 1974).
However, these authors indicate  that these  subpleural  alveoli may not be
representative  of the whole  lung and do not conclude that emphysema was present.
Thus, a  restudy of  these three reports  in the 1978 document appears appropriate.
      The precise  definition  of emphysema  is critical to  reevaluation  of these
reports.   Several  professional  groups  have  presented definitions of emphysema
(Fletcher  et al. ,  1959; World  Health  Organization,  1961; American Thoracic
Society, 1962).   The most recent  is the  report of a National Heart,  Lung  and
Blood Institute,  Division of Lung Diseases  Workshop (National  Institutes  of
Health,  1985).   In human lungs, "Emphysema  is  defined as a condition of the
lung  characterized by abnormal, permanent  enlargement of airspaces distal to
the terminal bronchiole,  accompanied by destruction  of their walls, and without
obvious  fibrosis"   (National  Institutes of  Health,  1985).   Destruction is
                                     9-49

-------
further defined:  "Destruction  in  emphysema is defined as non-uniformity in
the pattern of respiratory airspace enlargement so that the orderly appearance
of the acinus  and its components is disturbed and may be  lost,"  The report
further indicates "Destruction . .  .  may be recognized by subgross examination
of an  inflation-fixed lung slice .  .  .,"  In order  to  stimulate additional
research, the  definition  of emphysema  in animal  models was less restrictive.
The document states:   "An animal model of  emphysema  is defined as an abnormal
state of the  lungs  in which there is enlargement of the airspaces distal to
the terminal bronchiole.   Airspace  enlargement should be determined qualita-
tively in appropriate specimens and quantitatively  by stereologic methods."
Thus in animal models airspace wall  destruction need not be present.   However,
where information from air pollution exposures of animals is to be extrapolated
to hazards for humans, the definition of human emphysema must be considered and
the presence of airspace wall destruction documented.
     Stokinger  et al.  (1957) reported emphysematous changes in lungs from
guinea pigs, rats and hamsters, but not mice or dogs, exposed 6 hr/day,  5 days/
                                                                            3
week for 14.5  months  to a mean  concentration of slightly more than 1960 pg/m
(1 ppm) of  Og.   With  the  exception of the dogs,  mortality rates were high in
both control and exposed animals, ranging in the controls from 25 to 78 percent
and in exposed from 11 to  71 percent.    The published  report indicates that
emphysema was  present but does  not further characterize it as to the presence
of only enlarged air  spaces (Fletcher et al., 1959) or  enlarged air spaces
accompanied by destructive changes  i,n  alveolar walls (World Health Organiza-
tion, 1961; American  Thoracic  Society,  1962).   The  lungs  were fixed via the
trachea, making.them  suitable  for studies  of experimentally  induced emphysema
(American  Thoracic  Society, 1962;  National  Institutes of Health, 1985).
Stokinger et al.  (1957) attributed  the  emphysema in the  guinea  pigs  to the
observed bronchial  stenosis.   Also in the  guinea pigs were foci  of "extensive
linear fibrosis  . .  .  considered to be caused by organization of pneumonic areas."
In exposed  rats,  the  mild degree of emphysema "did  not exceed the emphysema
found in the unexposed control rats."  In exposed hamsters, "mild to moderate"
emphysema was present, but not in controls.  Emphysema is not mentioned in the
figure legends,  but three of them mention "alveoli  are  overdistended .  .  .
alveolar spaces  are dilated .  . .  dilation of alveolar ducts and air sacs."
Evidence of destruction of alveolar walls is not mentioned.   Later, however,
Gross et al.  (1965),  in an unrefereed publication abstracted from a presenta-
tion at the seventh Aspen Conference on Research in Emphysema,  reviewed the
                                    9-50

-------
lesions in the hamsters from this exposure and described a "destructive process"
that resulted in contraction of interalveolar septa not associated with enlarge-
ment of air spaces.
     Because  the  interpretation  in this EPA Criteria  Document  differs from
that in the previous document, the details need to be presented.  The signifi-
cance will be discussed in a paragraph at the end of this section.  The earlier
0, criteria document (U.S. Environmental Protection Agency, 1978) cites Stephens
et al.  (1976),  a  "long abstract"  that appears  not  to be  refereed.  This brief
article states  "rats  exposed continuously for long periods (3-5 months) to
28,200 M9/m3  (15.0  ppm) N02  or 1568 pg/m3  (0.8 ppm) 03 develop  a  disease  that
closely resembles emphysema", but does not provide additional evidence other
than citing five earlier studies by the Stanford group of investigators.   Each
of those  five references  was checked for  studies  of  animals exposed to 0.,.
Three articles describe only N09-exposed animals.  The fourth reference (Freeman
                                                             3
et al., 1973)  is  to an exposure  of dogs  to 1960  to 5880  |jg/m (1  to  3  ppm)  of  •
03 8 to  24 hr daily for  18  months and  was cited earlier  in this document.
Emphysema is not mentioned in that article.  Neither is emphysema mentioned in
the fifth  reference (Stephens ,et al., 1974b),  which was  also cited earlier  in
this document.   These investigators  did describe (Freeman et  al.,  1974) a
                                                                     3
group of month-old  rats exposed continuously for 3 weeks to 1725 jjg/m  (0.88 ppm)
of 0-j, half of which died and had "grossly inflated, dry lungs."   In this same
                                                                   3
study, they  also  exposed month-old rats to a  mixture  of 1690 pg/m   (0.9 ppm)
of NOp and 0, continuously  for 60 days, at which  time the lungs were  grossly
enlarged,  and "both grossly and microscopically,  the  appearance  of  the lungs
was characteristic  of advanced experimental emphysema" of the type they earlier
reported  following  N0? alon6 at much higher  concentrations  (Freeman et  al.,
1972).
     The  third  citation in the 0,  criteria- document (U.S.  Environmental Protection
Agency,  1978). is  to~-P'an et al.  (1972).   These  investigators exposed-rabbits
            3
to 784 M9/m  (0-4 ppm) of 03  6  hr/day,  5 days/week for 10 months.  Tissues
were fixed apparently by  immersion rather  than infusion  via  the trachea, which
is  not in accord with the American Thoracic  Society's  diagnostic standard for
emphysema,  making  emphysema lesions  much more  difficult  to evaluate  accu-
rately.   The lesions  related to emphysema are only very briefly described and
illustrated  in only  one  figure.  The  authors also  report that  "all  lungs
showed some  degree  of  inflammatory infiltrate" and "lungs  of the  sixth were so
                                     9-51

-------
congested that  visualization  of  the  mural  framework  of the alveoli was diffi-
cult."  This  is more reaction than reported in other species exposed to this
comparatively low  CL concentration.  The  rabbits were not specified pathogen-
free, nor was  the  possibility considered  that  some  lesions  could be due to
infectious agents.   Neither  did  these investigators consider the possibility
of spontaneous  "emphysema and associated inflammatory changes" which Strawbridge
(1960) described in  lungs from 155 rabbits of various ages and breeds.
     In the studies  reviewed in this section, enlargement of air spaces distal
to the terminal bronchiole have been described following 03 exposure of several
species of experimental animals.   In one study, the enlargement was quantitated
using morphometry  of air-dried lungs (Bartlett  et  al., 1974).  Destruction  of
alveolar walls  was only briefly described in two reports (P'an et al., 1972;
Gross et al.,  1965).  In one of these studies  (P'an et  al.,  1972) the lungs
were apparently fixed by immersion rather than infusion,  making a diagnosis of
emphysema less  reliable (American  Thoracic Society, 1962).   The  other study
(Gross et al.,  1965) appears to be an unrefereed long abstract rather than a
full  research  report article.   Neither  of these reports describes lesions
which unequivocally  meet the criteria for human emphysema as defined by either
the American Thoracic Society (1962) the  or  the National Institutes of Health
(1985).

9.3.2  Pulmonary Function Effects
9.3.2.1  Short-Term  Exposure.  Results of short-term 03 exposures of experi-
mental animals  are shown in Table 9-2.  These  studies were designed to evalu-
ate the acute changes in lung function associated with 03 exposure in a variety
of species (mice,  rats, guinea pigs, sheep,  rabbits, cats, monkeys, and dogs)
when compared to filtered-air exposure.
     The effects of  short-term  local Q~ exposure  of the lung periphery have
been examined in dogs by Gertner et al.  (1983a,b).   A fiber-optic bronchoscope ....
with an outside diameter  of 5.5 mm was wedged  into  a segmental airway and  a
                                   3
continuous flow of 196 or 1960 jjg/m  (0.1 or 1.0 ppm) of 0- was flushed through
this airway and allowed to escape through the  system of collateral channels
normally present  in the lung periphery.    During exposure  to either 196 or
         o
1960 ug/m  (0.1 or 1.0 ppm)  of 03, airflow resistance through the collateral
channels increased during  the first 2 min of  exposure.   Resistance of the
                                                                            3
collateral  channels  continued to increase throughout exposure to 1960 ug/m
                                   9-52

-------
TABLE 9-2.   EFFECTS OF OZONE ON PULMONARY  FUNCTION:   SHORT-TERM EXPOSURES
Ozone
concentration
Ma/n1*
196
1960


431
804
1568

470 to
2156
ppm
0.1
1.0


0.22
0.41
0.80

0.24 to
1.1

Measurement3
method
MAST



CHEM,
NBKI


NBKI

, Exposure
' duration
& protocol Observed effects
30 orin Collateral system resistance increased
rapidly during exposure, falling to
control levels at 0.1 ppm but con-
tinuing to increase at 1.0 ppm of Oa.
2 hr Concentration-dependent increase in fn
for all exposure levels. No change in
R., TV, or MV. Decreased C.dyn during
exposure to 0.4 and 0.8 pp> of 08.
12 hr Premature airway closure at 6 hr, and
1 and 3 days following exposure, ref lee-


Species
Dog



Guinea pig
(200-300 g)


Rabbit



Reference
Gertner et al.
1983a,b


Aindur et al . ,



Inoue et al . ,




,



1978



1979

                      ted by increased RV, CC,  and  CV  (6  hr
                      and 1 day only).  Maximum effect 1  day
                      following exposure, all values returned
                      to normal by  7  days.   Distribution  of
                      ventilation less uniform  6 hr following
                      exposure.  Increased lung distensibility
                      in the mid-range of lung  volumes (25-75%
                      TLC) 7 days following  exposure.
510
980
1960


666
1333
2117
2646
980

0.26 MAST
0.5
1.0


0.34 NBKI
0.68
1.08
1.35
0.5 NBKI

2.0 to
6.1 hr


2 hr



2 hr

Concentration-dependent increase in R,
during exposure. Decreased C, and D,CO
but less frequent and less marked than
changes in R, . No change in VC or
deflation pressure- volume curves.
Increased f« and decreased TV during
exposure to all 03 concentrations.
Increased R during exposure to 1.08
and 1.35 ppm of 03.
Slight increase in fg and R (to 113%
of control values) during exposure.
Cat



Guinea pig
(300-400 g)


Guinea pig
(280-540 g)
Watanabe et al . ,
1973


Murphy et al . , 1964



Yokoyama, 1969


-------
                                          TABLE 9-2.   EFFECTS OF OZONE ON PULMONARY FUNCTION:   SHORT-TERM EXPOSURES  (continued)
(Jl
Ozone L Exposure
concentration Measurement ' duration
yg/H85 ppM netted & protocol
1470 0.75 CHEM
1960 1.0
3920 2.0
1960 1 NBKI
1960 1 NBKI
Continuous
1, 3, 7 or
14 days
3 hr
6 hr/day,
7 to 8 days
Observed effects'" Species Reference
The validity of this study is ques- Rat Pepelko et al.,
tioned because of low airflow through 1980
the exposure chambers and high mortality
of exposed animals (66% mortality in rats
exposed to 1 pp« of Q3).
Reduced TLC at air inflation pressure Rabbit Yokoyaaa, 1972, 1973
of 30 cii H20, 1 to 3 days postexposure
but not at 7 days. No difference in
lung pressure- volume characteristics
during lung inflation with saline.
Increased R, and decreased C.dyn Rabbit Yokoyama, 1974
1 day following exposure. No change
in HEFV curves.
                   Measurement method:   HAST = Kl-coulometric (Hast meter); CHEM = gas phase chemiluminescence; NBKI = neutral buffered potassium iodide.
                    Calibration Method:   NBKI = neutral buffered potassium iodide.
                    See Glossary for the definition of pulmonary symbols.

-------
(1.0 ppm) of 0~, but decreased again to control levels during continued expo-
                3
sure to 196 ug/m   (0.1  ppm)  of 03-   Based on these observations, the authors
reported that tolerance appears to develop in the collateral airways to locally
                                                       3
delivered 0- at concentrations of 196 but not 1960 jjg/m  (0.1 but not 1.0 ppm)
of 03.
     Amdur et al.  (1978) measured bresrthing pattern (tidal volume, respiration
rate, and minute volume), pulmonary  resistance, and dynamic pulmonary  compli-
                                                                     3
ance in guinea  pigs  during 2-hr exposures to 431, 804, or  1568  ug/m  (0.22,
0.41, or 0.8 ppm)  of  0~,   Accelerated respiration rates  with no significant
changes in  tidal  volume were measured during exposures to  all 03 concentra-
tions,.   The onset  and magnitude of  these  changes in respiration rate were
concentration-dependent, and values of  respiration  rate remained  elevated
during a 30-min recovery period  following  exposure.  Pulmonary compliance was
significantly lower than pre-exposure  values  following 1  and 2 hr of exposure
                   3
to 804 or 1568 ug/m  (0.41 or 0.8 ppm) of 0-, and values of compliance remained
low  during  the  30-min recovery period.  Changes  in  dynamic compliance were
                                                                  3
essentially  the same  during exposure  to either  804  or 1568 ug/m   (0.41 or
0.8  ppm) of Q3.  These  investigators observed no  significant change in pulmo-
nary resistance during exposure to  0,.   If  anything,  resistance tended to
decrease throughout the exposure and recovery period.
     The lack  of  a significant  increase in pulmonary resistance  in  the Amdur
et al.  (1978) study is  in contrast to  the 113 percent increase over  pre-exposure
values in total respiratory  flow resistance measured in guinea pigs  exposed to
980  ug/m3 (0.5  ppm) of  03 for 2  hr by  Yokoyama (1969).   Watanabe  et  al. (1973)
also found  increased pulmonary flow  resistance in cats artificially  ventilated
through  an  endotracheal tube with  510,  980,  or 1960 ug/m  (0.26,  0.50, or
1.00 ppm) of  03 for 2  to  6.5  hr.   Pulmonary resistance had increased to at
least  110 percent  of  control values in all animals after 105 min of exposure
to 510 ug/m3 (0.26 ppm) of 0,, after 63 min of exposure to  980 ug/m  (0.50 ppm)
                                                3
of 03, and  after 49 min of exposure  to 1960 ug/m  (1 ppm) of 03-  Dynamic lung
compliance  was decreased during 03  exposure  in the, Watanabe et al.  (1973)
study,  as  it was in the Amdur et al.  (1978)  study.  However, changes  in pul-
monary  compliance  measured by Watanabe et al.  (1973) occurred  less  frequently
and  were less severe (based on  percentage changes from pre-exposure control
values)  than changes  in pulmonary resistance.
     Like Amdur et al. (1978) and  Yokoyama  (1969), Murphy  et al. (1964) also
measured  breathing pattern  and  respiratory flow resistance in  guinea pigs
                                    9-55

-------
during  2-hr 0~ exposures.   These  investigators found concentration-related
                                                                             3
increases in respiration rate during exposure to 666, 1333, 2117 or 2646 pg/rn
(0.34, 0,68, 1,08, or 1,35 ppm) of 0-.  Respiratory flow resistance was increased
(to 148 and 170 percent of pre-exposure values) in guinea pigs exposed to 2117
             o
and 2646 ug/m   (1.08 and 1.35 ppm) of 0~ respectively.  Pulmonary compliance
was not measured.
     The variability in measurements  of pulmonary resistance following 0-
exposure can be attributed  to a number  of  factors including the following:
frequency characteristics of the  monitoring equipment and measurement tech-
niques utilized,  the influence  of anesthetics, and the  intraspecies differ-
ences in airway reactivity  of guinea pigs.   The latter point was the subject
of critical review in the assessment of toxicological effects from particulate
matter and sulfur oxides (U.S. Environmental Protection Agency, 1982).
     Recovery of  guinea pigs from short-term Og  exposure  was substantially
different in the  above  three studies.  Animals  exposed by Amdur  et al. (1978)
showed little or  no  return  toward  pre-exposure  values for any  of the measured
parameters during a 30-min recovery period following exposure.  In guinea pigs
exposed by  Murphy et al.  (1964) and  Yokoyama  (1969),  respiration rates had
returned almost  to  pre-exposure values  by  30 min following exposure.  The
development of more  persistent  lung-function changes following 0~ exposure in
the Amdur et al.  study  (1978) may  be  attributed to the small  size and associ-
ated immaturity of these guinea pigs (200 to 300 g) compared with those in the
studies by  Murphy et al.  (1964) (300  to  400  g) and Yokoyama (1969) (280 to
540 g).  In an  earlier  study, Amdur et al.  (1952) showed  that young guinea
pigs 1  to  2 months old were significantly more sensitive to inhaled sulfuric
acid aerosols than  12-  to 18-month-old animals.  The use of ether anesthesia
and placement of  an intrapleural  catheter by Amdur et al. (1978) but not by
Murphy et al.  (1964) or Yokoyama (1969) may also  have sensitized the animals
exposed by Amdur et al.  (1978) to effects of 0~.
                                              O             .3
     Inoue et al.  (1979) exposed rabbits to 470 to 2156 M9/m  (0.24 to 1.1 ppm)
of 03  for  12  hr and performed lung function tests 6 hr,  and 1, 3, and 7 days
following exposure.   These  rabbits showed  functional  evidence of premature
airway closure with increased trapped gas at low lung volumes 6  hr, I day, and
3 days  following  exposure.    Functional  changes indicating premature airway
closure included  increased  values  of closing capacity,  residual volume, and
closing volume.   Lung quasistatic  pressure-volume measurements  showed higher
lung volumes at lung distending pressures from  0 to -10 cm of H~0.  These lung
                                   9-56

-------
function changes were greatest 1 day following exposure and had disappeared by
7 days following exposure.   Distribution of ventilation in the lung was less
uniform  in  03-exposed animals only  at  6 hr following exposure.  By  7  days
following the  initial 12<-hr  0-  exposure,  the  only significant functional
change was  an  increased  lung distensibility in  the midrange  of lung  volumes
(from 25 to 75 percent total lung capacity).
     Earlier studies  by Yokoyama (1972),  in which rabbits were exposed to
1 ppm of 0- for 3 hr, showed a timing of lung function changes similar to that
observed by Inoue  et al.  (1979).   For  both studies,  maximum  changes in 0,,-
exposed  animals were  observed  1 day  following exposure  and had  disappeared by
7 to 14 days following exposure.   However, in some aspects, the Yokoyama (1972)
study was substantially different from  that of  Inoue et al. (1979).   Yokoyama
(1972) found reduced maximum lung volume at an air  inflation pressure of 30 cm
of FLO, whereas Inoue et al. (1979)  found no difference in maximum lung volume.
Yokoyama (1972) does  not  show lung  pressure-volume curves at pressures less
than atmospheric pressure, so premature airway closure and gas trapping cannot
be evaluated in this study.  One factor  that may contribute to differences
between these two studies is the use of an excised  lung preparation by Yokoyama
(1972) compared with  evaluation  of  intact  lungs  in anesthetized rabbits by
Inoue et al. (1979).
     Yokoyama (1974) also evaluated  lung function in  rabbits following exposure
to 1960  ug/m   (1 ppm) of  03, 6 hr/day,  for 7 to 8 days.   He  found  increased
pulmonary resistance  and  decreased  dynamic compliance  in  0~-exposed  animals
                                                            «J
compared to air-exposed control  animals.  Static  pressure-volume curves and
maximum expiratory flow-volume curves were not significantly different between
the two  groups.
9.3.2.2   Long-Term Exposure.   Table 9-3  summarizes results of long-term 0^
exposures.   Raub et  al. (1983a) exposed neonatal  and young adult (6-week-old)
rats  to  157,  235,  or 490 ug/m3 (0.08,  0.12,  or 0.25 ppm) of 03  12  hr/day,
7 days/week for 6  weeks.   Lung function  changes  were observed primarily in
neonatal rats following 6 weeks of O- exposure.   Peak inspiratory flow measured
in these animals during,spontaneous  respiration was  significantly lower follow-
                                 3
ing  exposure to 235  or 490 ug/m   (0.12 or 0.25  ppm) of  03-   Lung volumes
measured at high  distending pressures  were significantly  higher in  neonatal
                            o
animals  exposed to 490 ug/m   (0.25  ppm)  of 03  for 6 weeks than  in  control
animals.  These results  are consistent with increased  lung volumes  measured
during  lung inflation with either air  or saline by  Bartlett et  al.  (1974)
                                   9-57

-------
                                                TABLE  9-3.   EFFECTS OF OZONE ON  PULMONARY  FUNCTION:   LONG-TERM EXPOSURES
vo


ui
oo
Ozone
concentration
(jg/«d ppi
157 0.08
235 0.12
490 0.25
392 0. 2
392 0.2
1568 0.8
784 0.4
882 0.45
980 0. 5
1568 0.8
Measurement3 '
method
CHEH
MAST,
NBKI
UV or CHEM
NBKI
NBKI
i
MAST
|w.
NBKI
i
Exposure
duration
& protocol
6 weeks,
12 hr/day,
7 days/week
28 to 32 days,
continuous
62 exposures,
6 hr/day,
5 days/week
6 weeks,
7 hr/day,
5 days/week
6 to 7 weeks,
6 hr/day,
6 days/week
7, 28, or
90 days;
8 hr/day
Observed effects0
Increased end expiratory lung volume
in adult rats and increased lung
volumes at high distending pres-
sures in neonatal rats exposed to
0.25 ppm of 03. Reduced peak inspira-
tory flow in neonatal rats exposed
to 0.12 or 0.25 ppm of 03.
Increased lung distensibility in
03-exposed rats at high lung
volumes (95-100% TLC) during in-
flation with air or saline.
Increased R. (not related to con-
centration) in rats exposed to
0.2 or 0.8 ppm of 03. Lung volumes at
high distending pressures (VC and TLC)
were increased at 0.8 ppm and FEF2s and
FEF10 were decreased at 0.2 and 0.8 ppm
of 03.
Increased alveolar wall extensibility
at yield and break points, increased
hysteresis ratio, and decreased stress
at moderate extensions. Fixed lung
volume increased 15%. Lung growth'
following pneumonectomy prevented
these changes to 03 exposure.
No effect of exposure on lung pressure-
volume curves.
Decreased quasistatic compliance (not
related to concentration).
Species
Rat
(neonate or
6-week-old
young adult)
Rat
(3 to 4 weeks)
Rat
(10 weeks)
Rabbit
Rat
Monkey
(Bonnet)
Reference
Raub et al . , 1983a
Bartlett et al. ,
1974
Costa et al . , 1983
Martin et al.,
1983
Yokoyama and
Ichikawa, 1974
Eustis et al., 1981

-------
                                         TABLE 9-3.  EFFECTS OF OZONE ON PULMONARY FUNCTION:  LONG-TERM  EXPOSURES   (continued)
vo



Ozone
concentration Meas
|jg/raa ppm • me
980 0.5 MA
1960 1.0 NB
. Exposure
urement * duration
thod & protocol Observed effects0 Species Reference
ST, 3 and 6 hr/day Increased resistance of central airways Rat Yokoyama et al.,
O for up to 60 after 3-hr daily exposures to 1.0 ppm 1984
days for 30 days; increased resistance of
peripheral airways after 6-hr daily
exposures to 0.5 ppm 03 for SO days.
1254 0.64 UV', 7 or 20 days Increased peripheral resistance in Rat Kotlikoff et al.,
NBKI rats exposed for 7 days but not 1984
. 20 days; decreased lung reactance
! at high frequencies in both groups.
1254 0.64 UV
NB
, 1 year, Following 6 months of exposure, venti- Monkey Wegner, 1982
O 8 hr/day, lation was less homogeneous and R. (Bonnet)
7 days/week was increased. Following 12 months
of exposure, R, remained elevated
and forced expiratory maneuvers showed
small airway dysfunction (decreased
FEVj and FEFja-s). During the 3-nonth
recovery period following exposure,
C.st decreased.
                 Measurement method:  MAJST = Kl-coulometric (Mast meter); CHEM = gas phase chemiluminescence; UV = UV photometry.

                 b-
                  Calibration method:  NBKI = neutral buffered potassium iodide.


                  See Glossary for the definition of pulmonary symbols.

-------
                                                                o
following exposure  of young  rats  (3-  to  4-week-old)  to  392 ug/m  (0.2 ppm) of
Og continuously  for 28  to 32 days.   Moore  and  Schwartz (1981)  also  found an
increased fixed  lung volume  (following  lung perfusion at  30 cm inflation
pressure with  Karnovsky's  fixative)  after 180 days of continuous exposure to
        2
980 ug/m  (0.5 ppm)  of  0-.   Yokoyama and Ichikawa (1974) found no change in
                                                                         3
lung  static pressure-volume  curves'   in  mature  rats  exposed  to 882 ug/m
(0.45 ppm) of Og 6 hr/day, 6 days/ week for 6 to 7 weeks.
     Martin et al.  (1983)  studied the mechanical properties  of the  alveolar
                                      3
wall from rabbits exposed to  784  ug/m (0.4 ppm)  of  0~,  7 hr/day, 5  days/week
for 6 weeks.   A  marked  increase  in the maximum extensibility of the alveolar
wall and a  greater  energy loss with  length-tension cycling (hysteresis) were
found following exposure.  A 15-percent increase in fixed lung volume following
perfusion at 20  cm  of H90 was  also  reported  following  0, exposure,  which is
                        
-------
former exposure while  R,  increased only at  lower pressures after the  latter
exposure.   The authors interpreted these changes as an indication of increased
                                                        3
central airway  resistance  in rats exposed to  1960  |jg/m  (1 ppm) of 0,  for
                                                                              3
30 days and increased peripheral airway resistance in rats exposed to 980 ug/m
(0.5 ppm)  of 0,, for 60 days.  These changes were also consistent with morpholog-
ical findings of  greater mucous secretions  in  large  bronchi  of the animals
                    2
exposed to  1960 jjg/m   (1 ppm)  of 0, and in the peripheral airways of animals
                    3
exposed to  980 (jg/m  (0.5 ppm)  of  0~.   No  changes in  static deflation  volume-
pressure curves of  the lungs were found after either exposure nor were there
any differences in  effects  that could  be attributed to  the age  of rats at the
start of exposure.
     Eustis et al.  (1981) evaluated.lung function in bonnet monkeys (Macaca
radiata) exposed  to 980  or  1568 jjg/m3  (0.5 or  0.8 ppm)  of 0,,, 8 hr/day for 7,
28, or 90 days.   This study appeared to be preliminary (range-finding) for the
long-term study reported by Wegner (1982).  Only a limited number of animals
were evaluated at each time point  (1 per exposure group at 7 days, 2 at 28 days,
and 3  at 90 days).  With so few animals tested and tests  made following  three
different exposure  periods,  little significant lung-function data related to
0-  exposure were  generated.  When  pooling  results from  all exposure  times and
0- concentrations, quasi-static  lung compliance was significantly different in
Og-exposed  animals  than  in  control animals.   Compliance  tended to  decrease
from pre-exposure values in control animals and increase  in 0,-exposed animals.
     Wegner (1982)  evaluated  lung function  in 32 bonnet monkeys, 16 of which
                         o
were exposed to 1254 ug/m   (0.64 ppm) of 03, 8 hr/day, 7 days/week for 1 year.
Lung function tests were performed pre-exposure, following 6 and 12  months of
exposure,  and following  a  3-month postexposure recovery period.  In addition
to  measurements of  carbon  monoxide diffusion  capacity  of the  lungs (D,co),
lung volumes,  quasi-static  pulmonary  compliance (C  .,.O and  partial and
maximum expiratory  flow-volume  curves by standard techniques, frequency depen-
dence  of compliance and  resistance and pulmonary impedance from 2-32Hz  were
measured by a forced  oscillation  technique.   The  addition of  these  latter
measurements may  elucidate  more clearly than  ever before the site and nature
of  lung impairment  caused by exposure to toxic compounds.
     Following six  months of 03 exposure, pulmonary  resistance and frequency
dependence  of  pulmonary compliance were  significantly  increased.   After 12
months, the 0-  exposure  had significantly increased  pulmonary  resistance and
inertance (related  to the pressure required  to accelerate air and lung tissue),
                                   9-61

-------
and forced expiratory  maneuvers  showed decreased flows at  low lung volumes
(12.5 percent VC) and decreased volume expired in 1 sec (FEV-,).  Wegner (1982)
suggested that because lung volumes and pulmonary compliance were not affected
in Og~exposed animals, changes in  forced expiratory function-were more  likely
caused by narrowing  of the peripheral airways than by decreased small airway
stiffness.   Rigid analysis  of the pulmonary impedance data by linear-lumped-
parameter modeling suggested that the increase in pulmonary resistance was due
to central  as well as peripheral  airway narrowing.
     During the 3-month recovery period following exposure, static lung compli-
ance tended to decrease  in  both 03-exposed  and control animals.  However, the
decrease in compliance was significantly greater in 0~-exposed animals than in
control animals.. No other significant differences were measured following the
3-month  recovery period, although  values  for 03-exposed  animals remained
substantially different  from  those for control animals, suggesting that full
recovery was not complete.
     The forced  oscillation technique has  also been utilized in rats exposed
            q
to 1254 (jg/m  (0.64  ppm)  of 0- for either  7  or  20 days (Kotlikoff  et  a!.,
1984).   In an attempt to further  characterize 0~-induced  changes in central
and peripheral distribution of mechanical properties of the respiratory system,
impedance spectra of Og-exposed  rats were  compared to the  spectra  of normal
rats.  The effective resistance  was higher at all  frequencies in the 7-day
exposed  rats  but no consistent differences were observed by  20 days.  The
effective reactance, however, was significantly lower than control in both the
7- and 20-day exposed rats.  These changes  in  respiratory system impedance
demonstrate evidence of  mechanical alterations in the peripheral airways of
rats for as long as  20 days of 0, exposure.
9.3.2.3  Airway  Reactivity.   Ozone potentiates the effects  of  drugs  that con-
strict airway  smooth muscle  in mice,  guinea pigs,  dogs,  sheep, and humans
(Table 9-4).   Early experimental  evidence for hyperreactivity to  broncho-
constrictive drugs  following  03  exposure was provided by  Easton and Murphy
(1967).  Although much of their work was done with very high 0, concentrations
                    o
(9800  to 11760 (jg/m ,  5  to  6  ppm), they did show that mortality from a  single
subcutaneous injection of histamine was higher in guinea pigs exposed to 980
or 1960  ng/rn3 (0.5  or 1 ppm) of  03 for 2  hr (33 and 50 percent  mortality,
respectively) compared with the mortality of air-exposed control animals.  The
animals  appeared to die  from massive  bronchoconstriction,  with the lungs
remaining  fully  inflated instead  of  collapsing  when  the chest was  opened.
                                    9-62

-------
TABLE 9-4.  EFFECTS OF OZONE ON PULMONARY FUNCTION:  AIRWAY REACTIVITY
Ozone
concentration
ug/m3
196
1568
196
1568
196
1960
vo
S 980
980
1568
980
2156
980
1960
3920
ppm
to 0.1 to
0.8
0.1
0.8
0.1
1.0
0.5
0.5
0.8
0.5
1.1
0.5
1
2
Exposure
Measurement duration
method & protocol
CHEM 1 hr
CHEM 1 hr
MAST 10-30 min
CHEM 2 hr
MAST Continuous,
13 to 16 days
of 03 ex-
posure in
four periods -»
(3 to 5 days
each) separ-
ated by 3 to
8 days of
breathing
air
NBKI 2 hr
CHEM 2 hr
Observed effects Species
Subcutaneous injection of histamine 2 hr following 03 Guinea pig
exposure caused a greater increase in R. following expo- (200-300 g)
sure to 0.8 ppm of 03 and a greater decrease in C.dyn
following exposure to all 03 concentrations (magni-
tude of C, changes not related to 03 concentration).
Decreased diaphragm and lung cholinesterase activity; Guinea pig
parathi on-treated animals had increased peak airway
resistance compared to controls, but the difference was
not statistically significant.
Bilateral vagotomy: completely blocked increased peri- Dog
pheral lung resistance from 0.1 ppm of 03 but not histamine;
only partially blocked response from 1.0 ppm of 03.
Histamine- induced airway reactivity increased during
1.0 ppm but not 0.1 ppm of 03 exposure and was not blocked
by atropine or vagotomy.
Increased number of mast cells and lymphocytes in tracheal Sheep
lavage 24 hr after exposure.
Repeated exposures to 0.5 or 0.8 ppm of 03 plus aerosolized Mouse
ovalbumin resulted in greater mortality from anaphy lactic
shock produced by intravenous injection of ovalbumin
compared with effects of ovalbumin injection in mice
repeatedly exposed to ovalbumin aerosols but no 03.
Increased histamine- induced mortality immediately Guinea pig
following exposure to 0.5 or 1.1 ppm of 03.
Increased airway reactivity to aerosolized carbachol Sheep
24 hr but not immediately following exposure to
980 ug/m3 (0.5 ppm) of 03 with no change in R,,
FRC, C.st, or tracheal mucous velocity. Increased
R. 24 hr following exposure and airway reactivity
immediately and 24 hr following exposure (1 ppm).
Reference
Gordon and Amdur,
1980
Gordon et al . , 1981
Gertner et al . ,
1983a,b,c, 1984
Kaplan et al. ,
1981
Sielczak et al., 1983
Osebold et al. ,
1980
Easton and Murphy,
1967
Abraham et al . , 1980

-------
                                          TABLE 9-4.   EFFECTS OF OZONE OH PULMONARY FUNCTION:  AIRWAY  REACTIVITY  (continued)
Ozone
concentration
pg/M^ ppm
980 0.5
980 0.5
1960 1.0
Exposure
Measurement duration .
nethod & protocol Observed effects
CHEH 2 hr/day No effect on airway responsss to inhaled carbachol
for 2 days 1 day after Oa exposure; airway reactivity increased
34X and airway sensitivity increased 31% with intra-
venous carbachol challenge.
CHEH 2 hr Airway responsiveness and airway permeability to hista-
mine increased 1 day after exposure to 0.5 ppm 03
(n=6) and in 4/7 exposed to 1.0 ppm 03; directional
changes in airway responsiveness paralleled direc-
tional changes in airway permeability.
Species Reference
Sheep Abraham et al., 1984a
Sheep Abraham et al., 1984b
1100 to    0.56 to       CHEM
1666       0.85
 I
O\
                                            2 hr
                               Abnormal,  rapid,  shallow breathing in conscious  dogs
                               while walking on  a treadmill  following 03  exposures.
                               Maximal  1- to 3-hr postexposure,  normal  24-hour  post- •
                               exposure.   Abnormal breathing not affected by drug-
                               induced  bronchodilatation (inhaled isoproteronol)  but
                               abolished  by vagal  cooling.   Increased respiration rate
                               caused by  inhalation of aerosolized histamine after  03
                               exposure also blocked by vagal  cooling but not by
                               isoproteronol.
                                                                                  Dog      Lee et al., 1979
      1313
           0.67
CHEH
                                            2 hr
                  Abnormal, rapid, shallow breathing during exposure to
                  air containing low 02 or high C02 immediately following
                  Q3 exposure.   Abnormal breathing not affected by
                  inhaled atropine aerosols or inhaled isoproteronol
                  aerosols but abolished by vagal  cooling.
                                                                Dog       Lee  et  al.,  1980
      1372 to
      2352
           0.7 to
           1.2
CHEH
2 hr
Greater increase in R,  caused by histamine aerosol
inhalation 24 hr following 03 exposure.  No hyper-
reactivity to histamine 1 hr following 03 exposure.
Drug-induced bronchodilatation (inhaled isoproteronol)
blocked any increase in R,  before or after 03 exposure.
Inhalation of atropine or vagal cooling (to block reflex
bronchoconstriction) prevented 03-induced reactivity
to hi stamina.
Dog       Lee  et  al.,  1977

-------
                                    TABLE 9-4.  EFFECTS OF OZONE ON PULMONARY FUNCTION:  AIRWAY REACTIVITY  (continued)
Ozone Exposure
concentration Measurement duration
W>
CTi
01
ug/m3
I960
2352
1960
4312
5880
5880
ppm method & protoco
1.0 CHEM 1 hr
1.2
1.0 UV 2 hr
2,2
3.0
3.0 UV 2 hr
1 . Observed effects
R. increased and C. decreased with subcutaneous
hTstamine 2 hr after exposure; responsiveness was not
blocked by atropine or vagotomy. No change in static
compliance after subcutaneous histamine injection.
Marked increase in airway responsiveness to inhaled ACh
and histamine 1 hr after exposure; increased to a lesser
degree 1 day later, and returned to control levels by
1 week. Effects possibly linked to acute inflamma-
tory response.
Airway responsiveness to inhaled ACh was prevented
by indomethacin pretreattnen not the airway infiltra-
tion by neutrophils. Both responsiveness and neutrophil
infiltration were prevented by hydroxyurea pretreatment.
Species Reference
Guinea pig Gordon et al.,
Dog Holtzman et al
1983a,b
Fabbri et al . ,
Dog 0' Byrne et al.
1984a,b

1984
" >
1984
1
5880      3.0
UV,
2 hr
SR   measured with intravenous ACh and/or inhaled ACh or
metnacholine increased similarly 14 hr after after expo-
sure; airway reactivity to inhaled bronchoconstrictors
returned to baseline levels 2 days after exposure while
responses to Intravenous ACh persisted.
Guinea pig   Roun and  Hurl as,  1984
5880      3.0
UV
2 hr
SR   increased with intravenous ACh; maximal response
2 nr after exposure; remission by the 4th day.  Airway
infiltration of neutrophils occurred later and lasted
longer than airway reactivity.
Guinea pig  Murlas  and  Roum  (1985)
 Measurement method:   MAST - Kl-coulometric  (Mast meter); CHEM = gas phase cherailuminescence; NBKI = neutral buffered potassium iodide.

 See Glossary for the  definition of pulmonary symbols.

-------
     Abraham et al.  (1980)  evaluated  airway  reactivity  in  sheep  from measure-
ments of pulmonary resistance following inhaled carbachol aerosols.   Carbachol
causes bronchoconstriction  by stimulating airway  smooth muscle  at  receptor
sites that are normally  stimulated by release of acetylcholine from terminals
of the  vagus nerve.   Pulmonary resistance  during inhalation  of carbachol
aerosols was significantly higher than pre-exposure values at 24 hr postexposure
                                                           o
but not  immediately  following a 2-hr  exposure to  980 jjg/m  (0.5 ppm) of Og.
This 07 exposure did not affect resting end-expiratory lung volume (functional
                                                                            3
residual  capacity) or  static  lung compliance.  In  sheep  exposed  to 1960 ug/m
(1 ppm)  of 0~ for 2  hr,  baseline resistance  (before carbachol aerosol inhala-
tion) was elevated 24 hr following exposure, and airway reactivity to carbachol
was increased immediately and 24 hr following 0~ exposure.
     To determine if 0,,-induced airway secretions  limit  the  penetration of an
inhaled  bronchoconstrictor,  Abraham  et al.  (1984a) compared airway respon-
siveness to inhaled and intravenous carbachol before and 24 hours after exposure
                                                   3
to Og.  Adult female sheep were exposed to 980 pg/m  (0.5 ppm) of 0~, 2 hr/day
for 2 consecutive days.  Airway sensitivity was defined as the largest increase
in specific lung resistance after carbachol challenge and airway reactivity as
the slope of the  dose-response curve.  There were no significant differences
between pre-  and postexposure responses to  inhalation challenge.  However, Og
exposure increased mean airway reactivity and sensitivity by 34 and 31 percent,
respectively, using  intravenous challenge.   Since carbachol causes direct
stimulation  of airway  smooth  muscle,  the authors  .suggested  that 0- may have
decreased penetration  of the  inhaled  carbachol to  the airway smooth muscle as
the result of increased  airway secretion.  This hypothesis is supported by a
previous study showing that changes  in airway responsiveness to inhaled his-
tamine following exposure  to  03 may  have  been  related  to changes in airway
permeability to histamine (Abraham et al., 1984b).
     A study in awake  guinea  pigs comparing  changes'in  airway reactivity with  ~
intravenous and inhaled bronchoconstrictors after  exposure to a  high concentra-
tion of CL has suggested that mechanisms other than increased airway permeabi-
lity may be  involved.   Roum and Murlas (1984) measured changes in specific
airway resistance with intravenous and/or inhaled  acetylcholine  or methacholine
                                                3
from 4 hr to 2  days  after exposure to 5880  M9/m   (3.0  ppm)  of 03 for 2 hr.
Airway hyperreactivity by  either  route was similar within 14 hr of exposure.
Two days  after exposure airway reactivity to bronchoconstrictor inhalation
                                   9-66

-------
returned to  baseline levels  while  responses 'to  intravenous acetylcholine
persisted.   The consistent early changes in airway reactivity after 0- exposure
with either  intravenous or inhaled bronchoconstrictors  indicated that this
response may be  independent  of the route  of  delivery.   However, it is also
possible that  there  may be more than  one mechanism  responsible  for  0,-induced
airway  hyperreactivity  depending  on  the  concentration  of  0, reaching the
airway  tissue  and  on interspecies differences in cellular  responsiveness  to
°3-                                            .
     Gordon  and  Amdur  (1980) evaluated airway reactivity to subcutaneously
injected histamine in  awake  guinea pigs following  a  1-hr exposure  to  196  to
         3
1568 |jg/m  (0.1 to 0.8 ppm) of 0,.  Airway reactivity to  histamine was maximal
2 to 6  hr following 03  exposure and returned to control  levels by 24 hr follow-
ing  exposure.   The  histamine-induced  increase in pulmonary  resistance was
                                           3
greater in guinea pigs  exposed to 1568 ug/m  (0.8 ppm) of 0,  than in air-exposed
control  animals.   Pulmonary  compliance  decreased more  following  histamine
injection in all  O^-exposed  groups than in air-exposed  controls, but there
were no differences  in  the histamine-induced decreases in pulmonary compliance
                                                           3
between any of the 03 concentrations (from 196 to 1568 ug/m  ; 0.1 to 0.8 ppm).
     Gordon  et al. (1981)  studied the effect  of 03 on tissue cholinesterases
to see  if they were  responsible for the bronchial reactivity  observed following
challenges with bronchoconstrictor analogs of  acetylcholine.  Guinea pigs were
                                         3
exposed  to  clean  air or 196  or 1568  ug/m   (0.1 or 0.8  ppm)  of  0~  for 1 hr.
After  2 hr,  brain,  lung, and diaphragm samples were  analyzed for  cholines-
terase  activity.   Brain cholinesterase activity was  not affected,  but lung
                                                                          3
cholinesterase  underwent a 17  percent decrease  in activity at 196  ug/m
                                                   3
(0.1 ppm) and  a 16  percent  decrease  at  1568 ug/m   (0.8 ppm).   Ozone at
         3
1568 |jg/m   (0.8 ppm) also  decreased the diaphragm cholinesterase activity by
14 percent.  To  provide long-term inhibition  of  cholinesterase, guinea pigs
were treated with parathion,  an irreversible cholinesterase  inhibitor.  Airway
 resistancei  tended  to increase following histamine challenge in the parathion-
 treated  guinea pigs,  but  the difference was not  statistically significant
 because  of  large variations  in response.  The authors  suggested that cholines-
 terase inhibition  by 0^ may  contribute  to the Q,,~induced bronchial  reactivity,
 as  already  reported.   Presumably,  the decreased cholinesterase activity could
 result  in  higher  acetylcholine  concentrations  in the bronchial  muscle.   A
 cholinergic-related  stimulus, such as  occurs with 0- exposure,  should then
                                    9-67

-------
 increase the contraction of the bronchus.  The persistence of this activity is
 not  known.
     Gordon et al.  (1984)  extended these studies to determine  the site of the
 airway  hyperresponsiveness  to histamine-induced airway constriction after Og
 exposure.  Anesthetized guinea pigs were evaluated  for response to subcutaneous
                                                  o
 histamine 2 hr after exposure to 1960 or 2352 ug/m   (1.0 or 1.2 ppm) of 0, for
 1 hr.   Respiratory  resistance increased and dynamic compliance  decreased in
 Qg-exposed animals, as previously  reported (Gordon  and Amdur, 1980).   However,
 static  compliance  changes  after histamine were  similar  in the air- and  Q~-
 exposed animals,  suggesting that the site of  hyperresponsiveness was  in  the
 conducting airways  rather  than the parenchyma.   In  addition, enhanced airway
 responsiveness to histamine was not blocked by atropine or vagotomy, indicating
 a minimal  level of vagal involvement.  The significantly greater increase in
 respiratory resistance  caused by  efferent electrical stimulation of the vagus
 in 0_-exposed  animals  suggested that  other mechanisms,  such  as changes in the
 airway  smooth  muscle,  may be responsible for the hyperexcitability following
 DO exposure.   However, i_n -vitro studies on isolated parenchyma! strips removed
 from the  lungs .of air- and Qg-exposed animals failed to show any differences
 in the  contractile responses to histamine or carbachol.
     Lee  et al. (1977) evaluated  airway  reactivity in  Og~exposed dogs from
 changes in pulmonary resistance induced by histamine aerosol inhalation.  Dogs,
 were exposed  to 1372 to 2352 ug/m3 (0.7  to  1.2  ppm) of 03 for 2 hr.   Airway
 reactivity to  inhaled  histamine aerosols was significantly greater 24 hr but
 not  1 hr after 0, exposure.  Bronchodilatation induced by  inhalation of isopro-
 teronol aerosols  prevented any change in resistance following  histamine expo-
 sure.   This  experiment showed that the increased resistance  normally observed
 following  histamine exposure was  caused by  constriction  of airway smooth
 muscle  and  not by edema or  increased mucous  production, which would  not be
 prevented  by  isoproterenol  bronchodilatation.   Administration of atropine
 (which  blocks  bronchoconstrictor  activity  coming from the"vagus nerve)  or
 vagal cooling (which blocks both  sensory receptor activity traveling from the
 lung to the brain and  bronchoconstrictor activity going from the brain to the
 lung) decreased the response to histamine both before and  following 03 exposure
 and  abolished  the hyperreactive  airway  response.   These experiments  showed
"that the  increased sensitivity to histamine following 03 exposure was caused
 by heightened  activity of vagal bronchoconstrictor  reflexes.
                                    9-68

-------
     Although the work of Lee et al. (1977) provides evidence that stimulation
of vagal reflexes by histamine is in part responsible for the increased airway
reactivity found  in  dogs  following 03 exposure, Kaplan  et  al.  (1981) found
that local responses to histamine in the lung periphery may not be mediated by
a significant  vagal  component.   When monodispersed  histamine  aerosols were
delivered to  separate sublobar  bronchi  in dogs through  a  5.5 mm diameter
fiber-optic bronchoscope, collateral airflow  resistance  increased both before
and after bilateral cervical vagotomy.   In follow-up studies that used similar
techniques,  Gertner  et al.  (1983a,b,c;  1984) described  the  role  of vagal
reflexes in the response of the lung periphery to locally administered histamine
and DO-   Collateral  resistance  increased during separate 30-min exposures to
               3                            ~6     3
either  196 (jg/m   (0.1  ppm)  of 0, or 1.5 xlO   mg/m  of  histamine.   However,
although parasympathetic blockade  (atropine  or bilateral cervical vagotomy)
prevented the  responses  to  03,  it did not prevent the responses to histamine
(Gertner et al. ,  1983b).  To  determine if 03~induced increases in collateral
resistance in the lung periphery were dependent on vagal reflexes, aerosolized
neostigmine was  administered  locally to maintain parasympathetic tone.   Re-
                   3
sponses to 196 (jg/m   (0.1 ppm) of  03 in  the  lung periphery  were enhanced  only
if the  vagi were intact and were  limited  to the challenged region (Gertner
et al. , 1984).   When larger areas were  exposed,  vagally mediated responses
occurred in both lungs.  In addition, the characteristics of responses to high
concentrations of 0- differ markedly from responses to the  lower concentrations.
                              3
A 30-min  exposure to  196 (jg/m   (0.1 ppm)  of 0, did not affect  the  airway
responsiveness to  histamine,  but when the 0- exposure was  increased to 1960
    3
|jg/m  (1.0 ppm) for 10 min,  histamine produced greater increases in collateral
resistance that were not abolished by parasympathetic blockade (Gertner et al.,
1983c).  Exposure to 1960 (jg/m  (1.0 ppm) of 03 for 30 min produced an increase
in collateral  resistance  that was mediated by  the parasympathetic system in
the early phase  of the response and related  in  part to  histamine release in
the late  phase of the response  (Gertner et  al.,  1983a).   Results from this
series  of studies by Kaplan et al. (1981) and Gertner et al. (1983a,b,c; 1984)
are difficult  to interpret  because of the small  numbers of animals in each
test group and large  variations  in response.   In addition,  because peripheral
resistance contributes  only a small part to  total pulmonary  resistance,  the
findings of these authors do not necessarily  contradict the work of Lee et al.
(1977).   Rather,  all  the  studies taken  together suggest  that  the periphery of
                                   9-69

-------
the lung may  respond differently from the  larger  conducting  airways during
exposure to Qg and that factors in addition to vagal bronchoconstrictor reflexes
can produce an increased airway reactivity to histamine.
     Holtzman  et  al.  (1983a)  reported the time course  of Q~-induced airway
                                                     ~      o
hyperreactivity in dogs exposed to 1960 and 4312 ug/m  (1.0 and 2.2 ppm) of 0-
for 2 hrs.   Airway responsiveness to acetylcholine in 7 dogs increased markedly
                                                               3
1 hr, and to  a lesser  extent,  24  hr after  exposure  to 4312 jjg/m   (2.2 ppm)  of
0~, returning  to  control  levels by 1 week after exposure.  Ozone- induced in-
creases in  airway  responsiveness to  histamine were  similar  following exposure
            3
to 1960 |jg/iii   (1.0 ppm)  of 03, but data were reported  for  only 2 dogs.  The
authors suggested that the time course of the 03 effect may be linked to acute
airway inflammation.   In a coincident publication,  Holtzman  et al.  (1983b)
found a strong association between airway hyperreactivity and trachea!  inflam-
                                                           2
mation in dogs 1  hr  following a 2-hr exposure to 4116 jjg/m  (2.1 ppm)  of 0,.
Airway reactivity  was assessed  from the  increase  in pulmonary resistance
following inhalation  of acetylcholine aerosols,  and airway reactivity was
increased in 6 of  10  0,,-exposed dogs.  The  number of neutrophils present in  a
trachea!  biopsy, a measure of inflammation, was increased only in the 6 dogs
that were hyperreactive  to acetylcholine.   These observations have recently
been extended  to show an association of 0,-induced increases in airway respon-
siveness with  inflammation in more distal  airways (Fabbri et al., 1984).  The
number of neutrophils as well  as ciliated  epithelial  cells  in  fluid recovered
from bronchoalveolar  lavage was increased in 5 dogs that were hyperreactive to
                                                     q
acetylcholine  following  a  2-hr exposure  to  5880 jjg/m  (3.0  ppm) of 03 without
significant changes in the numbers of macrophages , lymphocytes, or eosinophils.
In dogs depleted of neutrophils by treatment with hydroxyurea, 0™ exposure to
         3                                                  '
5880 ug/m   (3.0 ppm)  of 0- for 2 hr caused the desquamation of epithelial
cells but airway responsiveness to inhaled acetylcholine was prevented (O'Byrne
et al., 1984a).  This observation suggests that Ov"induced hyperresponsiveness
may depend on  the mobilization of neutrophils into the airways.
     The authors have expanded their work to speculate that the neutrophils
produce mediators  that are responsible  for the  increased responsiveness of
airways (O'Byrne  et al.,  1984b).  In dogs pretreated with indomethacin, a
                                                         3
prostaglandin  synthetase inhibitor, exposure to 5880 jjQ/m  (3.0 ppm) of 03 for
2 hr had no effect on airway  responsiveness to inhaled acetylcholine but there
was a  significant  increase in the number of neutrophils in  the airway epithe-
lium.  While  these results  suggest  that oxidation products of arachidonic
                                   9-70

-------
acid, possibly prostaglandins  or thromboxane (O'Byrne et al., 1984c), may be
released by the neutrophils, they are not conclusive.  Therefore, the identity
of the  specific  inflammatory cells or of the responsible mediators  is still
uncertain.
                                                                3
     The results of trachea! lavage in sheep exposed to 980 ug/m  (0.5 ppm) of
Og for 2 hr suggest that the migration of mast cells into the airways may also
have important implications  for reactive airways and allergic airway disease
(Sielczak et al., 1983).  Nasotracheal-tube exposure to 0., in 7 sheep resulted
in an increased number of mast cells and lymphocytes 24 hr after exposure, sug-
gesting an association between an enhanced inflammatory response and 03~induced
bronchial  reactivity  reported previously  in sheep (Abraham  et al.,  1980).
     Additional  evidence  in guinea pigs  suggests  that 03~induced bronchial
hyperreactivity may be due to airway mucosal injury and mast cell infiltration
(Murlas and Roum, 1985).  Specific airway resistance was measured as a function
of increasing intravenous acetylcholine doses for varying periods of from 2 hr
                                        g
up to 4 days after exposure to 5880 pg/m  (3.0 ppm) of 03 for 2 hr.  The largest
airway response to acetylcholine challenge occurred 2 hr after 0- exposure with
complete remission by the fourth day.  Meutrophil infiltration occurred later and
lasted longer despite the remission in airway hyperreactivity suggesting that the
influx of neutrophils was a result of the initial damage and  not a direct cause of
increased airway responsivensss.
     Increased drug-induced  bronchoconstriction  is not the only indicator of
airway hyperreactivity following 03 exposure.  Animal experiments were designed
to  investigate  the mechanisms responsible  for the abnormal,  rapid, shallow
breathing  found  in human subjects exercising during experimental 0- exposure
compared  with  subjects exercising  in clean air (Chapter  10).   Lee et al.
(1979,  1980) showed that  abnormal,  rapid,  shallow  breathing  in  conscious  dogs
                                                          3
immediately following  2-hr  exposures  to  1100 to  1666 pg/m  (0.56 to 0.85  ppm)
of 0-  was a hyperreaetive airway  response.   This  abnormal breathing pattern
was  elicited by  mild  exercise,  hi stand ne aerosol inhalation,  or breathing air
with reduced oxygen (02) or elevated carbon dioxide  (COp) concentrations.  The
rapid,  shallow breathing observed in dogs following 03 exposure was  not affec-
ted  by drug-induced  bronchodilatation (inhaled isoproteronol aerosols) or by
blocking  vagally induced bronchoconstriction with  atropine.   In all cases,
rapid,  shallow  breathing was abolished by  vagal cooling,  which blocked the
transmission of  sensory nerves located  in  the airways.   These  investigators
(Lee et al.,  1979, 1980) suggest that the  rapid,  shallow breathing observed
                                   9-71

-------
following 0~  exposure in dogs  is  caused  by heightened activity of  sensory
nerves  located  in the airways.  The  increased reactivity of these  sensory
nerves  is  independent of  smooth muscle  tone  (either bronchodilatation or
bronchoconstri cti on).
     Studies of  lung  morphology following 03 exposure  showed damage to the
respiratory epithelium (Section 9,3.1).   Damage to the epithelium overlying
sensory receptors may be responsible  for  the increased  receptor reactivity to
mechanical stimulation (increased  ventilation  with exercise, low Op, or high
COp) or chemical  (histamine)  stimulation  (Nadel, 1977;  Boushey et a!,,  1980).
The rapid, shallow breathing observed in guinea pigs during 03 exposures (Amdur
et al.  1978;  Yokoyama,  1969;  Murphy et al., 1964) may also be related to in-
creased sensory neural activity coming from the lungs, not to an indirect effect
of changes in airway diameter or lung distensibility as previously speculated.
     In their  study  of allergic  lung sensitization,  Osebold et al.   (1980)
showed additional functional  evidence for epithelial  disruption caused by 03
exposure.   These  investigators  studied the anaphylactic response of mice to
intravenous ovalbumin injection following repeated inhalations of aerosolized
                                                              3
ovalbumin.  Mice were continuously exposed to 980 or 1568 |jg/m  (0.5  or 0.8 ppm)
of 0~ for four periods of 3 to 5 days each, separated by 3 to 8 days  of ambient
air exposure. During periods of 03 exposure, mice were removed from the exposure
chambers for short  periods,  and they  inhaled  ovalbumin aerosols for 30 min.
Mice exposed to  03  and ovalbumin aerosols developed more severe anaphylactic
reactions and had a higher incidence  of  fatal  anaphylaxis  than air-exposed
mice receiving the same exposure to aerosolized ovalbumin.   In mice exposed to
                                                                   3
aerosolized ovalbumin, 34 percent of the 03~exposed mice (1568 ug/m  ; 0.8 ppm)
developed fatal  anaphylaxis following intravenous ovalbumin injection, compared
with  16 percent  of  the  air-exposed animals.   This study also  showed some
indication of an interaction  between 0-  exposures  and  exposures to  sulfuric
                                                3                            3
acid aerosols.   Of  the mice exposed to 980 pg/m  (0.5 ppm) of 03 plus 1 mg/m
of  sulfuric  acid aerosols, 55 percent died of anaphylactic  shock following
intravenous ovalbumin injection,  compared with 20-percent mortality  in mice
exposed to 0., alone and zero mortality in mice exposed to sulfuric acid aerosols
alone.  The  authors propose  that these data may indicate that pollutants can
increase not only the total  number of  clinical asthma  attacks, but  also the
number  of allergically  sensitized individuals  in  the population.  Matsumura
                                   9-72

-------
(1970) observed  a similar  increase  in the allergic  response of sensitized
                                                                           2
guinea pigs  to inhaled antigen  following 30 min of  exposure to 3920 ug/m
(2.0 ppm) of 03.

9.3.3  Biochemically Detected Effects
9.3.3.1   Introduction.  This section  will  include  some studies involving con-
centrations  above 1960 ug/m  (1 ppm) of  0~,  because the direction  of the
effect is opposite that for lower 0, concentrations (decrease vs. increase in
many  parameters).   An extensive body  of  data  on this topic  has been reviewed
by Menzel (1983), Mustafa  et al.  (1977,  1980,  1983),  Mustafa and Lee (1979),
Cross  et al. (1976), and  Chow  (1983).  To facilitate presentation of this
information,  it  has been categorized  by  broad classes of metabolic activity.
This  results in   some degree of  artificial separation,  particularly because
many  patterns  of response to 0"3  are similar across the classes of metabolism.
Lung  permeability is discussed in  this section, because it is typically detected
biochemically.   The final   subsection  presents  hypotheses about the molecular
mechanism(s) of action  of 0~, relying  on  the data presented  earlier by metabolic
class.
9.3.3.2   Antioxidant Metabolism.   Antioxidant  metabolism of  the  lungs  is
influenced by  03 exposure.  As  shown  in  the schematic diagram below  (Figure
9-3),  this  system consists of a number of enzymes.  As shall be  discussed, 03
has been shown to produce several reactive oxidant species jm vitro from  com-

                 -^—G-6-P——^v—-NADP+-*->«*•** GSH-
                          G-6-PD         GSH       GSH
                        HMP shunt     reductase   peroxidase
                                                         ROH
         Figure 9-3. Intracellular compounds active in antioxidant metabolism of
         the lung. (G-6-P = glucose-6-phosphate; 6-PG = 6-phosphogluconate;
         G-6-PD = glucose-6-phosphate dehydrogenase; HMP shunt = hexose
         monophosphate shunt; NADP+ = nicotinamide adenine dinucleotide
         phosphate; NADPH = reduced NADP; GSH = glutathicne; GSSG =
         glutathione disulfide; [O] = oxidizing moiety [i.e., hydrogen peroxide,
         free radical, lipid peroxide]; GSH peroxidase = glutathione perioxidase;
         GSH reductase = glutathione reductase; and ROH = reduced form of
         [O]).
         Source: U.S. Environmental Protection Agency (1978).

                                    9-73

-------
pounds found in the lung, as well as in other organs.  It is reasonably certain
that Og can produce such  reactive species  in the  lung after jji vivo exposure.
Many of these  oxidant species are metabolized by the glutathione peroxidase
system, rendering  them less  toxic.   Thus, this  system  is  involved  in the
toxicology of On.
                                                                   3
     Typically, following exposures to  levels of  (U  below 1960 ug/m  (1 ppm),
the activities  of most  enzymes  in  this system are  increased  (Table 9-5).
Whether this increase is  due to direct mechanisms  (e.g.,  de noyo synthesis
resulting in greater enzyme activity), indirect mechanisms (e.g., an increased
number of type 2 cells that naturally have a higher enzymic activity than type
1 cells), or a combination  of both has  not been proven.  The increase  in type
2 cells is  most  likely to  be the mechanism,  because with similar exposure
regimens,  the effects (increased enzyme activities  and  increased numbers  of
type 2 cells)  increase,  reaching a maximum at 3  to  4 days of exposure and a
steady state  on day  7 of exposure.   Whatever the  mechanism,  the increase
occurs at low  0^ levels in several species  under varying exposure regimens
(Table 9-5).  The  net result is that the  antioxidant metabolism of the lung
is  increased.   Whether this  is  a  protective or  a toxic response  is often
debated.   To resolve this debate scientifically will require more knowledge of
the mechanisms involved.  However, even attributing  it to be a protective  re-
sponse implies a physiological  need for protection  (e.g.,  an  initial  toxic
response occurred  which  required protection).   A more detailed discussion of
these effects follows.
     Acute  exposures  to high concentrations of 0^  generally decrease  anti-
oxidant metabolism, whereas repeated exposures to  low  levels  increase this
metabolism.   For example, DeLucia et al. (1975a)  compared the effects of acute
(2 to 8 hr) exposures to high 03 levels (3920 and 7840 H9/m3, 2 and 4 ppm) and
short-term  (8 .or 24 hr/day, 7 days)  exposures to lower  Q~  levels (392, 980,
         3
1568  ug/m  ,  0.2, 0.5, 0.8  ppm)  on  rats.   For'nohproteTn sulfhydryllevels
(principally glutathione, GSH), decreases  in the  level of GSH were progressive
                                                3
with time of  exposure (2 to  6 hr)  to 7840 |jg/m  (4 ppm).   For  glutathione
disulfide (6SSG),  decreases were less and had returned  to  normal by 6 hr  of
exposure.   These exposure regimens also decreased the activities of GSH reduc-
tase and glucose-6-phosphate dehydrogenase (G-6-PD).  After the first day  of a
                                      q
7-day continuous exposure to 1568 pg/m  (0.8 ppm) of 03, no significant change
was seen in the nonprotein sulfhydryl or GSH content or in the activities of
G-6-PD, GSH reductase, or disulfide reductase.  However, the levels/ activities
                                   9-74

-------
                                      TABLE 9-5.  CHANGES IN THE LUNG ANTIOXIDANT METABOLISM AND OXYGEN CONSUMPTION BY OZONE
f
-J
cn
Ozone
concentration
ug/m3 ppm
196 0.1
196 0.1
392 0.2
196 0.1
392 0.2
392 0. 2
980 0. 5
1568 0.8
1568 0.8
. Exposure
Measurement ' duration and
method protocol
NBKI Continuous
. ' for 7 days
I Continuous
for 7 days
MAST, Continuous
NBKI for 7 days
I Continuous
for 7 days; or
8 hr/day for
7 days
Continuous
for 1 to 30
days
Observed effect(s)c Species Reference
With vit E-deficient diet, increased levels of GSH and Rat Chow et a!., 1981
activities of GSH peroxidase, GSH reductase, and G-6-PD;
no effect on malic dehydrogenase. With 11 ppm vit E
diet, increased levels of GSH peroxidase and G-6-PD.
With 110 ppm vit E diet, no change.
With 66 ppm vit E- supplemented diet, increase in Rat Mustafa, 1975
oxygen consumption of lung homogenates only at Mustafa and Lee,
0.2 ppm. With 11 ppm vit E-supplemented diet, 1976
increase in 02 consumption at 0.1 and 0.2 ppm.
Increase due to increased amount of mitochondria
in lungs.
Increased activities of GSH peroxidase, GSH reduc- Rat Plopper et al.,
tase, and G-6-PD and of NPSH levels with (66 mg/kg) 1979
or without (11 mg/kg) vit E supplementation; at
0.2 ppm, effects less with vit E supplementation.
Morphological lesions unaffected by vit E supple-
mentation.
For the continuous exposure to the two higher con- Rat Mustafa and
centrations, increased activities GSH peroxidase, Lee, 1976
GSH reductase, and G-6-PD. At the lower concen-
tration (continuous), increased activities of GSH
peroxidase and GSH reductase. A linear concentration-
related increase in all three enzyme activities. In-
creased 02 consumption using succlnate-cytochrome C
reductase activity fairly proportional to 03 level.
Similar results for intermittent exposure groups.
Increased rates of 02 consumption, reaching a peak
at day 4 and remaining at a plateau for the re-
mainder of the 30 days. Also an initial decrease
(day 1) and a subsequent increase (day 2) in
activity of succinate-cytochrome C reductase
which plateaued between days 3 to 7.

-------
                              TABLE 9-5.  CHANGES IN THE LUNG AHTIOXIDANT HCTABOLISM AND OXYGEH CONSUMPTION  BY OZONE (continued)
f
Ozone

concentration
pg/a3
392
686
980
1568



392
980
1568




392
980
1568


392
980
1960

392
980
1960

2352-
16,072


ppm
0,2
0.35
0.5
0.8



0.2
0.5
0.8




0,2
0.5
0.8


0.2
0.5,
1.0

0.2
0.5
1.0

1.2-
8.2


b Exposure
Measurement ' duration and
method protocol
I 8 hr/day for
7 days





N8KI Continuous
for 8 days or
8 hr/day for
7 days



HAST, 8 or 24 hr/day
NBKI for 7 consecu-
tive days


NBKI 3 hr/day for
4 days


NO 4 hr/day for
up to 30 days


4 hr





Observed effect(s)c Species
Increased concentration-related activities of Monkey,
G-6-PD, NADPH-cytochrome c reductase, and
succinate oxidase. Significant increases
occurred in the bonnet monkey at 0.35 and
0.5 ppm; in the rhesus monkey at 0.8 ppn.
However, actual data were only reported for
succinate oxidase.
For continuous exposure to two higher concentrations, Rat
increased activities of GSH peroxidase, GSH reduc-
tase, and G-6-PD. At the lower concentration (con-
tinuous), increased activities of GSH peroxidase
and GSH reductase. A concentration-related linear
Increase in all three enzyme activities. Similar
results obtained for intermittent exposure groups.
Activities of G-6-PD and NADPH-cytochrome C reduc- Rat
tase and succinate oxidase increase- in a concen-
tration-dependent fashion. No significant
differences between the intermittent and continuous
exposure groups.
Reduced glutathione levels increased in a linear House
concentration-dependent manner. No effect at
0.2 ppm in the no-exercise. group. Exercise en-
hanced effect.
GSH content increased directly with 03 concentration Mouse
and exposure duration. Increase in activities of
G-6-PD, GSH reductase, and GSH peroxidase after 7
days of exposure to 0.5 and 1 ppm.
Decrease 1n GSH content after exposure to 8.2 ppm.
No change below 4.0 ppm. Two days postexposure to
4 ppm, GSH content increased, lasting for several
days.


Reference
Mustafa and
Lee, 1976





Chow et al . , 1974






Schwartz et al . ,
1976



Fukase et al . , 1978



Fukase et al., 1975








-------
                         TABLE 9-5.   CHANGES IN THE LUNG ANTIOXIOANT METABOLISM AND  OXYGEN  CONSUMPTION BY OZONE (continued)
Ozone
concentration
M9/m3
392
980
1568
1568
3920
7840


392
980
15S8
627
882
ppm
0.2
0.5
0.8
0.8
2
4


0.2
0.5
0.8
0.32
0.45
b Exposure
Measurement ' duration and
method protocol
MAST, 8 or 24 hr/day
NBKI for 7 days
8 hr/day
for 7 days
2 to 8 hr


I Continuous
for 7 days
UV 6 hr
UV Continuous
for 5 days
Observed effect(s)c
All 03 levels: increase in NPSH levels; increased
activities of G-6-PD, GSH reductase, NADH cyt. c
reductase. At 0.5 and 0,8 ppm, increased activity
of succinate cyt. c reductase. At 0.8 ppm, continuous
increase began at day 2 of exposure.
Increased NPSH, GSH, and G-6-PD; no change in other
enzymes.
Loss of GSH; loss of SH from lung
nitochrondrial and microsomal frac-
tions and inhibition of marker
enzyme activities from these
fractions.
Concentration- related increase in 02 consumption.
In both vit E-supplemented and nonsupplemented
groups: increased G-6-PD activities and GSH
levels; decreased ACHase activities.
Mice: increased levels/activities of TSH, NPSH,
GSH peroxidase, GSH reductase, G-6-PD, 6-P-GD,
isocitrate dehydrogenase, cytochrome c oxidase,
Species Reference
Rat DeLucia et al. ,
1975a
Monkey
Rat


Rat Mustafa et al . ,
1973
House Moore et al . , 1980
Mouse, Mustafa et al.,
3 strains 1982
of rats
                                                          and  succinate oxidase,
                                                          Rats:   increased  levels/activities  of NPSH,  GSH
                                                          peroxidase,  and G-6-PD  in  several strains. Gene-
                                                          rally  mice were more  responsive.  For both species,
                                                          no change in DNA  or protein  levels  or activity  of
                                                          GSH-S-transferase.
882
0.45
                       UV
                                  8 hr/day         Oa and 4.8 ppin of N0a alone produced  no  significant
                                  for 7 days        effects but 03 + NOZ produced  synergistic  effects:
                                                   increased total and nonprotein sulfhydryls;  increased
                                                   activities of succinate oxidase and cytochrome  c
                                                                                                                     Mouse
Mustafa et al.,
1984

-------
     PRELIMINARY DRAFT




TABLE 9-5,  CHANGES IH THE LUNG AMTIOXIDAHT METABOLISM AND OXYGEN  CONSUMPTION BY OZONE (continued)
Ozone
concentration
pg/«3 ppi
882 0.45
980 0.5
1372 0.7
1568 0.8
U3 1470 0.75
-J
CO-
1S68 0.8
1568 0.8
1568 0.8
3920 2
u Exposure
Measurement « duration and
Method protocol
ND Continuous
for 7 days
NBKI 8 hr/day for
7 days
NBKI Continuous
for 5 days
Continuous
for 7 days
NBKI Continuous
for up to
30 days
Conti nuous
for 7 days
NBKI Continuous for
3 days
I Continuous for
10 to 20 days
8 hr
Observed effect(s) Species Reference
Increased SOD activity at days 3 and 5, but not Rat Bhatnagar et al.,
days 2 and 7 of exposure. 1983
Increases in activities of GSH peroxidase, GSH Rat Chow et al., 1975
reductase and G-6-PD and In GSH levels of rats. Monkey
No effect In iionkeys.
Increased activities of GSH peroxidase, G-6-PD, Rat Chow and Tappel,
and GSH reductase. Malonaldehyde observed. 1972
The increases in the first two enzymes were partially
inhibited as a logarithmic function of vitanin E
levels in diet.
Increase in activities of GSH peroxidase, GSH Rat Chow and Tappel,
reductase, G-6-PD, 6-P-GQ, and pyruvate kinase 1973
at day 3, reaching a peak at day 10, at which
time beginning of a slight decrease (except for
GSH peroxidase which continued to increase). At
day 30 still elevated over controls.
Increased activities of hexose monophosphate shunt
and glycolytic enzymes of lung.
Increased activities of GSH peroxidase, GSH reductase, Rat Chow et al., 1976b
and G-6-PD; levels of NPSH; general protein synthesis;
and rate of mitochondria! succinate oxidation.
Decrease to control values 6 to 9 days postexposure.
Re-exposure using same reginen (S, 13, or 27 days
postexposure) resulted in similar elevations.
At the higher concentration: increase in the lung Rat Mustafa et al.,
mitochondria! 03 consumption in oxidation of 2- 1973
oxyglutarate and glycerol-1-phosphate and the number
of type 2 alveolar cells which are rich in Mito-
chondria. No change in malonaldeliyde. At the
lower concentration: increase in 02 consumption.

-------
                          TABLE 9-5.   CHANGES  IN THE  LUNG ANTIOXIDANT METABOLISM AND OXYGEN  CONSUMPTION BY OZONE (continued)
    Ozone
concentration
jig/ni3     ppm
1568
3920
0.8
                      Mea'surement'
                        method
                                 a,b
                                    Exposure
                                  duration  and
                                    protocol
Observed effect(s)
Species
Reference
1568

2940-
7840
0.8 I;

1.5-
4
< 24 hr

< 8 hr

NPSH level unaffected at 0.8 ppm of 03 for < 24 hr or Rat
or 1.5 ppm for < 8 hr; decreased at 2 ppm of 03 for
8 hr or 4 ppm for 6 hr. At 4 ppm of 03 for 6 hr,
decreased level of GSH; no change in GSSG level.
De Lucia et
1975b


al.,



                                  10 days          Lung  SH  levels  unchanged.   Increase  in  G-6-PD and
                                                   and cytochrome  c  reductase  activities.   No change
                                                   in malonaldehyde  levels.

                                  4 to  8 hr       Decrease in  lung  SH  levels  and in  G-6-PD,  GSH
                                                   reductase, and  cytochrome c reductase activi-
                                                   ties.  No change  in  malonaldehyde  levels.
                                                          High mortality in 7- and 12-day-old  rats.
                                                                                                                        Rat
                                                          DeLucia et al.,
                                                          1972
3920-
5880
1568
1568
2-
3
0.8 NO
0.8 UV
30 m1n
Continuous
for 7 days
72 hr
In vitro; decrease in SH levels; increase in
•alonaldehyde levels.
Increased activity of superoxide dismutase.
Succinate oxidase, cytochrome c oxidase, and iso-
citrate dehydrogenase: No effect at 24 days old,
increased in 90-day-old rats. G-6-PD, 6-PGH:
increased at 24 and 90 days of age, latter had
greater increase. Succinate oxidase and G-6-PD
decreased in 7- and 12-day-old rats and increased
in 18-day-old rats.

Rat
Rat
(7 to 90
days old)

Mustafa et al . ,
1977
Elsayed et al. ,
1982a
1568    0.8
                        UV
                                          Continuous       Diet was constant vitamin E and deficient  or  sup-
                                          for  5 days       plsmented (2 levels) with selenium  (Se).   No  change
                                                          in GSH peroxidase.  With 0 Se, decreased GSH  reduc-
                                                          tase activity; no change with low or  high  Se.
                                                          Progressive increase in activities  of G-6-PD  and
                                                          6-P-GD with increasing Se, beginning  at low Se  level.
                                                                                                                Mouse
                                                          Elsayed et al.,
                                                          1982b

-------
                               TABLE 9-5.  CHANGES IN THE LUNG ANTOXIDANT METABOLISM AND OXYGEN CONSUMPTION BY OZONE (continued)
 vo
•o
Ozone
concentration
ug/m3 ppm
1568 0.8
1568 O.B
k Exposure
Measurement * duration and
nethod protocol
UV Continuous
for 5 days
NBKI continuous
for 7 days
Observed effect(s)c Species
Diet was constant vitamin E and 0 ppm of Se or 1 ppm Mouse
in Se. +Se: increased G-6-PD, 6-P-GD; no change
in GSH reductase or GSH peroxidase. -Se: decreased
GSH reductase.
Both diet groups had increase in TSH and NP.SH, and
lung Se levels after 03.
Vitamin E partially prevented increased activities Rat
of G-6-PD, 6-P-GD, and malic enzyme. Activities
of phosphofructokinase and pyruvate kinase increased.
No effect on aldolase and nalate dehydrogenase.
Reference
El sayed et al . ,
1983
Chow and Tappel ,
1973
      1764    0.9
    CHEM
96 hr
Trend towards decreased activities of GSH reductase
GSH peroxidase, G-6-PD before 18 days of age, followed
by increases thereafter.   For G-6-PD:  no change at 5
and 10 days of age; decrease at 15 days, and increase
at 25 and 35 days.
Rat          Tyson et al.,
(5-180       1982
days old)
     1764    0.9
    MAST
>, 96 hr          96 hr:   No effect below 20 days of age;  G-6-PD in-
                 creases thereafter up to 35 days, after  which (40
                 and 50 days old) 1t decreases.   When exposure started
                 at 25 or 32 (but not 10 to 15)  days of age, the maxi-
                 mal increase in G-6-PD occurred at about 32 days of
                 age under continuous exposure conditions.
                                                             Rat          Lunan et al.,
                                                             (10-50       1977
                                                             days old)
    Measurement method:
MAST = Kl-coulonetric (Mast meter);  NBKI = neutral  buffered potassium iodide;  CHEM = gas solid cheniluminescence;  UV = UV photometry;
I =;iodometric; NO = not described.
    Calibration method:  NBKI = neutral buffered potassium iodide.

   cAbbreviations used:  GSH = glutathione; GSSG = reduced glutathione; G-6-PD = glucose-6-phosphate dehydrogenase; LDH = lactate dehydrogenase;
    NPSH = non-protein sulfhydryls; SH = sulfhydryls; 6-P-GD = 6-phosphogluconate dehydrogenase.

-------
of these constituents increased by day 2 and remained elevated for the remainder
of the exposure  period.   Comparable results were  reported with similar (but
not identical) exposure  regimens  by DeLucia et al.  (1972, 1975b)  using rats
and Fukase et al. (1975) using mice.
     Investigators (Table 9-5)  have found that for  lower  levels  of 0~, in-
creases in antioxidant  metabolism are linearly related to 0,  concentration.
Most such studies were conducted by using intermittent and continuous exposures.
No differences between these regimens were found,  suggesting that concentration
of exposure is more important than time of exposure.
     Chow et al.  (1974) exposed rats continuously or intermittently (8 hr/day)
                                    3
for 7 days to 392, 980, or 1568 \ig/m  (0.2, 0.5, or 0.8 ppm) of 0- and found a
concentration-related  linear  increase in activities of GSH  peroxidase, GSH
reductase, and G-6-PD.   Significant increases occurred for all measurements,
                                                 3
except G-6-PD  at continuous  exposure to 392  |jg/m   (0.2  ppm).   Although the
difference between continuous  and intermittent  exposure was  not examined  sta-
tistically, no major  differences  appeared to exist.   Schwartz et  al.  (1976)
made similar  observations  for G-6-PD activity when  using  identical  exposure
regimens  and  found concurrent  morphological  changes  (Section 9.3.1).   Mustafa
and Lee (1976), also by using identical exposure regimens, found similar effects
for G-6-PD activity.   DeLucia et al. (1975a) found similar changes and increased
nonprotein sulfhydryls at all three concentrations of 0^.   A similar study was
performed in mice by using a longer exposure period of 30 days (Fukase et al.,
1975).   The  increase  in GSH level  was  related  to  concentration and time of
exposure.  Fukase et  al.  (1978) also observed a linear concentration-related
increase  in GSH  levels of mouse lungs exposed 3 hr/day for 4 days to 392, 980,
            3
or 1960 |jg/m   (0.2, 0.5, or 1.0 ppm) of 0,.  Exercise enhanced the effect-  At  ..
the lower 0- level, the increase  in GSH was significant only in the exercising
mice.
     The  influence  of  time  of exposure was  examined  directly by Chow and
-Tappel (1973).   Rats'were exposed  continuously to  1470 |jg/m  (0.75 ppm) for 1,
3, 10,  or 30 days,  at which times  measurements of GSH reductase, GSH peroxi-
dase, G-6-PD,  pyruvate kinase, and  6-phosphogluconate dehydrogenase activities
were made.   No  statistical  tests  or  indications  of data variability were
presented.  A  few of the enzyme activities (GSH peroxidase and 6-phosphogluconate
dehydrogenase) may have decreased  at day 1 of exposure.  All enzyme activities
except GSH  peroxidase  increased by day  3  and reached a pe-ak  at 10 days and
then began to  return toward control values.  GSH peroxidase activity continued
                                    9-81

-------
to increase over this time of exposure.   In a similar study  (Mustafa and  Lee,
1976), G-6-PD activity was measured after rats were exposed for 7 days to 1568
    3
ug/m  (0.8 ppm)  continuously.   No effect was detected on day 1, but by day 2
the activity  had  increased.   The peak response was  on  day 4; the activity
remained elevated to an equivalent degree on day 7.  In  a  similar experiment,
DeLucia et al. (1975a) obtained equivalent results.
     The tolerance phenomenon has also been investigated for lung antioxidant
                                                                   3
metabolism.  Rats were exposed continuously for 3 days to 1568 ug/m  (0.8 ppm)
of Og,  allowed to remain  unexposed for 6, 13, or 27  days,  and then re-exposed
for 3  days to the same 03  level (Chow et al.,  1976b).   Immediately after
the first 3 days of exposure, the activities of GSH peroxidase, GSH reductase,
and G-6-PD were  increased,  as was the nonprotein  sulfhydryl content.   By 2
days  after this exposure ceased,  recovery had begun;  control values were
completely reached by 9 days postexposure.  Following a 30-day recovery period,
no changes were observed.   When re-exposure commenced on day 6 of recovery (at
which time incomplete recovery was observed), the metabolic activities returned
to levels  equivalent  to  those of the original  exposures.   Similar findings
were made  when  re-exposure  commenced on  days 13 and 27  days of  the  recovery
period.
     The  influence  of vitamin  E,  an antioxidant,  on 03 toxicity  has been
extensively studied, because it typically reduces the toxicity of 03 in animals.
This topic has  been recently reviewed by  Chow (1983).   Early studies  centered
on mortality.   For  example,  vitamin E-deficient rats are more susceptible to
                                 3
continuous exposure to  1960  vg/m  (1 ppm) of 03 than rats fed supplements of  ,
vitamin E  (LT50,  the  time at which  a 50  percent mortality is observed,   8.2
days  versus  18.5 days)  (Roehm  et al., 1971a,  1972).   Vitamin E protected
animals from mortality and changes in the wet to dry weight  ratios of the lung
                                                 o
(lung edema)  on  continuous  exposure to 1568 ug/m   (0.8 ppm) of  0- or  higher
for 7  days (Fletcher  and Tappel, 1973).  Vitamin  E  protection against 03 is
positively correlated to the log concentration of dietary vitamin E fed to the
rats.   Rats maintained  on vitamin E-supplemented  diets  and exposed  to 1568
    o
MS/"1   (0-8 ppm)  of  0- continuously for 7 days also  had changes  in 6-phospho-
                                                                            3
gluconate  dehydrogenase activity.  Rats  were exposed to  1372 to  31,360 jjg/m
(0.7 to 16 ppm) of 03 while being  fed  diets containing ascorbic  acid,  dl-
methionine, and  butylated hydroxytoluene (Fletcher  and  Tappel,  1973).   This
combination was  supposed  to  be  a more potent  antioxidant mixture than vitamin
E alone.  Animals fed diets  with the highest level of this antioxidant mixture
                                   9-82

-------
had the greatest  survival  rate.   Animals fed ortocopherol (vitamin E) in the
range of 10 to 150 mg/kg of diet had a survival rate slightly lower than those
fed the combination of antioxidants.
     Donovan et al.  (1977)  fed mice 0 (deficient diet), 10.5 (minimal diet),
or 105  (supplemental  diet)  mg/kg of vitamin  E acetate.   The diet was also
altered to  increase  the peroxidilability of the  lung by feeding either low or
high polyunsaturated fats (PUFA).  Mice were continuously exposed to 1960
    3
|jg/m  (1 ppm)  of  0^.  The mortality  (LT50 of 29 to 32 days) was  the same,
regardless of the large differences in peroxidizability of the lungs of animals
fed high- or low-PUFA diets.  High supplemental levels (105 mg/kg) of vitamin E
acetate were protective and  delayed the LT50 to 03 by an average of 15 days.
Although these experiments demonstrate clearly the protective effect of vitamin
E against  Og  toxicity,  they do  not support  the hypothesis that changes  in
fatty acid  composition  of the lung will increase  Oj  toxicity.   The results
could be  interpreted to indicate that the scavenging of radicals by vitamin  E
is more important than the relative rate of oxidation of PUFA.  These findings
led  to  biochemical  studies  that used graded  levels of dietary  vitamin E.
     Plopper et al.  (1979)  correlated biochemical and morphological (Section
9.3.1)  effects  in rats maintained on  a synthetic diet with 11 mg kg/vitamin  E
(equivalent to the average U.S. adult  intake) or commercial rat chow having 66
mg/kg vitamin  E.   The 11  mg/kg vitamin E group was  exposed continuously  for  7
                       3
days to 196 or 392 M0/m  (0.1 or 0.2 ppm) of 0.,, and measurements were made at
the end of exposure.  The rats on the  commercial diet were exposed to only the
higher  concentration.  All  exposures   increased  activities of GSH peroxidase,
GSH  reductase,  and G-6-PD, and the amount of  nonprotein sulfhydryl.   Although
statistical  comparisons  between  the  dietary  groups were   not made, greater
increases appear  to  have occurred  in the 11 mg/kg vitamin  E group; the magnitude
                                                            3
of  the  responses  in the higher  vitamin  E group at  392  |jg/m  (0.2  ppm)  of 0™
was  roughly  equivalent to the magnitude of the responses  of the  low vitamin  E
                          3
group exposed  to  196 (jg/m   (0.1  ppm)..  The  2 dietary groups showed  little
variation  in morphological effects.
     These  studies  were  expanded to  include  three  vitamin E dietary  groups:
                                                                      3
0,  11,  or 110 ppm  (Chow  et  al., 1981).  Rats were  exposed to 196  yg/m   (0.1
ppm) of 03 continuously for  7  days.   In the 0-ppm vitamin  E group, 03 increased
the  level  of GSH and the  activities  of GSH  peroxidase, GSH  reductase,  and
G-6-PD.   Increases  of  similar magnitude occurred  in  the 11-ppm vitamin E
group,  with the exception of  GSH  reductase activity, which was not affected.
                                    9-83

-------
In the highest  vitamin E group, no significant effects were observed.   Ozone
caused no changes in the activity of malic dehydrogenase in any of the dietary
groups.   Morphologically (Section  9.3.1}, only 1 of  6  rats of the 110-ppm
vitamin E group had lesions,  whereas more rats of  the two other groups had
lesions.   These  lesions  became  more severe as the vitamin E level decreased.
They occurred  at the  bronchio-alveolar  junction  and  were characterized by
disarrangement  of the  bronchiolar epithelium  and an increase in the number of
alveolar macrophages.
                                                                  3
     Chow and Tappel (1972) exposed  rats  continuously to 1372 ug/m  (0.7 ppm)
of 0., for 5  days.   The  animals  had  been maintained on diets with different
levels of dl-crtocopherol  acetate  (vitamin E) (0,  10.5,  45,  150, and 1500
mg/kg diet).   Ozone exposure increased  GSH  peroxidase,'GSH reductase, and
G-6-PD activities.   For GSH peroxidase and G-6-PD activities, the increase was
reduced as a function  of the  logarithmic  concentration of  vitamin E.  Vitamin
E did not alter the magnitude of the  effect  on  GSH reductase, a finding in
contrast to the results  of others (Chow et al.,  1981; Plopper et al., 1979).
Malonaldehyde,  which is produced by lipid peroxidation, increased; this increase
was also partially inhibited as a logarithmic function of vitamin E concentra-
tion.  However,  others (DeLucia et al.s 1972; Mustafa et al.,  1973) have not
observed the  presence  of malonaldehyde  in exposed  lungs.   The increase in
malonaldehyde and activity of GSH peroxidase were linearly correlated, leading
Chow and Tappel  (1972) to propose a compensatory mechanism in which the in-
crease in GSH peroxidase activity increases lipid peroxide catabolism.
     Chow and Tappel (1973)  observed the typical protection of vitamin E (0
                                                                3
and 45 mg/kg diet crtocopherol)  from the  effect of 03  (1568 ug/m  > 0.8 ppm;  7
days, continuous) on increasing G-6-PD activity in rat lungs.   Similar findings
occurred for  6-phosphogluconate dehydrogenase and  malic  enzyme activities.
The activities  of two  glycolytic regulating enzymes,  phosphofructokinase and
pyruvate kinase, were  increased by 0~ exposure but were  not influenced by
vitamin E levels in the diet.  Aldolase  and  malate dehydrogenase activities
were not affected.
     Elsayed et al.  (1982b,  1983)  examined the influence of selenium (Se) in
the diet on GSH peroxidase activity in the lung.   Selenium is an integral part
of one form of the enzyme GSH peroxidase.  Mice were raised on a diet contain-
ing 55 ppm vitamin E with either 0 ppm or 1 ppm of Se and exposed to 1568 ± 98
    3
fjg/fli  (0.8 ppm) of  0-  continuously for 5 days.  In these mice, Se deficiency

                                   9-84

-------
caused a sevenfold decline  in Se  level and a threefold decline  in GSH peroxi-
dase activity  in the lung.  Other  enzyme  activities (e.g., GSH reductase,
G-6-PD, 6-phosphogluconate  dehydrogenase)  were not  affected  by dietary Se.
After  03  exposure,  the  GSH peroxidase activity  in  the  Se-deficient group
remained unstimulated and  was  associated with a  lack  of stimulation of GSH
reductase, G-6-PD, and  6-phosphogluconate  dehydrogenase  activities.  In con-
trast, the 0--exposed Se-supplemented group exhibited increases in 6-phosphoglu-
conate dehydrogenase and G-6-PD activities.   Dietary deficiency or supplementa-
tion of Se, vis-a-vis alteration of GSH peroxidase activity, did not appear to
influence the effects of Q3 exposure as assessed by other parameters.  Although
the animals  received  the same level of  dietary  vitamin  E, after air or 0«
exposure, the  Se-deficient  group  showed  a two-fold  increase in  lung vitamin E
levels  relative  to the  Se-supplemented  group,  suggesting a complementary
relationship between Se and vitamin E in the lung.  This sparing action between
Se (i.e., GSH peroxidase activity) and vitamin E might explain similar effects
of 03  exposure in Se-deficient and supplemented mice.
     Several investigators  have studied the responsiveness of different species
to the effect of 0- on antioxidant metabolism.  DeLucia et al. (1975a) exposed
                                                                 3
both Rhesus  monkeys  and rats for 7 days (8  hr/day)  to 1568 ug/m  (0.8  ppm).
The nonprotein sulfhydryl and GSH content were increased, as was G-6-PD activity.
Activity of  GSH  reductase  was affected  in the rats  but  not the monkeys.   No
statistical  comparisons  were  made between the rats and monkeys.  In the only
parameter for which sufficient data were presented for comparison, G-6-PD, the
increase in  monkeys was  about 125 percent of controls; for rats,  it was about
130 percent  of controls.
     Rats  and  Rhesus  monkeys were compared more  extensively by Chow et al.
                                          3
(1975).  Animals  were exposed to 980 jjg/m   (0.5 ppm)  of 03  8  hr/day for  7
days.  The nonprotein sulfhydryl  content and the activities of GSH peroxidase,
GSH reductase, and G-6-PD increased in rats but not  in monkeys.  The magnitude
of  the increases in  rats was  20  to  26  percent.   The increases in monkeys were
between 10 and 15 percent and statistically insignificant,  "because of  relative-
ly  large  variations," according  to  the authors.   The variation  in  the monkeys
was approximately double that of  the rats.  The sample size of  the monkeys  (6)
was  lower  than that of  the rats  (8).  Statistical tests of the Type II error
(e.g.,  false negative  error)  rates were not  reported.   Thus,  the monkeys
apparently were  not  affected to  the same  degree as the rats.  However, the
experiments  with monkeys were apparently not conducted with as  much  statistical
                                   9-85

-------
power as those  with rats.   Thus, under the actual study designs used, ozone
would have had  to  have  substantially greater effects on monkeys than rats for
a statistically significant effect to  be detected.   This  did  not happen,
leading to  the conclusion  of  the investigators that monkeys are  not  more
responsive.   Studies  of improved  experimental  design would  indicate  more
definitively whether monkeys are less responsive.   Mustafa and Lee (1976) also
alluded to different  G-6-PD responses  of rats  and Bonnet and Rhesus monkeys
                                                                         Q
after exposures for 8  hr/day  for 7  days to levels as  low as  392 Mfl/m
(0.2 ppm).   However, no data for G-6-PD were presented,  and the  description  of
these results was incomplete.
     Mice (Swiss Webster)  and  3 strains of rats (Sprague-Dawley,  Wistar, and
                                                                       3
Long Evans)  were-compared after a 5-day continuous exposure to 882 pg/m  (0.45
ppm) of OQ (Mustafa et a!., 1982).  Total  sulfhydryl  content increased only  in
mice.  However,  nonprotein  sulfhydryl content increased in  both rats and mice
to a roughly equivalent degree.   GSH-S-transferase was not  affected in any of
the animals.  Mice  exhibited the typical  increases in the  activities of GSH
peroxidase,   GSH reductase,  G-6-PD,  6-phosphogluconate dehydrogenase,  and
isocitrate dehydrogenase.   Rats  were  less affected;  no changes were seen in
the  activities  of GSH  reductase or isocitrate dehydrogenase,  and not all
strains of rats showed an increase in the activities  of GSH peroxidase, G-6-PD,
and  6-phosphogluconate  dehydrogenase.   For GSH reductase  and  G-6-PD,  the
increased activities  in exposed mice were significantly greater  than  those
in exposed rats.
     At present, it is not possible to determine whether these apparent species
differences in  responsiveness  were due to differences in the total deposited
dose of Og,  an innate difference in species  sensitivity,  or differences in
experimental design  (e.g.,  small  sample  sizes, insufficient concentration-
response studies).
     Age-dependent  responsiveness to Q-Hnduced changes in GSH systems  has
been observed.   Tyson  et  al. (1982) exposed rats (5  to 180 days old) to 1764
    a
ug/m  (0.9 ppm) of 03 continuously for 96 hr, except for suckling  neonates (5
to 20 days old) which received an intermittent exposure (4 hr of exposure,  1.5
hr no exposure, 4 hr exposure).  Given that  others  (Mustafa and Lee,  1976;
Chow et al., 1974; Schwartz et al., 1976) have observed no differences between
continuous and  intermittent exposures for these  enzymatic  activities,  this
difference in  regimen  can  be considered inconsequential.   All ages given are
ages at the  time of initiation of exposure.   They were calculated from those
                                   9-86

-------
given in the report to facilitate comparisons with other reports on age sensi-
tivity.   Weanlings  (25 and  35  days  old)  and  nursing  dams  (57  and 87 days old)
had higher  lung  to  body weight ratios.  Generally,  the DNA content of the
younger animals  was  unchanged.  When  the activities  of  G-6-PD,  GSH reductase,
and GSH peroxidase  were  measured after 0^ exposure, the trend was a decrease
in activities  at and before 18 days of age,  followed by increases thereafter.
For G-6-PD, this trend was most pronounced; at 5 and 10 days of age,  no signi-
ficant changes were seen;  at 15 days  of age,  a  decrease  was  seen; and at  25
and 35 days of age, progressively greater increases were seen.
                                                                   3
     Ten- to 50-day-old rats were exposed continuously to 1764 |jg/m  (0.9 ppm)
of 0- (Lunan et al», 1977).  Ages of rats reported are presumably ages at ini-
tiation of  exposure.   In rats (10  to 40 days old) exposed for 3 days, G-6-PD
activity was measured periodically during exposure.  No statistical  analyses
were reported.   Ozone caused a possible increase  (20 percent of control)  in
the activity of G-6-PD in the 20-day-old group, and the magnitude continued to
increase as age  increased up to about 35 days (~ 75 percent of control), after
which (40 and 50 days of age) the effect became less  (40 percent of control at
50 days  of age).  In another  experiment,  a  complex  design  was used  in which
rats at  10, 15,  25, and 32  days  of  age were  exposed up  to 32 to 34  days  of
age; thus,  the duration of  exposure  for each  group  was different.  When the
animals were younger than 20 days, .no effect  was  observed.   When older mice
were used,  the greatest magnitude  of  the increased activity occurred  at about
32 days of  age,  regardless  of  the absolute length  of  exposure.
                                                                       •2
     Elsayed et  al.  (1982a) exposed  rats  of various ages to 1568 \ig/m  (0.8
ppm) of  0~ continuously for 72  hr.  Ages  given are  those  at initiation of
exposure.   Ozone increased lung weights, total  lung protein, and total lung
DNA  in  an age-dependent fashion, with the older (90-day-old)  rats being more
affected  than  24-day-old animals.   For  isocitrate dehydrogenase, activity, no
effect was seen  in the  24-day-old  rats,,  but an increase was observed in the
90-day-old  animals.   For G-6-PD and 6-phosphogluconate  dehydrogenase, increases
were observed  in the 24- and 90-day-old  rats,  with a greater magnitude  of  the
effect  occurring in the 90-day-old group.   Younger  rats (7 to 18 days old)
were also examined.  The exposure  caused >60 percent mortality to the 7-  and
12-day-old  rats.  Glucose-6-phosphate  dehydrogenase  activity  decreased in  both
the  7-and-12-day old groups, with  the younger rats  being more affected.   The
18-day-old rats  had an  increase in this activity.   These trends were similar
                                    9-87

-------
to those  observed  by Tyson  et  al.  (1982),  although  the exact age for signifi-
cant changes differed slightly.
     The  reason  for  these age-dependent changes is  not  known.   The younger
animals (< 24 days of age} have lower basal levels of the studied enzymes than
the older animals  examined  (>  38 days of age)  (Tyson et  al., 1982;  Elsayed  et
al., 1982a).   It is  conceivable that age influenced the dosimetry of 03. The
decreased activities observed in the neonates are reminiscent of the decreased
activities that  occur  at higher 03 levels  in  adults (DeLucia et al.,  1972,
1975a; Fukase  et al.,  1975).   The  increased activities in the later stages  of
weanlings or  in  young,  growing adults is consistent with the effects observed
in other  studies of adult rats (Table 9-5).
     Mustafa et  al.  (1984)  are the only researchers to report the effects of
combined  exposures to  0- and nitrogen dioxide  on the GSH peroxidase system.
                                                         q
Mice were exposed 8 hr/day for 7 days to either 9024 ug/m  (4.8 ppm) of nitro-
                         3
gen dioxide,  or  882 ug/m (0.45 ppm) of 0~, or a mixture of these.  Ozone and
nitrogen  dioxide alone caused no significant effects on most of the endpoints;
however,  synergistic effects were  observed in  the  mixture group.   The total
sulfhydryl and nonprotein sulfhydryl contents  were  increased.  The  activities
of GSH peroxidase, G-6-PD,  6-phosphogluconate dehydrogenase,  and  isocitrate
dehydrogenase  increased,  but the activities of GSH  reductase and GSH S-trans-
ferase were  unchanged.   Tissue 0«  utilization  was also increased, as shown  by
the increase in succinate oxidate and cytochrome c oxidase activities.
     Superoxide  dismutase (SOD) catalyzes  the dismutation of (and  therefore
destroys) superoxide (02~), a  toxic oxidant species thought to be formed from
0-, exposure,  and is  thus involved in antioxidant metabolism.   Rats exposed
                                     3
continuously for 7 days to  1568 ug/m  (0.8 ppm) of 0., exhibited an increased
activity  of SOD in cytosolic and mitochondria! fractions of the lungs (Mustafa
et al., 1977).   In a more complex  exposure  regimen  in which rats were exposed
                        o
for 3 days  to 1568 pg/m  (0.8  ppm) and then various days of combinations of
2940 ug/m3 (1.5 ppm) and 5880 ug/m3 (3 ppm) of 03,  SOD activity also increased.
Bhatnager et al.  (1983) studied the time course of the increase in SOD  activity
                                             o
after continuous exposure of rats  to 882 ug/m   (0.45 ppm) of On.  On day 2  of
exposure, there  was  no effect.  On days 3  and 5, activity had increased; by
day 7 of  exposure, values were not different from control.
9.3.3.3.   Oxidative and Energy Metabolism.   Mitochondrial  enzyme activities
are typically studied to evaluate effects on 02 consumption,  which is a funda-
mental parameter of cellular metabolism.   Mitochondria are cellular organelles
                                   9-88

-------
that are the major sites of Op utilization and energy production.  Many of the
enzymes in mitochondria have  functional sulfhydryl groups, which are  known to
be affected by  0»,  and mitochondria! membranes  have  unsaturated fatty acids
that are also susceptible to CL.   The patterns of Ck effects on Op consumption,
as will be  discussed,  are quite similar to effects on antioxidant metabolism
(Section 9.3.3.2).
     Mustafa et al.  (1973) showed in rats that acute (8-hr) exposure to a high
                               3
concentration of  0^  (3920 |jg/m ,  2 ppm)  decreases  Op consumption using the
substrates  succinate,  a-oxoglutarate,  and  glycerol-1-phosphate.   Similar
findings were  made  by DeLucia et  al.  (1975a).   Decreases in  mitochondria!
total sulfhydryl levels were also observed (Mustafa et al., 1973).   Equivalent
changes occurred  in whole-lung homogenate and the mitochondria!  fraction.  No
change in malonaldehyde  levels  was found.  When rats were exposed to high 0-
                 3
levels (5880 |jg/m ,  3 ppm; 4 hr), the immediate depression in succinate oxidase
activity was followed by an increase that peaked about 2 days postexposure and
returned to normal  by  20 days postexposure (Mustafa et al., 1977).  A 10- or
                                                                  3
20-day continuous exposure  to a lower 0., concentration (1568 |ag/m , 0.8 ppm)
caused an increase  in  Op consumption of  lung homogenate which was greater at
20 days (Mustafa et al., 1973).  When the activity of the mitochondrial fraction
per mg of protein was  measured, the increased actfvity was less than that of
the lung homogenate  per mg of protein.   Morphological  comparisons  indicated
that the exposed lungs had a threefold increase in type 2 cells, which contain
more mitochondria than  type 1 cells.  Thus, the  increase in Op consumption
appears to reflect changes in cell populations.
     Schwartz et  al. (1976) exposed  rats  for 7  days  continuously or  intermit-
tently (8 hr/day)  to 392, 980, or 1568  (jg/m3  (0.2, 0.5, or 0.8 ppm) of 03.
Succinate oxidase activity increased linearly with 03 concentration.  No major
differences were  apparent between continuous and intermittent exposures.  No
statistical  analyses  were  reported.   Concentration-dependent morphological
effects were  also observed (Section  9.3.1).  When  using rats  and an  identical
exposure regimen, Mustafa and Lee (1976)  found similar responses for succinate
oxidase and succinate-cytochrome  ,c reductase activity.  These increases were
statistically significant.  Mustafa  et al.  (1973), when using 7 days of con-
tinuous exposure, also  showed that Op consumption of rats increased with in-
creasing 03 level (392, 980, 1568 yg/m3,  0.2, 0.5, 0.8 ppm).
     Although concentration appears to be a  stronger determinant of the effect,
time of exposure also plays a role (Mustafa  and Lee, 1976).  Rats were exposed
                                   9-89

-------
            o
to 1568 ug/m  (0.8 ppm) of 0, continuously for 30 days, and 0^ consumption was
measured as the  activities  of succinate oxidase, 2-oxoglutarate oxidase, and
glycerol-1-phosphate  oxidase.   On day  I,  the" effect  was  not significant.
However, at day  2  and following,  these  enzyme  activities increased.  The peak
increase occurred  on  day 4 and  remained at that  elevated level throughout  the
30 days of exposure.   Mitochondrial succinate-cytochrome c reductase exhibited
a similar pattern  under similar 07 levels for  7  days of exposure.   Equivalent
                                                                           3
results occurred in rats during  a 7-day continuous exposure to 1568 ug/m
(0.8 ppm) of 0. (DeLucia et a!., 1975a).
                                                        3
     In rats, recovery from  an ozone-induced  (1568 ug/m ,  0.8  ppm; 3 days,
continuous) increase  in  succinate oxidase activity occurred  by 6 days post-
exposure (Chow et  a!., 1976b).   When the rats were re-exposed  to the same
exposure regimen at 6, 13, and 27 days of recovery, the increased activity was
equivalent to that of the initial exposure.   Thus, no long-lasting tolerance
was observed.
     Dietary vitamin  E can  also reduce  the effects of 0,, on  0« consumption.
                                                       3
After 7 days  of  continuous  exposure to  196 or 392  ug/m  (0.1 or 0.2 ppm)  of
Og, the lung  homogenates of  rats  maintained on diets with either 11 or 66  ppm
of vitamin E  were  examined  for changes  in 02  consumption (succinate oxidase
activity) (Mustafa, 1975; Mustafa and Lee,  1976).  In the  11-ppm vitamin E
group, increases in 00 consumption occurred at both 0., levels.  In  the 66-ppm
                              3
vitamin E group,  only 392 ug/m  (0.2 ppm) of 03 caused an increase.   Mitochon-
dria were  isolated from the lungs and  studied.   Neither dietary group  had
Og-induced changes in the  respiratory rate of mitochondria  (on  a per mg of
protein basis).   However, the amount of mitochondria (measured as total  protein
content of the mitochondria!  fraction of the lung) from the 0~-exposed rats of
the 11-ppm vitamin E group did increase  (15-20 percent).
     Similar to earlier discussions for antioxidant metabolism (Section 9.3.3.2),
responsiveness to  effects of 0, on 09 consumption  is age-related (Table  9-5).
                                                                3
Elsayed et al. (1982a) exposed rats of various ages to 1568 ug/m  (0.8 ppm) of
0- continuously  for  72  hr.   The 0.,-induced increase  in the  activities  of
succinate oxidase  and cytochrome  c oxidase increased with age (from 24 to 90
days of age),  with no significant change at 24 days of age.   When younger rats
were examined, succinate oxidase activity decreased in both the 7- and 12-day-
old animals,  with  the younger ones more affected.   The 18-day-old rats had an
increase in this activity.
                                   9-90

-------
     Species and  strain  differences were observed after  a 5-day continuous
exposure to  882 ug/m  (0.45 ppm)  of 03 (Mustafa et al.,  1982).   Mice and
Sprague Dawley rats (but not other strains of rats) had an increase in activi-
ty of  succinate oxidase.   Cytochrome c oxidase was increased  in mice, Long-
Evans  rats,  and Sprague Dawley rats,  but  not Wistar rats;  the  increase in the
mice was greater than that in the rats.
     Rats have also  been  compared to two strains of monkeys after a 7-day (8
hr/day)  exposure  to various concentrations of  0.,  (Mustafa and Lee, 1976).
                                           o
Rats were exposed  to 392,  980, or  1568 M9/m  (0.2, 0.5, or 0.8 ppm) of 03 and
exhibited increases  in succinate oxidase  activity.  Rhesus monkeys exposed to
                       o
either 980 or 1568 ug/m  (0.5 or 0.8 ppm) of 03 had an increase in this enzyme
activity only at the higher  exposure concentration.  When  Bonnet monkeys were
exposed  to 392, 686,  or 980 ug/m   (0.2,  0.35,  or 0.5 ppm) of 03, succinate
oxidase activity increased at the two higher 0« levels.  The number of animals
                                                                              3
used was not specified, which makes  interpretation difficult.  At the 392-ug/m
(0.2 ppm) level, the increase in rats was to 118 percent of controls (signifi-
cant); in Bonnet monkeys,  it was to  113 percent of controls (not significant).
                2
At  the 980-(jg/m   (0.5 ppm)  of 03 level, the magnitude of the significant
increases was not  different  between  rats  (133 percent  of  controls) and Bonnet
monkeys  (130 percent of controls).   Rhesus monkeys may have been slightly less
responsive than rats,  but no statistical analyses were performed to assess
this question.
     The increase  in  levels  of nonprotein sulfhydryls, antioxidant enzymes,
and  enzymes  involved in  02 consumption is typically attributed to concurrent
morphological changes  (Section  9.3.1) in the lungs, principally  the  loss of
type 1 cells and the increase of type  2 cells and the  infiltration of alveolar
macrophages.  Several  investigators  have  made such correlated  observations  in
rats (Plopper et al., 1979; Chow, et a!., 1981; Schwartz et a!., 1976; DeLucia
et  a!.,  1975a).   Type  2  cells  are  more metabolically active  than  type  1  cells
and  have more abundant mitochondria  and endoplasmic reticula.  This hypothesis
is  supported by the findings of Mustafa  et  al. (1973), Mustafa (1975), and
DeLucia  et  al.  (1975a).    For example,  succinate  oxidase  was studied  in  both
lung homogenates  and isolated mitochondria of  rats after  a 7-day exposure to
1568 ug/m  (0.8 ppm) of 03 (DeLucia  et al., 1975a).  The increase in the homo-
genate was  about  double that of the isolated  mitochondria (on a per  mg of
protein  basis).   As mentioned previously, this indicates  that an increase in
                                   9-91

-------
the number  of mitochondria, rather than an increased activity within a given
mitochondrion, in exposed lungs is the probable dominant cause.
9.3.3.4  Honooxygenases.   Multiple microsomal enzymes function in the metabo-
lism of both endogenous (e.g., biogenic amines, hormones) and exogenous (xeno-
biotic)  substances.   These  substrates  are either activated or detoxified,
depending on  the substrate and the enzyme.   Only a few of the enzymes have
been studied subsequent to G"3 exposure (Table 9-6).
     Monoamine oxidase  (MAO) activity has been investigated (Mustafa et a!.,
1977) in view of its importance  in catalyzing  the metabolic degradation of
bioactive amines  like  5-hydroxytryptamine and norepinephrine.   Although MAO
activity is located principally  in the mitochondria, it also  is found in
microsomes.   Activity  levels of MAO in  rats were  determined  after exposure to
         3                               3
3920 ug/m   (2  ppm)  for 8 hr or 1568 ug/m  (0.8 ppm) continuously for 7 days.
Substrates  used  included  n-amylamine, benzylamine,  tyramine, and  3-hydroxyty-
ramine; three  tissue preparations were  used (whole  lung  homogenate, mitochon-
dria, and microsomes).  The acute high-level  exposure reduced MAO activity in
                                                          2
all tissue preparations.  The longer exposure to 1568 ug/m  (0.8 ppm) increased
MAO activity  in  all tissue  preparations.  This  pattern  is  similar to  that
found for mitochondria!  enzymes  and antioxidant metabolism (Sections 9.3.3.2
and 9.3.3.3).
     The cytochrome P-450-dependent enzymes have been studied because of their
function in drug and carcinogen metabolism.   Palmer et al. (1971S 1972) found
                                    Q
that hamsters  exposed to 1,470 ug/m  (0.75 ppm)  of 0,  for  3  hr  had lower
benzo(a)pyrene hydroxylase  activity in  the lung.   Goldstein et  al.  (1975)
                                           q
showed that rabbits exposed to  1,960 ug/m  (1 ppm) of  03  for 90 min  had
decreased levels of lung cytochrome P-450.  Maximal decreases occurred 3.6 days
following exposure.  Recovery to  control values occurred somewhere between 8
days and  45 days.   Cytochrome  P-450-mediated activity of benzphetamine N-
                                                                     3
demethylase in the lung was  lowered by a 24-hr exposure to 1,960 ug/m  (1 ppm)
of 0, in rats (Montgomery and Niewoehner, 1979).  The cytochrome P-450 dependent
activity began to recover by 4 days  postexposure but was still  decreased.
Complete recovery occurred by I we.ek.   Cytochrome bn-mediated lipid desaturation
was stimulated by Og 4, 7, and 14 days postexposure.  Immediately after expos-
ure, the desaturase  activity was  quite depressed, but this was attributed to
anorexia in the rats, and not to 03.  Cytochrome P-450-dependent enzymes exist
in multiple forms,  because  they  have  different  substrate  affinities that
overlap.   Measuring  activity with only one substrate does not characterize a
                                   9-92

-------
TABLE 9-6.   MQNQOXYGENASES
Ozone
concentration
ug/«a
392
980
1568


1470
5880
19,600
1470
5880
19,600
1568


3920
1568


3920
392
1568
1568
1568

1568


ppni
0.2
0.5
0.8


0.75
3.0
10.0
0.75
3.0
10.0
0.8


2.0
0.8


2.0
0.2
0.5
0.8
0.8

0.8


b Exposure
Measurement ' duration and
method protocol
MAST Continuous or
N8KI 8 hr/day for 7
days


I 3 hr


I 3 hr


NO Continuous for
7 days

8 hr
I Continuous
for 7 days

8 hr
I Continuous or
8 hr/day

8 hr/day

I Continuous for
7 days



Observed effect(s) Species
Concentration-related 'linear Rat
increase in NADPH cytochrome c
reductase activity. No
difference between continuous
and intermittent.
Decreased activity of benzpyrene Hamster
hydroxylase in lung parenchyma.

Decreased activity of benzpyrene Rabbit
hydroxylase in tracheobronchial
inucosae.
High 03 level reduced monoamine Rat
oxidase activity; low 03 level
increased it.

High 03 level decreased activity Rat
of NADPH cytochrome c reductase;
low level increased it.

Increased activity of NADPH cyto- Rat
chrome c reductase. At 0.8 ppm,
increase began at day 2 of exposure.
No change in NADPH cytochrome c Monkey
reductase activity.
Increased activity of NADPH cyto- Rat
chrome c reductase on days 2 through
7. Maximal increase on day 4.


Reference
Mustafa and Lee,
1976; Schwartz
et al. , 1976;
Mustafa et al . ,
1977
Palmer et al., 1971


Palmer et al., 1972


Mustafa et al. ,
1977


DeLucia et al. ,
1972, 1975a


DeLucia et al. ,
1975a



Mustafa and Lee,
1976


-------
                                                          TABLE 9-6.   MONOOXYGENASES (continued)
 i
UD
Ozone
concentration
ug/nr1 ppm
1960 1. 0
1960 1. 0
5880 3
b Exposure
Measurement ' duration and
method protocol
NO 90 min
MAST 24 hr
NBKI 10 min before
lung perfusion
and continuous
throughout
experiment.
Observed effect(s) Species
Decreased levels of lung cyto- Rabbit
chrome P-450. Maximal decrease
at 3.6 days postexposure.
50% decrease in benzphetamine Rat
N-demethylase activity 1 day
postexposure; return to control
levels by 1 wk postexposure.
Stimulation of cytochrome
bs-mediated lipid desaturation.
Decreased enzymatic conversion of Rat
arachidonic acid to prostaglandins
when using isolated ventilated per-
fused lung.
Reference
Goldstein et al. ,
1975
Montgomery and
Niewoehner, 1979
Menzel et al . , 1976
        Measurement method:
MAST = Kl-coulometric (Mast meter); CHEM = gas solid chemiluminescence; NBKI = neutral buffered potassium iodide;
I = lodometric; ND = not described.
        Calibration method:  UKI = unbuffered potassium iodide.

-------
single enzyme.   More importantly, the relatively long time for recovery suggests
that cell injury, rather than enzyme destruction, has occurred,   Benzo(a)pyrene
hydroxylase is the  first major enzyme  in the activation of benzo(a)pyrene and
several  other  polycyclic  hydrocarbons  to  an active  carcinogen.   However,
additional enzymes  not  studied  after 0- exposure are involved in the activa-
tion, which makes  full  interpretation  of the effect of 0, on this metabolism
impossible.  The impact of  the  decrease in cytochrome P-45Q  depends on the
activation or  detoxification of  the metabolized compound by this system.
     Also  involved  in mixed  function oxidase metabolism is NADPH cytochrome  c
reductase.  As  with other classes  of  enzymes  (Sections  9.3.3.2;  9.3.3.3),
                                             3
acute exposure to  a high  03 level  (3920 ug/m  ,  2 ppm; 8 hr) reduced NADPH
cytochrome c reductase  activity  (DeLucia et a!., 1972, 1975a).   After a con-
tinuous  or 8 hr/day exposure of rats for 7 days, the activity of NADPH cyto-
chrome c reductase increased linearly in a concentration-related fashion (392,
980, and  1568 ug/m3;  0.2,  0.5,  and 0.8 ppm) (Mustafa and Lee, 1976; Schwartz
et al.,  1976;  Mustafa et  al.s 1977).   Continuous and  intermittent exposures
                                                                  2
were not  different.   The  time course of the response to 1568 |jg/m  (0.8 ppm)
was  an  increase  in activity that began  at  day  2, peaked at  day 4,  and was
still  increased  at  day  7  of continuous exposure in the rat (Mustafa and Lee,
1976).   The  rat, but  not  the Rhesus monkey, is  apparently  affected after
                                             o
exposure for 8 hr/day for 7 days to 1568 ug/m  (0.8 ppm) of 03 (DeLucia et al.,
1975a).  However, monkey data were not reported in any detail.
9.3.3.5   Lactate Dehydrogenase and  Lysosomal Enzymes.   Lactate dehydrogenase
(LDH)  and  lysosomal  enzymes  are  frequently  used  as markers of cellular damage
if  levels  are  observed to increase  in  lung lavage  or serum/plasma, because
these  enzymes  are  released by cells  upon certain  types of damage.   Effects of
0-  on  these  enzymes are described in Table 9-7.   No lung lavage studies have
been reported; whole-lung homogenates were used.   Therefore,  it is not possible
to determine whether the observed increases reflect a leakage into lung fluids
and  a compensatory resynthesis  in  tissue  or cellular changes (See  Section
9.3.1),  such as  an  increase  in type  2 cells and alveolar macrophages and poly-
morphonuclear leukocytes rich in lysosomal hydrolases.  In some instances, cor-
relation with plasma values was sought.  They are described briefly here; more
detail  is  given  in  Section 9.4.3.
      Lactate dehydrogenase  is an intracellular enzyme that  consists of two
subunits  combined as a tetramer.  Various combinations of the two  basic subunits
change  the electrophoretic  pattern of  LDH  so that its various isoenzymes can
                                   9-95

-------
                                                          TABLE 9-7.   LACTATE DEHYDRQGEHASE AND  LYSQS0HAL EHZYHES
vo
Ozone
concentration
{JgTP ppl
196
392
980
1568
980
1568
1568
1372
1568
1372-
1568
1568
1568
0.1
0.2
0.5
0.8
0.5
0.8
0.8
0.7
0.8
0.7-
0.8
0.8
0.8
h Exposure
Heasurenent ' duration and
nethod protocol
NBKI
HAST,
NBKI
NBKI
NBKI
MAST,
NBKI
HAST,
NBKI
HAST,
NBKI
Continuous
for 7 days
Continuous
for 8 days or
8 hr/days for
7 days
8 hr/day for
7 days
Continuous
for 5 days
Continuous
for 7 days
Continuous
for 7 days
Continuous
for 7 days
Continuous
for 7 days
Observed effect(s) Species
Increase in total LDH activity 1n Rat
diet group receiving 0 ppm
vitamin E. Groups with 11 or
110 ppu of vitamin E had no effect.
Increased lung lysozytne activity only Rat
after continuous exposure to 0,8 ppn.
Increased LDH activity in lungs. Rat
Change in LDH isoenzyme distri-
bution at 0.8 ppffl.
No change in total LDH activity or Monkey
isoenzyne pattern in lungs.
Specific activities of various Rat
lysosomal hydrolases increased.
Increase in lung acid phosphatase Rat
activity; no observed increases
in B-glucuronidase activity.
Bronchiolar epithelium had decreased Rat
NADH and NADPH activities and In-
creased ATPase activity.
Increase in LDH activity not Rat
affected by vitamin E (0 or
45 mg/kg diet).
Reference
Chow et al . ,
Chow et al . ,
Chow et al . ,
Dillard et al
1972
Castleman et
1973a
Castleman et
1973b

1981
1974
1977
* 1
al.,
al.,
Chow and Tappel ,
1973
                 Measurement method:
                  Calibration method:
MAST = Kl-couloiiietric (Mast meter); NBKI = neutral buffered potassium iodide.
NBKI = neutral buffered potassium iodide.

-------
be detected.  Chow and Tappel (1973) found an increase in LDH activity in the
                                            3
homogenate of rat lungs exposed to 1568 ug/m  (0.8 ppm) of 0, continuously for
                                3
7 days.   Lower levels  (196  [jg/m ,  0.1 ppm;  7 days continuous) only increased
LDH activity of lung homogenate when rats were on a diet deficient in vitamin" E
(Chow et al., 1981).   Vitamin E levels in the diet did not significantly in-
fluence the response.   In following up this  finding, Chow et al. (1977) studied
LDH activity and isoenzyme pattern (relative ratios of different LDH isoenzymes)
                                                                            3
in the lungs, plasma,  and erythrocytes of (L-exposed rats (980 or 1,568 pg/m ,
                                       3
0.5 or 0.8  ppm) and monkeys  (1568 M9/m >  0.8 ppm).  Exposure was for 8 hr/day
for 7 days.  In monkeys, no significant changes in either total LDH activity or
isoenzyme pattern in  lungs,  plasma, or erythrocytes were detected.  The total
LDH activity  in the  lungs of rats was increased after exposure to 1,568 or
        o
980 ug/m  (0.8 or 0.5 ppm),  but no changes in the  plasma or erythrocytes were
detected.   The isoenzyme  pattern of LDH  following  0-  exposure  was  more com-
plex, with  the  LDH-5  fraction significantly decreased in lungs and plasma of
                          3
rats exposed to 1,568 ug/m  (0.8 ppm).  The LDH-4 fraction in lungs and plasma
and the  LDH-3 fraction  in lungs were  increased.   No changes were discernible
                            3
in rats  exposed to  ,980 ug/m  (0.5 ppm)  of 0-.   The changes in  LDH  isoenzyme
pattern appeared to be due to a relative  increase in the LDH isoenzymes contain-
ing the  H  (heart type) subunits.  Although the increase  in  LDH  suggests cyto-
toxicity after 0- exposure,  no clear-cut interpretation can be placed on the
importance  of the isoenzyme pattern.   Some specific cell types  in the lung may
contain  more  H-type  LDH  than others and  be  damaged by 0- exposure.  Further
studies  of  the  fundamental  distribution  of LDH  in lung cell types are needed
to clarify  this point.
     Lysosomal enzymes have  been found to increase in the  lungs of animals
exposed  to  0~ at  concentrations of 1,372 ± 294  ug/m3 (0.70 ± 0.15 ppm) for 5
                          3
days and 1,548 ± 274  |jg/m  (0.79 ± 0.14 ppm) for 7 days, whether detected by
biochemical  (whole-lung  homogenates and  fractions) or histochemical  means
(Dillard et al., 1972).   Dietary vitamin E  (0 to 1500  mg/kg diet)  did  not  in-
fluence  the effects.  These  increased  activities were attributed to the infil-
tration  of  the lung by phagocytic cells during the  inflammatory response phase
from 0~  exposure.   Similarly, Castleman  et al.   (1973a,b) found that activity
of  lung  acid phosphatase was increased in young  rats that had been exposed to
                   3
1,372 to 1,568 ug/m   (0.7 to  0.8 ppm)  of  03 continuously for 7  days.  Increases
in p-glucuronidase activity were not observed.   The histochemical and cytochem-
ical  localization  suggested that 03 exposure results  in damage to  the lung's
                                   9-97

-------
lysosomal membranes.  Castleman et al. (1973b) also found that the bronchiolar
epithelium in  infiltrated  areas had lower NADPH-  and  NADH  diaphorase  activ-
ities and  higher  ATPase activities than similar epithelium of control lungs.
They discussed  in greater  detail  the enzymatic  distribution within the  lung
and suggested  that some of the pyridine nucleotide-dependent reactions could
represent  an enzymatic  protective mechanism operating locally in the centri-
acinar  regions  of  03-exposed  lungs.   Chow et al.  (1974)  also observed an
increase in  lysozyme  activity (lung homogenate) in rats exposed continuously
            3
to 1568 pg/m   (0.8 ppm)  for 8 days but not in rats exposed intermittently (8
hr/day, 7 days).  No effect was seen in continuous or  intermittent exposure of
                       •3
rats to 392 or 980 pg/m  (0.2 or 0.5 ppm) of Oo.
     Lysosomal  acid hydrolases include enzymes  that digest protein  and can
initiate emphysema.  The contribution of the increases in these enzymes obser-
ved by  some  after 0, exposure  to  morphological  changes  has not been demon-
strated.
9.3.3.6  Protein  Synthesis.   The  effects of 0,  on protein  synthesis  can be
divided into two  general areas:   (1)  the effects on the  synthesis  of  collagen
and related  structural  connective tissue proteins, and (2)  the effects on the
synthesis  or secretion  of  mucus.   The studies  are summarized  in Table 9-8.
     Hesterberg and Last (1981) found that increased collagen synthesis caused
                                                          o
by continuous  exposure  to  03  at 1568,  2352,  and 2940 ug/m   (0.8, 1.2,  and 1.5
ppm) for 7 days could be  inhibited by concurrent treatment with methylpred-
nisol one (1 to 50 mg/kg/day).
     Hussain et al. (1976a,b) showed that lung prolyl  hydroxylase activity and
                                                                           3
hydroxyproline  content  increased  on exposure of rats  to 980 and 1568 ug/m
                                                   3
(0.5 and 0.8 ppm) of 0™ for  7 days.   At 392 ug/m (0.2 ppm), 03 produced a
statistically  insignificant increase  in  prolyl  hydroxylase activity.   Prolyl
hydroxylase is the  enzyme  that  catalyzes the conversion of proline to  hydroxy-
proline in collagen.   This conversion is essential  for  collagen to form the
fibrous conformation necessary  for its structural  function.  Hydroxyproline is
                                                                            3
an indirect  measure of  collagen content.  When rats were exposed to 980  ug/m
(0.5 ppm)  of 0,, for 30  days,  the augmentation of activity seen earlier at
       £       -C?
7 days  of  exposure had diminished, and  by  60  days, the enzyme activity was
within the normal  range despite continued 0» exposure.   When rats were exposed
             3
to 1568 ug/m  (0.8 ppm) of 03> the prolyl  hydroxylase activity continued to
rise for about 7 days;  hydroxyproline content  of the  lung rose to a  maximum
value  at about 3  days after exposure  began  and remained  equivalently  elevated
                                   9-98

-------
                                                        TABLE 9-8.   EFFECTS OF OZONE ON  LUNG  PROTEIN  SYNTHESIS
vo
Ozone
concentration
pgTi3
392
784
1176
1568

980



392
980
1568




392
784
1176
1568

392
1568
3920
882





1568

980-
3920

ppm
0.2
0,4
0.6
0.8

0.5



0.2
0.5
0.8




0.2
0.4
0.6
0.8

0.2
0.8
2.0
0.45





0.8

0.5-
2

b Exposure
Measurement ' duration and
method protocol
ND 8 hr/day for
3 days

Continuous for
1 through 90 days
Continuous for
3 or 14 days and
combined with
H2S04
MAST Continuous for
7 days





UV 8 hr/day for
1 to 90 days



UV, 6 hr/day, 5 days/
NBKI wk, 12.4 wk (62
days of exposure)
ND Conti nuous
for 7 days




NO Continuous
for 90 days
UV 1, 2, or 3 wk




Observed effect(s) Species
Decreased rate of glycoprotein secretion by Rat
trachea] explants at 0.6 ppm.

Decreased rate of glycoprotein secretion.

Increased rate of glycoprotein secretion.



Concentration-dependent increase in lung prolyl Rat
hydroxylase activity. No effect at 0.2 ppm. Meta-
bolic adaptation suggested at 980 ng/m8 (0.5 ppm)
At 0.8 ppm, collagen and noncollagenous protein
synthesis increased; effect on prolyl hydroxylase
returned to normal by about 10 days postexposure,
but hydroxyproline was still increased at 28 days.
At 0.8 ppm, tracheal explants had decreased rate Rat
of glycoprotein secretion for up to 1 wk, followed
by increased rate up to 12 wks. Three day exposure
to three lower concentrations caused decrease at
only 0.6 ppm.
Decrease in collagen and elastin at 0.2 and Rat
and 0.8 ppm; increase at 2 ppm.

Increased collagen synthesis at 5 and 7, but Mouse
not 2 days of exposure. Similar pattern
for increase in superoxide dismutase activity.
Increased prolyl hydroxylase activity at 2, 3,
5, and 7 days of exposure; maximal effect
at day 5.
Increase in prolyl hydroxylase activity through Rat
7 days. No effect 20 days and beyond.
Increased rate of collagen synthesis; fibrosis Rat
of alveolar duct walls; linear concentration
response.


Reference
Last and Kaizu,
1980; Last and
Cross, 1978






H us sain et al. ,
1976a,b





Last et al., 1977




Costa et al., 1983


Bhatnagar et al . ,
1983






Last et al., 1979



-------
TABLE 9-8.  EFFECTS Of OZONE ON LUNG PROTEIN SYNTHESIS  (continued)
Ozone
concentration Measurement '
ug/»3 pjS nethod
vo
H
O
O
980
980-
2940
5000
980
1000
1254
5000
1882
5000
1254
1254
1882
0.5 W
0.5- NBKI
1.5
(NH4)2S04
0.5 NBKI
H2S04
0.64 . W
0.96 OM
(NH4)2S04
0.64 (N
0.64 UM
0.96
. Exposure
duration and
protocol
Continuous for
up to 180 days
Continuous for
7 days
Continuous for
3 to 50 days
Continuous
for 3 days
Continuous for
7 to 14 days
8 hr/day,
361 days
Continuous for
90 days; inter-
Observed effect(s)
Increase 1n protein and hydroxyproline content of
lungs. No change 2 no postexposure.
03 caused linear, concentration-related increases
in collagen synthesis; (HH4)2S04 combined with
03 Increased collagen synthesis rates by 180% at
1.2 and 1.5 ppn 03.
03 increased collagen synthesis at 3, 30, and 50
days; H2S04 combined with 03 increased collagen
synthesis rates by 220%.
No significant effect of 03 or (NH4)2S04 on
collagen synthesis; (NH4)2 S04 + 03 increased
collagen synthesis rates by 230%.
Interstitial edema and inflammation of proximal
alveolar ducts; (NH,t)2 S04 increased the severity
of OB effects at lesion sites without Increasing
the number of lesions.
Increase in collagen content
Equivalent increase in collagen in all but the
0.64 ppn continuous group which only had a
Species Reference
Rat Last and Greenberg,
1980
Rat Last et al., 1983
Rat Last et al., 1984a
Monkey Last et al . , 1984b
Rat Last et al., 1984b
(young
aiittent units of
5 days (8 hr/
day) of 03, and
9 days of air,
repeated 7 times
with i total of
35 exposure
days over a 90-day
Interval
•arginal (p <0.1) increase.
adult)

-------
                                            TABLE 9-8.  EFFECTS OF OZONE ON LUNG PROTEIN SYNTHESIS  (continued)
Ozone
concentration
VD
H
O
H
|jg/m3
1254
1882
1882
2352
1882
1882
1254
1568
1568
1568
2352
2940
ppm
,64
,96
0.96
1.2
.96
.96
0.64
0,8
0.8
0.8
1.2
1.5
. Exposure
Measurement * duration and
method protocol
8 hr day
7 days/wk,
6 wk
Continuous
for 3 wk
Continuous for
1 wk, then
2 wk of air
Intermittent
units of 3 days
(8 hr/day) of 03,
and 4 days of air,
repeated 6 times
7 wk, 8 hr/day,
7 days/wk
UV Continuous
for 7 days
NO Continuous
for 7 days
NBKI Continuous
for 3 days
UV Continuous
for 7 days
Observed effect(s)
Increased collagen content at 0.96 ppm 03.
At 6 wk post-exposure, both 03 levels in-
creased collagen. Suggestion of progressive
effects.
Both groups had an equivalent increase in lung
collagen content.
No effect on collagen content.
Increase in collagen content.
Increased lung collagen and protein synthesis
rates; results of statistical analyses were not
reported.
Decreased protein synthesis on day 1; increased
synthesis day 2 and thereafter; peak response on
days 3 and 4.
Increased protein synthesis; recovery by 6 days
later; after re-exposure 6, 13, or 27 days later,
protein synthesis increased.
Net rate of collagen synthesis by lung minces
increased in concentration-dependent manner;
methyl prednisolone administered during 03
exposure prevented increase.
Species Reference
Rat
(wean-
ling)
Rat,
(young
adult)
Rat,
(wean-
ling)
Rat,
(young
adult)
Rat Myers et al . , 1984
Rat Mustafa et al . ,
1977
Rat Chow et al . , 1976b
Rat Hesterberg and Last,
1981
aMeasurement method:
 Calibration method:
MAST = Kl-coulometric (Mast meter); UV = UV photometry; NBKI = neutral buffered potassium iodide; ND = not described.
NBKI = neutral buffered potassium iodide.

-------
through day 7  of exposure.  Incorporation of  radiolabeled  amino acids into
collagen and noncollagenous protein rose to a plateau value at about 3 (colla-
genous) or 4 to 7 (noncollagenous) days after exposure.  Synthesis of collagen
was about  1.6  times greater than that  of  noncollagenous proteins  during  the
first few  days  of exposure; no major differences were  apparent  by 7 days of
exposure.   After  the  7-day exposure ended, about 10  days  were required for
recovery to  initial values of  prolyl  hydroxylase.   However,  hydroxyproline
levels were still increased 28 days postexposure.  This suggests that although
collagen biosynthesis returns to normal, the product of that increased synthesis,
collagen, remains stable for some time.
     The shape of the concentration-response curve was investigated by Last et
al.  (1979)  for biochemical  and histological  responses of  rat lungs after
                                                                              3
exposure to ozone for 1, 2, or 3 weeks at levels ranging from 980 to 3920 ug/m
(0.5 to 2  ppm).   A general correlation was  found  between fibresis detected
histologically and the quantitative changes in collagen synthesis in minces of
Og-exposed rat  lungs.   The stimulation of collagen  biosynthesis was essen-
tially the same, regardless of whether the rats had been exposed for 1, 2 or 3
weeks; it  was  linearly  related  to the 03 concentration  to which  the  rats  were
exposed.
     Protein deficiency  and food restriction do  not  have a  major influence on
the effects of 03 on  lung  hydroxyproline,  lung elastin, or  apparent  rates  for
lung collagen  synthesis and elastin accumulation (Myers  et al., 1984).   In
                                                                              3
this study weanling or young adult rats were exposed continuously to 1254 ug/m
(0.64 ppm) Oo  for 7 days.   It appears that 0~ caused an increase in apparent
lung collagen  and protein  synthetic rates and no major  change  in elastin
accumulation, but  the results  of  statistical  analyses were  not reported.
                                                         3
     Continuous exposure of mice  for 7 days to  882  |jg/m   (0.45 ppm)  of  03
caused an increase in collagen synthesis after 5 or 7, but not 2 days of expo-
sure (Bhatnagar  et  al.,  1983).  The  5-  and 7-day results showed  little if any
difference.  The effect on synthesis of noncollagen protein was not significant.
Prolyl hydroxylase  activity was also increased  at  2t 3,  5, and 7 days  of
exposure, with the  maximal  increase at day 5.   The  day 7 results were only
                         \
slightly different  (no  statistical analysis) from the day 2 data.  Superoxide
dismutase activity was  investigated,  because it has  been observed (in other
studies) to prevent  a superoxide-induced increase in collagen synthesis  and
prolyl hydroxylase.   Activity  of  superoxide  dismutase increased in a pattern
parallel  to that for collagen synthesis.
                                   9-102

-------
     Bhatnagar , et al .  (1983)  also studied rats exposed  continuously  for 90
days to 1568 ug/m  (0.8 ppm) of 0,,.  Prolyl hydroxylase  activity continued to
increase through 7 days  of exposure.   By 20,  50,  and 90 days of exposure, no
significant effects were observed.
     Last and his  colleagues  performed a series of  studies  to evaluate the
effects of subchronic exposure to Q3 on rats.   In one experiment (Last et al.,
19845), young adult (60-65 days old) rats were exposed continuously for 90 days
to 1254 ug/m  (0.64 ppm) or 1882 ug/m  (0,96 ppm) 03; the higher level increased
lung collagen content,  while  the lower level  only caused a marginal (p <0.1)
increase.   Rats were  exposed  to these same concentrations in an intermittent
regimen consisting of 5 days of exposure  (8 hr/day),  9 days  of air, 5  days of
exposure, 9 days of air, etc., for a total of 35 days of 0- and 54 days of air
                                                          C/
within the 90-day  experiment.  Both concentrations of 03 caused an  equivalent
increase in collagen  content.  The magnitudes  of the  effects after  continuous
or intermittent exposure were not statistically different.
     Weanling rats (28 days old) were examined after a 6-wk exposure (8 hr/day,
7 days/wk) to 1254 yg/m   (0.64 ppm) or 1882 ug/m3  (0.96  ppm) 0-, (Last  et al.,
                                                                             3
1984b).  Immediately  after exposure, only those animals  exposed to  1882
(0.96 ppm) 03  exhibited  an increase in lung collagen content.  However, 6 wk
postexposure,  animals  exposed  to both the high  and  low concentrations  of Q~
had  increases  in collagen  content.   It appears  that  the collagen content
increased during this postexposure period, but no statistical comparisons were
reported.
                                                                   Q
     Young adult rats exposed continuously for 3 weeks to 1882 ug/m  (0.96 ppm)
DO  had  an  increase  in lung collagen  (Last  et al.,  1984b).   Rats exposed to
         3
2352 ug/m  (1-2 ppm)  0,  for I wk and examined  2 wk  later had an  increase in
                                                                 3
collagen equivalent to that of the 3-wk exposure group (1882 ug/m , 0.96 ppm).
                                                                             2
     Weanling  rats were also studied after intermittent  exposure to 1882 |jg/m
(0.96 ppm) 03  (Last  et al. ,.. 1984b)..  -The regimen.-was  3 days of exposure for
8 hr/day,  followed  by 4  days  of air; this  unit was repeated 6 times.   No
significant changes in collagen  content occurred.
     Lung  collagen  of juvenile  cynomolgus  monkeys  (6-7 mo  old at start of
                                 3
exposure),  exposed  to 1254 ug/m  (0.64 ppm)  03  for  8 hr/day, 7 days/wk for
I yr  (361  days),  was  studied by  Last et  al.  (1984b).   Collagen content was
increased.   Collagen  type ratios were determined, and there were no apparent
shifts  in  collagen  types; however,  the  authors report that  given the  surgical
variation,  small  shifts  in collagen types would not  be likely to be  detected.
                                    9-103

-------
                                   3
     Rats were  exposed to  980 jjg/m  (0.5 ppm)  continuously  for up to 180 days
and  examined  at various  times during  exposure and 2 months after exposure
ceased (Last  and  Greenberg,  1980).  The total  protein  content  of the lungs
increased during  exposure, with  the  greatest  increase occurring after 88 days
of exposure.  By  53 days postexposure,  values  had  returned  to control levels.
Hydroxyproline content of the lungs also increased following 3,  30,  50,  or 88,
but  not  180,  days of exposure.   No  such effect was observed at the 53-day
postexposure  examination.  The  rates of protein and hydroxyproline synthesis
were also measured.  Protein synthesis was not affected significantly.   Although
the  authors  mentioned that  hydroxyproline  synthesis rates  "appeared to be
greater," statistical significance was not discussed and values appeared to be
only slightly increased, considering the variability of the  data.   In  dis-
cussing  their data,  the  authors referred to a concurrent morphological  study
(Moore and Schwartz, 1981)  that showed an increase in lung volume, mild thick-
ening of the  interalveolar septa and alveolar  interstitium,  and an increase in
collagen  (histochemistry)  in these areas.     Different  results  for  collagen
levels were observed  after another longerrterm exposure (6  hr/day, 5 days/wk,
12.4 wk)  to 392,  1568,  or 3920  ug/m3  (0.2, 0.8, or  2.0 ppm) of 03 (Costa et
a!., 1983).   At the  two  lower concentrations,  rats  exhibited  an  equivalent
decrease  in hydroxyproline.   At the highest  concentration,  an  increase was
observed.  Similar findings were made for elastin levels.
     Most reports, such as those described  above,  are on the effects of Q3 on
collagen synthesis.   Very  little is  known about the  effects of 03 on collagen
turnover (i.e., the integration  of synthesis  and degradation).  Curran et al.
(1984) found  that in  vitro exposure  to  high levels of 0,, (19600 ug/m3, 10 ppm
                                                       ^
for 1-4 hr) caused degradation of collagen.   A lower level,  490 ug/m  (0.25 ppm)
did not cause degradation,  but the collagen became more susceptible to proteoly-
tic degradation.
     Last (1983)  and  Last  et al. (1983) investigated the interaction between
0~ and aerosols  of ammonium  sulfate [(NH.^SO.] and sulfuric acid (H^SO^)  in
rats.  Rats were  exposed continuously  for 7 days to  four concentrations of 0,,
                      OO                                       O
ranging  from  980  ug/m to  2940  ug/m   (0.5 to  1.5  ppm) Og.  According to the
authors, these levels were  determined by the KI method and should be multiplied
by 0.8  to compare to UV photometric methods  (i.e., 0.4 to 1.2  ppm).   The
authors'  actual  measured values  are noted here.  The 03 exposure resulted in a
linear,   concentration-related increase  in  collagen synthesis rate.   A 7-day
                              3
continuous exposure to 5 mg/m  (NH»)2SO. (0.8-1.0 urn mass median aerodynamic
                                   9-104

-------
diameter, MMAD)  had  no  effect.   However, when mixtures  of 0-. from 1568  to
2940 (jg/m3  (0.8  to 1.2 ppm) and 5 mg/m3  (NH4)2$04 were used, the collagen
synthesis rate increased to about 180% over the  rates of 07  for  the higher 07
                                     Q                     O                  O
levels (apparently 2352 and 2940 (jg/m , 1.2 and 1.5 ppm).
     Mixtures of 03  and H2$04 were also  reported  (Last  et al. ,  1983).   Rats
were exposed to  980  (jg/m   (0.5 ppm)  0, or this  0, level in  combination  with
1000 (jg/m3  H2S04 (0.38  (jm, MMAD) continuously  for  3 to  50  days.   As expected,
Oo exposure  increased collagen  synthesis  rates at  the three  times of  examina-
tion (3, 30, and 50 days).   The mixture of 0- and H,,S04 caused greater effects.
Examination of the slope ratio of regression lines  indicated that hUSO. in the
mixture  caused a 220 percent  enhancement.  The authors  state that FLSO.  alone
had no effects.
     The statistical  procedures applied in these collagen  studies (Last et al.,
1983) were  questioned by Krupnick and  Frank  (1984) in a letter to the editor.
The general  criticism was  that  too few statistical  tests were applied to test
the hypothesis of  synergism and that too few pollutant data points were  used
in the  design  to develop robust regression  lines.   The authors  (Last et al.
1983) responded  that they  did apply most of the statistical  tests and found
significant  differences between regressions  comparing  03  only  to 03 plus
aerosols,  but  the journal  would not accede to  publishing these analyses.
     The  proximal  acinar regions of rats from the above-mentioned collagen
                                        3
studies, exposed for 1  days to 2352 (jg/m  (1.2 ppm) 0Q alone  or  in combination
              3
with 5000 ug/m   (NH4)2S04, were also evaluated.  Ammonium  sulfate alone caused
no morphological or morphometric effects.  Ozone exposure  resulted in  a thicken-
ing of  the interstitium and an influx of inflammatory cells.  The mixture of
0-  and   (NH4)2$04  caused  the same response plus  an apparent deposition  of
fibrous  material.  These  lesions were  then examined morphometrically.   In the
03 and  03 plus (NH4)2$04 groups, there were  no changes  in  the volume  ratio of
the lesion  per lung or  the volume ratio of the extracellular  connective  tissue
per lesion.   In  examining volume density of different cell  types  in  the  lung
lesions,  both  groups had an  increased percentage  of fibroblasts and  smooth
muscle.   When  the  number of cells  of each type present  per area  of lesion was
calculated,  there  was  a 340 percent increase  in the number  of fibroblasts in
the 0-  plus (NH.)pSO, group compared to the 0-, group.   These  findings  correlate
well with  the  biochemical  effects.
     These  studies were expanded by  Last  et al.  (1984a)  to better evaluate the
influence  of length  of  exposure.  All exposures  were continuous  to 03,  (NH4)2S04,
                                   9-105

-------
                                                                               3
or a mixture of the two; the ammonium sulfate concentration was about 5000 ug/m
(0.5 |jm MMAD), the precise concentration depending upon the experiment.   After
                                      3
3 days of exposure to either 1254 ug/m  (0.64 ppm) 03 or (NH4)2S04, no significant
effects on collagen  synthesis  of rats were observed;  the  mixture of 03 and
(NH*)2SO» more than  doubled the collagen synthesis  rate over controls.  The
pulmonary lesions  (i.e.,  aggregations  of interstitial inflammatory cells  in
terminal bronchioles)  of the  proximal  acinus in these  rats  were examined
histologically and morphometrically.  Ammonium sulfate alone caused no effects.
In the  0, and 03 plus  (NH4)2S04  groups,  there was no  fibrosis, but thickening
by edema  and inflammatory cell  infiltrate was  observed.   Morphometrically,
(NH4)2S(L caused  no effects.   However,  03 and 0~ plus (NH4)2S04  increased
total cell  numbers in  lesions,  with  the mixture producing a significantly
greater  increase  over  03  alone.   This increase was  principally  due  to an
increase in  the numbers  of macrophages, monocytes,  and fibroblasts.   The
increase in  the number  of fibroblasts  is consistent with the biochemical
findings.
     Similar measurements were made on rats exposed continuously for 7 days to
1882 ug/m  (0.96 ppm)  03,  5000 ug/m  (NH4)2S04,  or a mixture of the two (Last
et a!., 1984a).   Control data were not presented and all statistical  comparisons
reported were in  reference to the 0~ alone group.  Ammonium sulfate exposure
resulted in fewer, numbers of cells than Q3 alone.  For macrophage and monocyte
numbers, 0~  and (NH4)2S04 were apparently additive.   For fibroblast and total
cell numbers, 03 and (NFL)2SO, were apparently synergistic.
     The frequency of occurrence of lesions was also examined in the 3-day and
the 7-day studies described above (Last et al., 1984a).  There was no difference
in the  frequency  of lesion occurrence between 0, and Q3 plus (NH»)2SO,.  The
individual  lesions were larger in the 03 plus (NH4)2S04 group; the (NH4)2S04
group had almost no  lesions.   Using collagen  staining procedures,  the 0, plus
(NH^pSO, group had  a greater volume of collagen than did lesions in the  0,
alone group.  The  authors summarized these studies  (Last  et al.,  1983; Last
et al., 1984a) by  stating that  (NH.)2SO, increases the severity of 03 effects
at lesion sites without^ increasing the  number of  lesions; and the cellular
changes correlate  with biochemical  and histological indications of potential
later fibrosis.  From  all  these  studies  it is apparent that synergism occurs.
However, it further appears that the synergism is dependent on the concentration
of 03.  For instance, Last et al. (1983) demonstrated that 5000 ug/m  (NH4)2SQ4

                                   9-106

-------
                                                               3             3
increased collagen synthesis rates in rats exposed to 2352 [jg/m  or 2940 ug/m
                                                            3
(1.2 or 1.5 ppm)  0,,  but not in those exposed  to 1568 ug/m  (0.8 ppm) CU.
     Protein synthesis (incorporation of  radiolabeled leucine) was increased
                                                          3
in rats after  3  days  of continuous exposure  to 1568 |jg/m  (0.8 ppm) of Q,
(Chow et  al. ,  1976b).   Recovery  had occurred by 6  days  postexposure.   No
adaptation was observed, because  when animals were re-exposed to the same Qg
regimen 6, 13 or 27 days after the first exposure, protein synthesis increased
as it had earlier.  Mustafa  et al.  (1977)  investigated the time course  of  the
increased jjn vivo incorporation of radioactive ami no acids.  Rats were exposed
                                    2
continuously for 7 days to 1568 ug/m  (0.8 ppm) of 0,.  No statistics were re-
ported.   One day of exposure caused a decrease in protein synthesis.  However,
by day 2, an increase  occurred, which peaked  on  days  3 and 4  of exposure.  On
day 7, the effect had  not diminished.  The authors attributed this  finding to
synthesis of noncollagenous protein.  They also found no radioactive incorpora-
tion into blood  or alveolar macrophages.   Hence, the observed increases were
due to lung tissue protein  synthesis, and  not a concurrent  influx  of  alveolar
macrophages or serum into the lungs.
     The production of mucous glycoproteins  and  their secretion by trachea!
explants have  been reviewed by Last and  Kaizu  (1980).   Mucous glycoprotein
synthesis and secretion were measured by the rate of  incorporation of  H-glucos-
amine into mucous glycoproteins and their  subsequent  secretion into supernatant
fluid of  the  trachea!  explant culture medium.  This  method has been found to
be a reproducible index  of  mucous  production, and these  authors maintain  that
this measurement ex y 1 vo following  exposure  |ri  vivo  is  representative of
                                                                3
injuries occurring in vivo.  When  rats were exposed to 1568 ug/m  (0.8 ppm) of
0., for 8  hr/day for 1 to  90 days, Last et al.  (1977) found  a depression  of
glycoprotein synthesis  and  secretion into the  tissue culture medium  for  the
initial week that was  statistically significant only  on days  1 and 2 of exposure.
Rebound occurred subsequently,  with increased glycoprotein secretion  for at
least 12 weeks of  continued  exposure to 03 (only  significant  at 1 and 3 months
of exposure).   Rats were also exposed intermittently for 3 days to 1176,  784,
            o
and 392 ug/m   (0.6,  0.4, 0.2 ppm)  of  O^.   Glycoprotein  secretion  decreased
only  at  the higher  concentration.  Tracheal  explants from  Bonnet monkeys
exposed to 0, 980, or  1568 ug/m3 (0, 0-5 or 0.8  ppm)  Og for 7 days  appeared to
have increased  rates  of secretion of mucus (Last and Kaizu,  1980).  However,
few monkeys were used  and statistical analysis was not reported.
                                   9-107

-------
                                             3
     A combination  exposure to  0-  (980 |jg/m ; 0.5 ppm)  and  sulfuric acid
                         3
(HgSO*) aerosol  (1.1 mg/m )  caused   complex effects on mucous secretions  in
rats  (Last  and  Kaizu,  1980;  Last and Cross,  1978).   A 3-day exposure to
        3
980 ug/m  (0.5 ppm) of 0,  decreased mucous, secretion rates, but HpSO, had no
effect.  In  rats exposed to the combination of H-SO,  aerosol and 0~, mucous
secretion significantly  increased.  After 14 days of continuous exposure,  the
rats  receiving a combined  exposure  to both H?SO.  aerosol  and 03 had elevated
values (132 percent) over the control group of animals.  Because mucous secretion
and synthesis are intimately involved in diminishing the exposure of underlying
cells to 03  and  removing adventitiously  inhaled particles,  alterations in  the
mucous secretory rate may have significant biological importance.   Experiments
reported to  date do not clearly indicate what human  health effects may be
likely, nor their importance.
9.3.3.7  Lipid Metabolism and Content of the Lung.   If 0™  initiates  peroxi-
dation of unsaturated  fatty  acids in the lung, then changes in the  fatty acid
composition  of  the lung indicative  of  this process  should be  detectable.
Because the  fatty  acid content of the lung  depends  on the dietary  intake,
changes in fatty acid content due to 0. exposure are difficult to determine in
the absence of rigid dietary control.  Studies on lipid metabolism and content
of the lung  are  summarized in Table 9-9.   Generally,  the unsaturated fatty
                                                  3
acid  content  decreased in rats exposed to 980 |jg/m   (0.5  ppm) of 03  for up to
6 weeks (Roehm et a!., 1972).
      Peroxidation of polyunsaturated  fatty  acids  produces pentane and ethane
to be exhaled in the  breath of  animals  (Donovan and Menzel, 1978;  Downey  et
al.,  1978).   A  discussion  of the use of ethane and pentane as indicators  of
peroxidation  is  presented  by Gelmont et al. (1981) and Filser et al. (1983).
Normal animals  and humans  exhale  both  pentane and  ethane in the  breath.
Dumelin et al. (1978b) found that exhalation of ethane decreased and exhalation
of pentane increased  when  rats were deficient in  vitamin E and exposed to
         3
1960 ug/m  (1 ppm)  of 0., for 60 min.  The  provision  of 11 (minimum vitamin
requirement) or 40 IU (supplemented level) of vitamin E acetate per kg of diet
resulted in  a decrease  in  expired  ethane  and pentane after 03 exposure.
Dumelin et  al.   (1978a)  also measured breath ethane and  pentane  in Bonnet
                                       a
monkeys exposed to 0,  980,  or 1568 M9/m  (0, 0.5,  or 0.8 ppm) of 03 for 7, 28,
or 90 days.   They failed to detect  any  additional  ethane or pentane in the
breath of these  monkeys and attributed the lack of such additional evolution
to be due to the high level of vitamin E provided in the food.
                                   9-108

-------
                                             TABLE 9-9.  EFFECTS OF OZONE EXPOSURE ON LIPID METABOLISM AND CONTENT OF THE LUNG
H
O
Ozone
concentration
ng/m3
980
980
1568
1960
1960
ppm
0.5
0.5
0.8
1.0
1.0
Exposure
Measurement duration and
method protocol
I Continuous for
2, 4, or 6 wk
UV 8 hr/day for 7,
28, or 90 days
NBKI . 60 min
ND 4 hr
Observed effect(s) Species
Increase in arachidonic and palmitic Rat
acids; decrease in oleic and lino-
leic acids.
No effect on ethane and pentane Monkey
production in animals fed diets
supplemented with high levels of
vitamin E.
With vitamin E-deficient diet, in- . Rat
creased pentane production and de-
creased ethane production. With
vitamin E supplementation of 11 or
40 IU vitamin E/kg diet, decreased
ethane and pentane production.
Decreased incorporation of fatty Rabbit
acids into lecithin.
Reference
Roehm et al . , 1972
Dumelin et al. ,
1978a
Dumelin et al. ,
1978b
Kyei-Aboagye
et al. , 1973
                 Measurement method:  NBKI = neutral buffered potassium iodide; UV = UV photometry; I = iodometric; ND = not described.

-------
     Kyei-Aboagye et al, (1973) found that the synthesis of lung surfactant in
rabbits, as  measured by  dipalmitoyl  lecithin synthesis,  was  inhibited by
                      3
exposure to  1960  pg/m  (1 ppm) of 03  for 4 hr.   Pulmonary lavage showed an
increase in radiolabeled lecithins.   The authors proposed that 0~ may decrease
lecithin formation while  simultaneously stimulating the  release of surfactant
lecithins.   This  may  suggest the presence  of a larger disarrangement of  lipid
metabolism following  03 exposure.   However, although changes  in  lipid com-
position of  lavage  fluid  occur, the changes apparently do not alter the sur-
face tension  lowering properties  of the fluid, as  shown by Gardner et  al.
                                                                          3
(1971) and Huber  et al. (1971) when using high  levels  of 03 (> 9800 ug/m ;
5 ppm).
9.3.3.8  Lung Permeability.   Table  9-10 summarizes studies of the effects
on lung permeability of exposures to different concentrations of 0,,.
     The lung possesses several active-transport mechanisms  for  removal of
substances from the airways to the capillary circulation.  These removal mech-
anisms have  been  demonstrated  to be carrier-mediated and specific for  certain
ions.  Williams et al.  (1980)  studied  the  effect of 0- on  active transport of
                                                            3
phenol red in the lungs of rats exposed to 1176 to 4116 (jg/m  (0.6 to 2.1 ppm)
of 03  continuously for  24  hr.   Ozone inhibited the carrier-mediated transport
of intratracheally instilled  phenol  red from the lung to the circulation and
increased the nonspecific diffusion of phenol red from the lung. These changes
in ion permeability may also explain in part the effects of 03 on the respira-
tory response  of  animals  to bronchoconstrictors (Lee et  al., 1977; Abraham et
al., 1980).
     As another index of  increased  lung permeability following Q~ exposure,
the  appearance of albumin and  immunoglobins  in  airway secretions has  been
                                                                            3
examined.   Reasor et  al.  (1979) found that dogs breathing 1960 to 2940 ug/m
(1.0 to 1.5  ppm)  of 0,, had increased  albumin  and immunoglobjn G content of
their airway secretions. Alpert et al.  (1971a), using rats exposed for 6 hr to
                 3                           .
490  to  4900  |jg/m   (0.25 to 2.5 ppm) 0_, also found increased albumin in lung
                                     3
lavage in animals exposed to 980 ug/m  (0.5 ppm) or more.
     In a  series  of  experiments, Hu et al.  (1982) exposed  guinea pigs  to 196,
510, 1000, or 1960  ug/m3  (0.1, 0.26,  0.51,  or  1.0  ppm) of 03 for 72 hr and
found increased lavage  fluid protein content sampled immediately after exposure
to concentrations >  510 ug/m   (0.26 ppm)  as compared with controls.   Ozone-
exposed guinea pigs had no accumulation of proteins when the exposure time was
                                   9-110

-------
                                                        TABLE 9-10.  EFFECTS OF OZONE ON LUNG PERMEABILITY
H
H
Ozone

concentration
pg/m3
196
510
1000
1960
353

510
980


1000

490
880
1960
4900
1176
2156
3136
4116
1960-
2940
ppm
0.1
0.26
0.51
1.0
0.18

0.26
0.5


0.51

0.25
0.5
1.0
2.5
0.6
1.1
1.6
2.1
1.0-
1.5
k Exposyre
Measurement ' duration and
method protocol
CHEM, 3 hr or 72 hr
UV


3 hr; 8 hr/day
for 5 or 10 days
3, 24, or 72 hr



3 hr

MAST, 6 hr
NBKI


NBKI 24 hr



HAST 2 hr



Observed effect(s) Species Reference
Increased levels of lavage fluid protein Guinea Hu et al., 1982
>_ 0.26 ppm immediately after 72-hr exposure pig
or 15 hr after a 3-hr exposure. Vitamin C
deficiency did not influence sensitivity.
No effect on protein levels.

Lavage was 15 hr postexposure. At 0.26 ppm in-
creased protein only after 24 hr exposure; at
6.5 ppn, increased protein after 24 hr
of exposure.
Increase in protein levels at 10- and 15- (but not
0, 5, or 24) hr postexposure.
Increased alveolar protein accumulation at 0.5 ppm Rat Alpert et al.,
and above. 1971a


Concentration-dependent loss of carrier-mediated Rat Williams et al.,
transport for phenol red. 1980


Increased albumin and immunoglobin G in airway Dog Reasor et al.,
secretions. 1979
       Measurement method

       b,
MAST = Kl-coulometric (Mast meter); CHEM = gas  phase  chemiluminescence;  UV = UV photometry; NBKI = neutral buffered
potassium iodide.
       Calibration method:  KBKI = neutral buffered potassium iodide; UV = UV photometry.

-------
reduced from 72 to 3 hr, unless the time of lavage was delayed for 10 to 15 hr
following exposure.  The protein content of the  lavage fluid determined 10 to
15 hr following'a  3-hr exposure increased in a  concentration-related manner
                                                                            c
                                                                             3
from 500 to 1470 pg/m3 (0.256 to 0.75 ppm).  Again 196 pg/m3- (0.1 ppm) had no
effects.  The lavage fluid protein content of guinea pigs exposed to 353 pg/nf
(0.18 ppm) of  DO  for 8 hr per day for 5  or 10 consecutive days was not dif-
ferent from air controls.   No effect-of vitamin C deficiency could be found on
the accumulation  of  the lavage fluid protein  in  guinea pigs exposed to 196 to
1470 pg/m3 (0.10  to  0.75  ppm) 03 for  3  hr (Hu et al.s 1982).   In contrast,
vitamin Odeficient guinea pigs have increased sensitivity to NO,, (Selgrade et
a!., 1981).  Polyacrylamide  gel electrophoresis  of  lavage fluid proteins from
                                            3
animals exposed for  3 hr  to  196 to 1470 |jg/m   (0.1  to 0.75 ppm) QS showed the
appearance of extra protein bands which co-migrated with serum proteins and of
increased intensity  of  bands that also occur  in air-exposed controls.  This
led the authors to conclude  that  the main source of the  increased protein was
serum.
     Prostaglandins  are intermediates  in  tissue  edema resulting from a wide
variety of mechanisms  of  injury.   Increased lung permeability and edema pro-
duced by  03  might also  be mediated by prostaglandins.  Non-steroidal anti-in-
flammatory drugs  (aspirin and  indomethacin) at appropriate doses inhibit lung
edema in rats from exposure to 7890 pg/m  (4 ppm) of 0, for 4 hr (Giri et al.,
1975).  Prostaglandins  F2  and £„ were markedly increased in plasma and lung
lavage  of  rats  exposed  to 7840 pg/m   (4  ppm) for up to 8 hr  (Giri et al.,
1980).  Ozonolysis of arachidonic acid jm vitro  produces fatty acid peroxides
and other products having prostaglandin-like activity (Roycroft et al., 1977).
Fatty acid cycloperoxides are produced directly by ozone-catalyzed peroxidation
                                                                3
(Pryor, 1976;  Pryor  et  al., 1976).   Acute exposure to 5880 ug/m  (3 ppm) 03
inhibits uncompetitively rat lung prostagl-andin cyclooxygenase (Menzel et al.3
1976).
     Earlier  reports that  prostaglandin  synthesis  inhibitors exacerbated
Q3-produced  edema (Dixon  and Mountain, 1965; Matzen, 1957a,b), as did methyl
prednisolone (Alpert et al., 1971a), tend to confuse the interpretation of the
role of prostaglandin in 03~produced injury.   Prostaglandin synthesis inhibitors
clearly inhibit the  degradation of prostaglandins and alter the balance between
alternative pathways of fatty acid peroxide metabolism.  Thus, while prostanoids
are highly likely to be  involved in Q3~produced edema,  their exact role is
still unexplained.
                                   9-112

-------
9.3.3.9  Proposed Molecular Mechanisms of  Effects.   Experts generally agree
that the toxicity  of 0,, depends on  its  oxidative properties.   The precise
mechanism of O3's  toxicity at the subcellular  level  is  unclear, but  several
theories have been advanced.   These theories include the following:

     1.   Oxidation of  polyunsaturated  lipids  contained mainly in cell
          membranes;
     2.   Oxidation of  sulfhydryl, alcohol,  aldehyde,  or amine  groups in
          low molecular weight compounds or proteins;
     3.   Formation of  toxic  compounds  (ozonides and peroxides) through
          reaction with polyunsaturated lipids;
     4.   Formation  of  free  radicals,  either  directly  or  indirectly,
          through lipid peroxidation; and
     5.   Injury mediated  by  some pharmacologic  action, such  as  via a
          neurohormonal mechanism, or release of  histamine.

     These mechanisms have been discussed  in several reviews:  U.S. Environmen-
tal Protection  Agency  (1978); National  Air  Pollution  Control Administration
(1970); North Atlantic  Treaty Organization  (1974); National Research Council
(1977); Shakman (1974); Menzel (1970, 1976); Nasr (1967); Cross et al. (1976);
Pryor  et al.  (1983);  and Mudd and Freeman (1977).   From these reviews and
recent  research detailed here or  in  previous biochemistry  sections, two hypo-
theses  are favored and may in fact be related.
9.3.3.9.1  Oxidation of polyunsaturated lipids.   The  first hypothesis is  that
03  initiates peroxidation  of  polyunsaturated fatty  acids (PUFA) to peroxides,
which  produces  toxicity through  changes in  the properties of cell membranes.
Ozone addition to ethylene groups of PUFA  can take place in membranes yet give
rise to water-soluble  products that  can  find their  way to  the cytosol.  Alde-
hydes,  peroxides,  and  hydroxyl  radicals formed by peroxidatien all can react
with proteins.   In addition to the  direct oxidation  of amino  acids  by 0,,
                            *i                                             j
secondary  reaction products  from  O^-initiated PUFA  peroxidation  can also
oxidize amino acids or  react  with proteins to alter the  function of the proteins.
Since  large numbers of  proteins are embedded with lipids in membranes and rely
on  the associated  lipids to  maintain the  tertiary  structure of the protein,
alterations  in  the lipid  surrounding the  protein can result in  structural
                                   9-113

-------
changes of the membrane-embedded protein.  At present, methods are not avail-
able to differentiate  effects  on  membrane proteins from effects on membrane
lipids.
     Some of  the strongest evidence  that the  toxic reaction of 03  can  be
associated with  the  PUFA of membranes is  the  protective  effect of dietary
vitamin E on  03  toxicity (Roehm et al.,  1971a, 1972; Chow and Tappel, 1972;
Fletcher and  Tappel, 1973; Donovan et al., 1977; Sato et al., 1976a;  Chow and
Kaneko, 1979; Plopper  et al,,  1979; Chow  et al., 1981; Chow, 1983; Mustafa et
al., 1983; Mustafa, 1975).   Generally, vitamin  E reduced the CU-induced increase
in enzyme activities of the  glutathione peroxidase system (Section 9.3.3.2)
and those involved in  oxygen consumption  (Section 9.3.3.3).  More details are
provided  in Table 9-5.  Morphological effects  due  to  CL  exposure are also
lessened by dietary  vitamin  E  (Section 9.3.1.4.1.1).   Although this  is  not
the strongest evidence,  vitamin  E supplementation also prevented 0~-indueed
changes in red blood cells (Chow and Kaneko,  1979).
     These data  consistently show that vitamin  E has a profound effect on the
toxicity of 0- in animals.   They also support indirectly lipid peroxidation as
a toxic lesion in animals.   However,  although the influence of dietary vitamin
E is clear,  its  relation to vitamin  E  levels  in the lung, where presumably
most lipid peroxidation  would  occur,  is poorly understood.  Rats of  various
ages (5,  10 and  90 days  old and 2 yrs old) were fed normal diets; and 90-day-
old rats  were fed diets containing 0, 200, and 3000 mg of vitamin E/kg diet
(Stephens et al., 1983).  (Most of the biochemical studies of vitamin E protec-
tive effects  were conducted  with  diets having far less than 200 mg/kg diet.)
                               3
Rats were exposed to  1764 pg/m   (0.9 ppm) 03  continuously  for 72  hr,  and
periodic morphological  observations were  made.  Those animals on normal diets
had equivalent levels of vitamin E in the lung, but responses of the different
age groups differed.   Animals of a given age (90 days) maintained on the three
vitamin E diets  had  different  levels  of vitamin E in the  lungs (5.3 to 325 yg
of  vitamin  E/g  of tissue), but morphological  responses  were very similar.
Stephens  et al.  concluded  that 03~induced responses in the lung are  indepen-
dent of the vitamin E content of the lung.  Independent interpretation of this
study  is  not  possible,  since very minimal  descriptions of  responses were pro-
vided  and the number of animals was  not  given.  Also,  others (Chow et al.,
1981;  Plopper et al., 1979) found that for a  given age  of  rat,  different
dietary levels of vitamin  E (and perhaps  different  lung  levels as shown by
Stephens  et al.,  1983) influenced the morphological responses of rats to 03-
                                   9-114

-------
9.3.3.9.2  Oxidation of sulfhydryl or amine groups.  The  second  hypothesis  is
that ozone exerts its toxicity by the oxidation of low-molecular-weight compounds
containing thiol, amine, aldehyde, and alcohol functional groups and by oxida-
tion of  proteins.  Mudd and Freeman  (1977) present a summary  of  the  arguments
for oxidation of  thiols,  amines, and proteins as the primary mechanism of 03
toxicity based  upon  i_n vitro  exposure data.  Ami no acids are readily oxidized
by 03  (Mudd et  al.,  1969; Mudd and Freeman, 1977;  Previero  et a!., 1964).   In
the following descending order of rate, 0~ oxidizes the ami no acids cysteine >
methionine >_ tryptophan > tyrosine >  histidine  > cystine  >  phenylalanine. The
remaining common  ami no acids  are not oxidized  by  03,   Thiols are the  most
readily  oxidized  functional  groups of proteins and  peptides (Mudd  et  al.,
1969; Menzel, 1971).   Tryptophan in  proteins is also oxidized in vitro by 0~
                                                               	 "" '"'"'	     O
as  shown by studies  of avidin,  the  biotin-bitiding  protein.   Oxidation of
tyrosine in egg albumin by 0, occurs  in  vitro,  converting the 0.,-oxidized egg
                            *5         «—-  , IJTTT,,„,, -».,- m,.,                  J
albumin  to a form immunologically distinct from native  egg  albumin  (Scheel  et
al., 1959). Ozone inactivated human alpha-1-protease inhibitor i_n vitro (Johnson,
1980). When treated  with  0^,  alpha-1-protease  inhibitor  lost its ability to
inhibit  trypsin, chymotrypsin, and elastase.
     Meiners et al.  (1977)  found that 0- reacted  in vitro with tryptophan,
                                         J          —-~• T~  """
5-hydroxytryptophan, 5-hydroxytryptamine, and 5-hydroxyindolacetic  acid.  One
mole of  0~ was  rapidly consumed  by each mole  of indole compound.  Oxidation of
tryptophan by Q?  a^so generates  hydrogen peroxide.  Hydrogen peroxide is a
toxic  substance in  itself and initiates peroxidation of lipids (McCord and
Fridovich, 1978).  Other active  Op species such as HO- and Op~ are formed from
hydrogen peroxide (McCord and Fridovich, 1978).
     The results  of the oxidation of functional  groups in proteins  can  be
generally observed by  reduction  of enzyme activities at high  concentrations of
                            3
0,  (viz.,  1960  to 7840 M9/m , 1 to  4 ppm for  several hours).   Many  enzymes
examined (Tables  10-5  to  10-10)  in tissues  have decreased activities (in  many
cases  not  statistically significant) immediately following  even lower level
          3
(1960 p:g/m , 1  ppm)  0, exposures.  The enzymes  decreased,include those  cataly-
zing  key steps  supplying reduced cofactors to  other processes  in the  cell,
such as  glucose-6-phosphate dehydrogenase (DeLucia et al.,  1972) and succinate
dehydrogenase (Mountain,  1963).   Cytochrome  P--450 (Goldstein et al., 1975),
the  lecithin  synthetase system  (Kyei-Aboagye  et al., 1973), lysozyme (Holzman
et  al.,  1968),  and the prostaglandin synthetase system  (Menzel  et al., 1976)
are  decreased  by 0™ exposure.   Respiratory control  of  mitochondria is lost,
                                   9-115

-------
and mitochondria! energy production is similarly decreased (Mustafa and Cross,
1974).
     Mudd and  Freeman  (1977)  point out that proteins are the major component
of nearly all cell membranes, forming 50 to 70 percent by weight.   The remainder
of the weight  of the cell membranes  is  lipids  (phospholipids,  glycolipids,
glycerides, and  cholesterol).   Polyunsaturated  fatty acids are components of
membrane phospholipids, glycolipids, and glycerides.  Mudd and Freeman contend,
however, that  proteins  are far more easily oxidized  by 03 than are lipids.
Indirect evidence in support of the idea that amines in particular are oxidized
preferentially by 03 is the protective effect of p-aminobenzoic acid (Goldstein
et  al.»  1972a).   Rats  injected with  p-aminobenzoic acid were  partially
protected from the  mortality  due to high  concentrations- of 03.  Presumably,
the added p-aminobenzoic acid is oxidized by 03 in place of proteins.   Goldstein
and Balchum (1974) later suggested that the protection of p-aminobenzoic acid,
allylisopropy!acetamide, and  chlorpromazine was due to  the induction of mixed
function oxidase  systems  rather than  a direct free  radical scavenging effect.
However, they  also  recognized that chlorpromazine can mask free radicals and
result  in  membrane stabilization,  which could account for  the protective
effects of these compounds preventing edema and inflammation.   Acetylcholines-
terase  found  on  red blood cells is protected from  inhibition  by  jm  vitro
                        o
exposure to  78,400 ug/m   (40 ppm) 03 through p-aminobenzoic  acid  treatment
(Goldstein et  a!.,  1972a).  Amines  are  also  efficient  lipid antioxidants,
                                        t
so the  results of  these  experiments  could be interpreted in  favor of the
theory of peroxidation  of the cell membrane as a mechanism  of toxicity, as
we! 1.
9.3.3.9.3  Formation of toxic compounds  through reaction with polyunsaturated
lipids.   The effects of 03 could be  due to  the elaboration  of products  of
peroxidation as  well as peroxidation of the membrane itself.  Menzel et  al.
(1973) injected  10  pg  to  10 ug of fatty  acid ozonides  from oleic,  linoleic,
linolenic, and arachidonic acids into animals and  found increased vascular
permeability measured by  extravascularly located  Pontamine Blue dye bound to
serum proteins.   Extravascularization  of serum  proteins could be  blocked by
simultaneous antihistamine injection or by prior treatment with compound 48/80,
a substance that depletes  histamine stores.  Cortesi and Privett (1972) injected
methyl linoleate  ozonide  into rats or gave the compound orally; in each case
acute pulmonary edema resulted.   The pulmonary edema and changes in fatty acid
                                   9-116

-------
composition of the serum and lung lipids were similar to those occurring after
03 exposure (Roehm et al., 1971a,b, 1972; Menzel et a!., 1976).
     Peroxidation of lung lipids could lead to cytotoxic products.  Phosphatid-
yl choline  liposomes  (spheres  formed by emulsifying phosphatidyl choline  in
water) were  lysed  on  exposure to 0- (Teige et al., 1974).  Liposomes exposed
to On were  more  active  than 0,  alone  in  lysing  red  blood  cells.   The products
of ozonolysis of phosphatidyl choline could be stable,  yet toxic, intermediates.
9.3.3.9.4    Formation of  free  radicals and  injury mediatedby  pnarmacologic
action.   The  other theories may be  linked to  the consequences  of peroxidation
of PUFA.  Ozonation of  PUFA results in the  formation of peroxides (e.g., ROOH
or RQOR) rather than oxidation of alkenes to higher oxidation states (e.g., RON,
RCHO or RC02H).  Peroxides (ROOH or ROOR) are chemically reactive and may be the
ultimate toxicants, not simply products of oxidation.   During the process of
peroxidation  or  direct  addition of  0, to PUFA,  free radicals may be  generated
(Pryor et al.,  1983),  and these free radicals may be the ultimate toxicants.
     Because  the metabolism of PUFA peroxides in  the  lungs  is  intimately
linked with the metabolism of  thiol  compounds  (such as  glutathione, GSH),
direct oxidation of thiols by 0, (hypothesis B) may link the two major hypothe-
ses A and  B with  hypothesis C,  formation of toxic products.  Ozone depletion
of GSH could  render the peroxide detoxification mechanism ineffective.   However,
                                 q
at levels of Oo below 1960 |jg/m  (1 ppm),  increases in glutathione have been
observed  (Plopper  et al.,  1979; Fukase et al., 1975;  Moore  eta!., 1980;
Mustafa et  al.,  1982).  The rat  lung is  sensitive to the  increase of glutathione
peroxidase,  glutathione  reductase,  and  glucose-6-phosphate  dehydrogenase
activities  at levels  as low  as 196  yg/m   (0.1 ppm)  0- continuously  for  7 days
(Plopper  et al.,  1979; Chow-et al.,  1981;  Mustafa, 1975; Mustafa and  Lee,
                                                                              3
1976) in vitamin E-deficient rats.  With vitamin E-supplemented  rats, 392 pg/m
                                                                      q
(0.2 ppm) caused similar  effects.   After acute exposures to >1960 jjg/m  (1 ppm)
Oq, decreases are  observed in these enzyme  activities and glutathione  levels.
Other species (mice and monkeys) exhibit similar effects at different concentra-
tions  (Table 9-5). These changes at the lower levels  of Q,  appear to  be
coincidental  with  the initiation of repair and proliferative  phases of lung
injury  (Cross et al.,  1976;  Dungworth et al., 1975a,b;  Mustafa and  Lee,  1976;
Mustafa et  a!.,  1977, 1980; Mustafa and  Tierney, 1978).  The parallel increase
in the number of type 2 cells  having  higher levels  of metabolic  activity would
be expected to cause an  overall  increase  in the metabolic activity of  the
                                    9-117

-------
lung.  This was  substantiated by Mustafa (1975)  and  DeLucia et al.  (1975a),
who  found  that the  02 consumption per mitochondrion was  not  increased in
Og-exposed lungs  but that  the number  of mitochondria  was increased.  Chow and
Tappel  (1972)  proposed that  alterations of  enzymatic activity were due to
stimulation of glutathione peroxidase pathway.   Chow and  Tappel  (1973) also
proposed that the changes  in the pentose shunt and glycolytic enzymes in lungs
of the  Oj-exposed rats were due to the demands of the glutathione peroxidase
system for reducing equivalents in the form of NADPH.   Cross et al. (1976) and
DeLucia (1975a,b)  reported that 03 exposure oxidized glutathione  and formed
mixed disulfides between proteins and non-protein sulfhydryl compounds.
     The general  importance of glutathione in preventing lipid peroxidation in
the absence of 0- has been shown by Younes and Siegers (1980) in rat and mouse
liver where depletion of glutathione  by treatment with vinylidene  chloride or
diethylmaleate led to  increased  spontaneous peroxidation.  These authors
suggest that glutathione prevents  spontaneous peroxidation by suppression of
radicals formed by the enzyme cytochrome P-450 or already produced hydroperox-
ides.
     Chow and Tappel  (1973)  suggest that peroxides  formed  via  lipid  peroxida-
tion increase glutathione peroxidase activity and, in turn, increase levels of
the  enzymes necessary to  supply reducing equivalents (NADPH)  to  glutathione
reductase.   Vitamin E suppresses spontaneous formation of lipid peroxides and,
therefore decreased  the glutathione peroxidase activity in  mouse red  cells
(Donovan and Menzel,  1975; Menzel  et  al., 1978).  The supplementation of rats
with vitamin E could, therefore,  decrease the  utilization of  glutathione by
spontaneous reaction  or by ozone-initiated peroxidation.   The two mechanisms
could then interact  in a concerted fashion  to  decrease  ozone  cytotoxicity.
Eliminating vitamin E from the diet increases the chances of increases  of this
system.
     In support of hypothesis E, Wong and Hochstein (1981) found that thyroxin
enhanced the osmotic fragility of human erythrocytes exposed to 0- jjn  vitro.
They found also that 125I  from radio!abeled thyroxin was incorporated into the
major membrane glycoprotein, glycophorin, of red cells.   When these events had
occurred, the cation permeability  of  the human  red  cells was enhanced without
measurable inhibition of ATPase or membrane lipid peroxidation.  They suggested
that thyroid hormones play an important role  in  03 toxicity.  Fairchild and
Graham (1963) found that thyroidectomy, thiourea, and antithyroid drugs protec-
ted  animals from  lethal exposures to 03 and nitrogen dioxide.   Fairchild and
                                   9-118

-------
Graham ascribed  the mortality  following 03 to pulmonary  edema.   Wong and
Hochstein (1981) suggested that Q3 toxicity in the lung may be altered through
a free  radical  mechanism  involving  iodine transfer from  thyroxin to lung
membranes.  The hormonal  status of animals could alter a  variety  of  defense
mechanisms and 03 sensitivity.
9.3.3.9.5  Summary.  The  actual toxic mechanism of CL may involve a  mixture
of all of these chemical  mechanisms because of the  interrelationships between
the peroxide  detoxification  mechanisms and glutathione (See Section 9.3.3.2)
and the  complexity  of  the products produced from ozonation of  PUFA.   A single
chemical  reaction may  not be adequate to  explain 0-. toxicity.  The relative
importance of any one reaction, oxidation of proteins, PUFA, or small  molecular
weight compounds, will depend upon a number of factors such as the presence of
enzymatic pathways  of  decomposition  of products formed (peroxides), pathways
for regeneration of thiols, the presence of non-enzymatic means of terminating
free radical reactions (vitamins E and C), and differences  in membrane composi-
tion of PUFA (relative ease of  attack of 03), for example.

9,3.4  Effects on Host Defense  Mechanisms
     The  mammalian  respiratory tract  has a number of closely coordinated  pul-
monary .defense  mechanisms that, when  functioning normally, provide protection
from the  adverse  effects  of  a wide variety of  inhaled  microbes and other  par-
ticles.   A  variety  of  sensitive and reliable methods  have  been used to assess
the effects  of  (L on  the" various components  of this defense system to provide
a better  understanding of  the health effects of this pollutant.
     The  previous  Air Quality Criteria  Document for Ozone and Photochemical
Oxidants  (U.S.  Environmental  Protection  Agency,  1978) provided a  review  and
evaluation  of the scientific literature published  up  to  1978 regarding the
effects of 03 on  host defenses.  Other reviews have recently been  written that
provide  valuable  references  to the complexity of the host defense  system and
the effects  of  environmental chemicals  such as 0,  on its  integrity (Gardner,
1981; Ehrlich, 1980; Gardner  and Ehrlich,  1983; Goldstein,  1984).
     This section describes the existing  data base and,  where appropriate,
provides  an  interpretation of the data,  including an assessment of the different
microbial defense parameters  used, their sensitivity in detecting  abnormalities,
and the  importance of the abnormalities with  regard  to the pathogenesis of
infectious  disease in  the exposed host.   This  section also  discusses the
                                    9-119

-------
various components of host defenses, such as the mucociliary escalator and the
alveolar macrophages, which clear the lung of both viable and nonviable parti-
cles, and integrated mechanisms, which are studied by investigating the host's
response to experimentally  induced pulmonary infections.  The immune system,
which defends the overall host against both infectious and neoplastic diseases,
is also discussed.
9.3.4.1  Mucociliary Clearance.  The  mucociliary transport system is one of
the lung's primary defense  mechanisms against inhaled particles.   It protects
the conducting airways by trapping and quickly removing material that has been
deposited on  the mucociliary  escalator.   The effectiveness of mucociliary
clearance can be  determined by measuring such biological  activities  as the
rate of transport of deposited particles; the frequency of ciliary beating;
structural integrity of the ciliated cells; and the size, number,  and distribu-
tion of mucus-secreting  cells.  Once  this defense mechanism has been altered,
a buildup of  both nonviable and possibly viable inhaled substances can occur
on the epithelium and may jeopardize  the  health  of the  host, depending  on the
nature of the uncleared  substance.  As an example, regardless of the severity
of the influenza  infection, the virus concentrates initially in the lining of
the airways.   The virus then spreads to alveolar cells of the lung parenchyma.
It is suspected  that the macrophage might be the site of replication of this
virus (Yilyma et a!., 1979; Nayak et a!., 1964;  Raut et a!., 1975).
     A number of  studies with various animal species  have reported morpho-
logical damage to the  cells of the tracheobronchial  tree from acute and sub-
                                     3
chronic exposure  to  490  to 1960 ug/m  (0.20 to 1.0 ppm) of 03.   (See Section
9.3.1.)  The  cilia were either completely  absent or had become noticeably
shorter or blunt.   By  removing these animals to a clean-air environment, the
structurally  damaged cilia  regenerated and appeared normal.  Based on such
morphological observations,  related effects  such  as ciliostasis,  increased
mucus secretions, and  a  slowing of mucociliary  transport  rates might be ex-
pected.  However, no measurable changes  in  ciliary beating activity have been
reported due  to  0~ exposure alone.   Assay  of isolated trachea! rings  from
                                                      3
hamsters immediately after a 3-hr exposure to 196 ug/m  (0.1 ppm)  of 03 showed
no significant loss  in ciliary beating activity  (Grose  et  a!.,  1980).   In the
                                                                             3
same study, when  the animals were  subsequently exposed  for 2 hr to 1090 ug/m
HSO*  (0.30 urn volume  median diameter), a  significant  reduction  in ciliary
                                   9-120

-------
beating frequency occurred.   The magnitude of this effect was, however, signi-
ficantly less than  that  observed due to  the effect of H?SCL  expbsure  alone.
In either case, the animals completely recovered within  72 hr when allowed to
remain in a clean-air environment.  These authors found only a slight decrease
in beating frequency with a simultaneous exposure to 196 ug/m  (0.1 ppm) of 0_
            3
and 847 yg/m  H^SCL (Grose et a!., 1982).  These data indicate that 0, appears
to partially protect against the effects of H-SO^ on ciliary beating frequency.
Grose et al. (1982) proposed that the ciliary cells may be partially protected
due to the  increase in mucus trachea! secretion  of  glycoproteins (Last and
Cross, 1978) resulting from exposure to these chemicals.
     Studies cited  in the  previous' criteria document  (U.S.  Environmental
Protection  Agency,  1978) gave  evidence  on the  effect of 0,,  on  the host's
                          *                                 O
ability to  physically remove  deposited particles  (Table 9-11).   The  slowing
of mucus  transport  in both rat and rabbit trachea as a result of Oj exposure
was reported in the early literature (Tremer et al., 1959; Kensler and Battista,
1966).  Goldstein,  E.,  et  al.  (1971a,b,  1974)  provided evidence that the
primary effect of 0., on the defense mechanism of the mouse lung was to diminish
                   *5                        *
bactericidal activity  but not to  significantly  affect physical removal  of  the
deposited bacteria.   In  these  studies,  mice were  exposed to an  aerosol of
32
  P-labeled  Staphylococcus  aureus either  after  a 17-hr exposure to  0^ or
before a  4-hr exposure.   Concentrations  of Q,  were 1180,  1370, 1570,  or
         3
1960 ug/m   (0.6,  0.7,  0.8,  or  1.0 ppm).   The physical clearance  and bacteri-
cidal capabilities  of the lung were then measured 4 to  5 hr  after  bacterial
exposure.   Exposure  17 hr before infection caused a significant  reduction in
bactericidal activity beginning at 1960 jjg/m  (1.0 ppm) of 0_.  When mice were
exposed to  0,  for 4 hr after  being  infected, there was  a significant  decrease
in bactericidal  activity for each 0., concentration, and with increasing 0,
concentration, there was  a progressive decrease in bactericidal activity.  The
investigators proposed that because mucociliary clearance was unaffected  by
subsequent  0,,  exposure,  the bactericidal  effect was  due  to  dysfunction of  the
alveolar  macrophage.   Warshauer et al.  (1974) reported  that  a deficiency  in
vitamin E would further  reduce  the lungs'  bactericidal activity.
      Friberg et  al.  (1972) studied the  effect of  a 16-hr/day exposure to  980
    •3
H9/m   (0.5  ppm)  of 0, on the  lung clearance rate of radiolabeled monodisperse
polystyrene  and  iron particles in the  rabbit and of bacteria in the  guinea
                                    9-121

-------
TABLE 9-11,  EFFECTS OF OZONE OH HOST OEFEHSE MECHANISMS:   DEPOSITION AND CLEARANCE
Ozone
concentration Measurement Exposure
jjgTi3 pjS method duration and protocol
196
784
784
*P 785, 1568,
H 1960
NJ
K)
1764
980
980
980, 1960
784-3979
0.1 CHEM
0.4 NBKI
0.4 NO
0.4, 0.8, UV
1.0
0.5 HAST
0.5 NBKI
0.5 NBKI
0.5, 1.0 CHEM
0.57-2.03 M
3 hr
3 hr
4 hr
4 hr
1.4 hr
16 hr/day,
7 months
2 months
2 hr
17 hr before
bacteria
Observed effects
No effect on ciliary beating frequency.
Initially lowers deposition of inhaled bacteria,
but subsequently a higher number are present
due to reproduction.
Bactericidal activity inhibited. Silicosis
did not enhance 03 effect.
Delay in nueociliary clearance, acceleration
in alveolar clearance.
Increases nasal deposition and growth of
virus; no effect in the lungs.
No effect on clearance of polystyrene and
iron particles.
•Reduced clearance of viable bacteria.
Reduced tracheal mucus velocity at 1.0 pprn.
Ho effect at O.S ppn.
• Physical clearance not affected, but bactericidal
activity affected at 0.99 ppm. Decrease in
deposition of inhaled organisms at 0.57 ppm.
Species Reference
Hamster Grose et a!
Mouse Coffin and
., 1980
Gardner, 1972b
Mouse Goldstein et al. , 1972b
Rat Kenoyer et
Mouse Fairchild,
Rabbit Friberg et
Guinea pig Friberg et
Sheep Abraham et
al., 1981
1974, 1977
al., 1972
al., 1972
al., 1980
Mouse Goldstein et al., 1971a

-------
                             TABLE  9-11.   EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:  DEPOSITION AND CLEARANCE  (continued)

Ozone
concentration Measurement
Exposure
ug/ms ppm method duration and protocol
y3
H
NJ
W
1176 0.62-4.25 M
1372 0.7 H
1372 0.7 G
1S68 0.8 UV
1960 1.0 NBKI
2352 1.2 UV
4 hr after
bacteria
7 days
3-4 hr
4 hr
3 hr
4 hr
Observed Effects
No effect on bacterial deposition and
clearance; reduced bactericidal activity
at each exposure level.
Deficiency of Vitamin E further reduced
bactericidal activity after 7 days.
Reduced bactericidal activity in lungs.
Slowed tracheobronchial clearance and
accelerated alveolar clearance. Effects
greater with higher humidity.
Bacteria clear lung and invade blood.
Delayed mucociliary clearance of particles.
Species Reference
Mouse Goldstein
Rat Warshauer
Mouse Bergers et
Rat Phalen et
et al . , 1971b
et al., 1974
al . , 1982
al., 1980
Mouse Coffin and Gardner, 1972b
Rat Frager et
al., 1979
Measurement method:   NO  = not  described; CHEM = gas phase chemiluminescence; UV = UV photometry; NBKI = neutral buffered potassium iodide;
 MAST = Kl-coulonetric (Mast meter); M = microcoulomb sensor; G = galvanic meter.

-------
pig.  *The results  from the guinea pig  studies  showed a reduced clearance of
viable bacteria.  The rabbit's lung clearance rate was not affected by 0~.   In
this latter study,  however,  a large number  of  the test animals died during
exposure from  a respiratory  disease,  and the  results must  be  viewed with
caution.
     Recent studies have continued to examine the effects of 03 on mucociliary
transport in the intact  animal.   Phalen et al.  (1980) attempted to quantitate
the removal rates  of deposited material  in  the upper and lower respiratory
tract of the rat.   In this study, the clearance rates of radio!abeled monodis-
perse polystyrene  latex spheres  were  followed after 0~  exposure.   A 4-hr
                     3
exposure to 1568 |jg/m  (0.8 ppm) of 03 significantly slowed the early (tracheo-
bronchial) clearance and accelerated the late  (alveolar) clearance  rates at
both low (30 to 40 percent) and  high (> 80 percent)  relative  humidity.  These
effects were even  greater at higher humidity, which produced nearly additive
effects.  Combining  03 with various sulfates  [Fe2(SO.)2,  H^SO.,  (NH^gSO^]
gave clearance  rates  very similar to those  for 03 alone.   Accelerating the
long-term clearance  from the  alveoli may not  in  itself be harmful;  however,
because this process may result from an influx of macrophages into the alveolar
region, the  accumulation of excess  numbers  of macrophages  might present a
potential health hazard  because  of their high content of proteolytic enzymes
and Oy  free radicals,  which  have the  capability for tissue destruction.
Essentially no  data  are available on the effects of prolonged exposure to 0,
on  ciliary functional  activity or on mucociliary transport rates measured in
the intact animal.
     Frager et  al.  (1979)  deposited insoluble, radioactive-labeled particles
via inhalation  and monitored  the  clearance rate after a 4-hr  exposure to 2352
    3
ug/m  (1.2 ppm) of 0~.  This exposure  caused  a substantial  delay in rapid
(mucociliary) clearance  in  the rat.   However,  if  the animals were exposed 3
                          3
days earlier to 1568 ug/m  (0.8 ppm) of 03 for 4 hr, the pre-exposure elimi-
nated this effect,  resulting in a clearance rate that was essentially the same
as for controls.  Thus, the pre-exposure to a lower level 3 days before rechal-
lenging with a  higher  concentration  of  03 appeared to afford  complete protec-
tion at 3  days.  After a 13-day  interval between the pre-exposure and the
challenges, this adaptation or tolerance was lost.
                                   9-124

-------
     These results were  confirmed  when Kenoyer et al. (1981) repeated these
                                                                            2
studies, with three different concentrations of 0-, 784, 1568, and 1960 jjg/m
(0.4, 0.8, and 1.0 ppm).   At each of the three concentrations, a delay was ob-
served in early  (0  to 50 hr postdeposition  of particles)  clearance,  and an
acceleration was seen in long-term (50 to 300 hr postdeposition) clearance, as
compared to controls.   Concentration-response curves showed that clearance was
affected more by the higher concentrations of 0,.
     The velocity of the trachea! mucus of sheep was not significantly altered
                                                                      3
from a baseline value of 14.1 mm/min after a 2-hr exposure to 980 ug/m  (0.5 ppm)
of 03 (Abraham et al., 1980).  The authors state that 1960 ug/m3 (1 ppm) of 03
for 2 hr did significantly reduce, both immediately and 2 hr postexposure, the
trachea! mucus velocity.
9.3.4.2   AlveolarMacrophages.   Within the  gaseous  exchange region of the
lung, the first line of defense against microorganisms and nonviable insoluble
particles is the resident population of alveolar macrophages.  These cells are
responsible for  a  variety of important activities, including detoxification
and  removal  of  inhaled  particles,  maintenance of pulmonary  sterility, and
interaction with  lymphoid cells for immunological protection.   In addition,
macrophages act as scavengers  by  removing cellular debris.  To adequately
fulfill their purpose,  these defense cells must  maintain  active mobility, a
high  degree of  phagocytic activity, an integrated membrane structure,  and a
well-developed  and  functioning  enzyme  system.  Table 9-12  illustrates the
effects of 03 on the alveolar macrophage (AM).
      Under normal conditions, the number of free AMs located  in the alveoli is
relatively  constant when measured  by lavage  (Brain  et al., 1977, 1978).
Initially, Q-,  through  its  cytolytic action  that  is probably mediated  through
its  action  on  the cell  membrane,  significantly reduces  the total  number of
these defense  cells immediately after exposure (Coffin  and Gardner, 1972b).
The  host  responds  with  an  immediate  influx of cells to  aid the  lung in
combating  this  assault.   Little is  known about the mechanisms of action that
stimulate  this  migration or about  the  fate  of these immigrant cells.  The
source  of these new cells may be either (1)  the influx  of  interstitial  macro-
phages,  (2)  the  proliferation  of interstitial macrophagic  precursors  with
subsequent migration  of  the  progeny  into the air  space,  (3)  migration of blood
monocytes,  or  (4) division  of  free  AMs.  The  rapid increase  in the number of
macrophages  is  evidently a  biphasic response, arising from  an  early  phase
                                   9-125

-------
TABLE 9-12.  EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:  HACROPHAGE ALTERATIONS
Ozone
concentration
{jg/B3
1%
1960
392
490
980
K m
a\
980
980
980
1313
980
1960
1058
PP»
0.1
1.0
0.2
0.25
0.50
0.5
0.5
0.5
0.5
0.67
0.5
1
0.54
Measurement3 ' Exposure
tnethod duration and protocol
NBKI
MAST
NBKI
NBKI
NBKI
NBKI
CHEN
NBKI
NBKI
UV
2.5 hr or
30 nin in vitro

8 hr/day for
7 days
3 hr
(in vivo and
Tn vitro)
8 hr/day for
7 days
3 hr
3 hr
3 hr
2 hr
(in vitro)
23 hr/day for
34 days
Observed effects0
Lung protective factor partially inactivated,
increasing fragility of utacrophages (concen-
tration-related).
Increased number of macrophages in lungs
(norphology).
Decreased activity of the lysosomal
enzymes lysozyme, acid phosphatase,
and p-glucuronidase.
Increased osmotic fragility.
Decreased enzyme activity and increased influx
of PMHs.
Decreased red blood cell rosette binding to
macrophages.
Decreased ability to ingest bacteria.
Decreased agglutination in the presence of
concanavalin A,
Increased number of raacrophages (morphological).
Species
Rabbit
Monkey
Rat
Rabbit
Rabbit
Rabbit
Rabbit
Rabbit
Rat
Mouse
Reference
Gardner et al . , 1971
Castleaan et al., 1977
Dungworth, 1976
Stephens et al., 1976
Hurst et al., 1970
Hurst and Coffin, 1971
Dowel 1 et al., 1970
Alpert et al., 1971b
Hadley et al., 1977
Coffin et al., 1968
Coffin and Gardner, 1972b
Goldstein et al., 1977
Zitnik et al., 1978

-------
                              TABLE 9-12.   EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:  MACROPHAGE  ALTERATIONS  (continued)
Ozone
concentration
ug/ma
1168
1568
1568
1568
& I960
1 9800
— i
SJ
^ 1960
3136
6860
ppn
0.8
0.8~
0.8
0.8
1
5
1
1.6 to
3.5
Measurement ' Exposure
method duration and protocol
NO 11 days
NO 90 days
UV 7 days
MAST 3, 7, 20 days
CHEM 3 hr
UV 4 hr
N8KI 2 hr to 3 hr
Observed effects0 Species
No effect on in vitro interferon production Mouse
with alveolar macrophages but did inhibit
the production of interferon by tracheal
epithelial cells.
Eightfold increase in number of macrophages at Rat
7 days, reducing to fourfold after 90 days.
Decreased number of migrating macrophages Monkey
and total distance migrated.
Increased phagocytosis. Rat
Decreased ability to produce interferon Rabbit
in vitro.

Decreased in vitro niarational ability, as Rat
evidenced by decreased number of macrophages able
to migrate.
Decreased superoxide anion radical production. Rat
Reference
Ibrahim et al., 1976
Booraan et al . , 1977
Schwartz and Christman, 1979
Christman and Schwartz, 1982
Shlngu et al., 1980
McAllen et al., 1981
Amoruso et al . , 1981
Witz et al., 1983
4900
2.5
                                       5 hr
 Loss of p-glucuronidase and acid phosphatase in
. PAH with ingested bacteria; decreased rate of
 bacterial ingestion.
                                                                                                     Rat
Goldstein et al., 1978b
4900
2.5
                                       5 hr
 Diminished rate of bacterial killing, increased
 numbers of intracellular staphylococcal clumps;
 lack of lysozyme in macrophages with staphylococcal
 clumps.
                                                                                                     Rat
Kimura and Goldstein, 1981
aMeasure«ent method:   NO = not described;  CHEM = gas  phase chemiluminescence; UV = UV photometry;  NBKI = neutral  buffered  potassium iodide;
 MAST = Kl-coulometrlc (Mast meter);  H = nicrocoulomb sensor.
 Calibration method:   NBKI = neutral  buffered potassium iodide.

 Abbreviations used:   PMN = polymorphonuclear leukocytes;  PAM = pulmonary alveolar macrophage.

-------
apparently correlated to a local cellular response and a later phase of inter-
stitial cell proliferation, which  is responsible for the maintenance of the
high influx of macrophages (Brain et al., 1978).
     Morphological studies have  supported the observation that exposure to 03
can result in a raacrophage influx in several animal  species.   Exposure of mice
            3
to 1058 ug/m   (0.54 ppm) of 03  23 hr/day for a maximum of 34 days resulted in
an increased number of macrophages within the proximal alveolar ducts (Zitnik
et al., 1978).   These cells were highly vacuolated and contained many secondary
                                  i
phagocytic vacuoles filled with cellular debris.   The effect was most prominent
after 7 days of exposure and became less evident as the exposure continued.
This observation  correlated with the finding in  rats of an eightfold  increase
                                                                  q
in the  number  of pulmonary free cells after exposure to 1568 pg/m  (0.8 ppm)
of 0-  for  7 days, but only a  fourfold increase  after exposure for 90 days
(Brummer et al., 1977;  Boorman et al., 1977).   In  similar  studies,  other
authors found  that exposure of  both monkeys and  rats to concentrations as low
           g
as 392 ug/m  (0.2 ppm) of 0, for 8 hr/day on 7 consecutive days resulted in an
accumulation of  macrophages  in  the lungs of these exposed animals (Castleman
et al., 1977;  Dungworth,  1976;  Stephens et al.,  1976).  The data from these
studies suggest  that  these two species of animals are approximately equal in
susceptibility to the short-term effects of CU.
     Thus, the  total  available  data would indicate that,  after short periods
of Og insult, there is a significant reduction in the number of free macrophages
available for  pulmonary  defense, and that these  macrophages are more  fragile,
are less phagocytic,  and have decreased enzymatic activity  (Dowel!  et al.,
1970; Coffin et al.,  1968; Coffin and Gardner,  1972b;  Hurst  et al., 1970).
However, histological studies have reported that with longer exposure periods,
there is an influx and accumulation of macrophages within the airways.  Such a
marked accumulation of macrophages within alveoli may appear to be a reasonable
response to the immediate  insult,  but it has been speculated that the conse-
quences of  this mass  recruitment may also be instrumental in the development
of future  pulmonary  disease  due to the release of proteolytic enzymes by the
AMs (Brain, 1980; Menzel et al., 1983).
     A number  of integrated steps are  involved in phagocytosis processes, the
first being the ability  of the  macrophage to migrate to the foreign substance
on stimulus.   McAllen et al.  (1981) studied the effects  of 1960 ug/m3 (1.0
ppm) of 03  for 4 hr on the migration rate of AMs.  Migration was measured by

                                   9-128

-------
determining the area  macrophages  could clear of  gold-colloid particles that
had been previously precipitated  onto  cover  slips.   In  this study,  it was not
clear whether the  gold was actually ingested or  merely adhered to  the outer
surface of the  cell.   Nevertheless, the cells  from  0~-exposed rats appeared
less mobile,  in that  they migrated 50 percent less than  the  sham-exposed
group.
     A decrease in the ability of macrophages  to phagocytize bacteria after
                                              3
exposure to concentrations  as  low as  980 jjg/m  (0.5 ppm) of 03 for 3 hr was
demonstrated by Coffin et al.  (1968)  using  rabbits.  However, Christman and
Schwartz (1982) may  have demonstrated that with  longer exposure periods,  the
effects may be  different (i.e.,  the phagocytic rate may increase).  In this
study,  rats were exposed to 1568 (jg/m  (0.8 ppm) of 0, for 3,  7 or 20 days;  at
those times the macrophages were isolated,  allowed  to  adhere  to  glass,  and
incubated with carbon-coated latex microspheres.  The percentages  of phagocytic
cells were determined at 0.25, 0.5, I, 2, 4, 8 and  24  hr of incubation.  At
all exposure time periods tested, the  number of spheres engulfed had increased.
The greatest  increase in phagocytic activity was observed after  3 days of
exposure.  The exposed cells engulfed  a greater number of spheres  than controls,
and a  larger  percentage of macrophages from exposed animals was phagocytic.
This enhancement correlated well  with  a significant  increase in cell spreading
of AMs  from exposed  rats as compared to controls.  If longer-term 0-, exposure
enhances macrophage  function and causes a migration of macrophages into the
lung, the  comparisons of function of these new  cells with controls  may not be
valid, because these new cells are biochemically younger.   Another significant
problem with  this  study is  that  the cells  examined were a selected  population
because only  the  cells that adhered to the  glass surface were available for
study.   The cells that did  not adhere  were removed by washing.   In this study,
only 51  percent of the  collected cells adhered after 3 days of Q3  exposure,
compared to 85  percent of  the controls.   Although the  effects of 0- on cell
attachment have not  been studied directly,  there is evidence that 03 affects
AM membranes  involved in the attachment process  (Hadley et al., 1977; Dowell
et al.,  1970;  Aranyi  et al.,  1976; Goldstein et  al., 1977).  The cells most
affected by the cytotoxic action of the 03  exposure might never have been
tested, because they were discarded.
     Goldstein  et  al.  (1977) studied  the effect  of  a 2-hr exposure on the
ability  of AM to be agglutinated by concanavalin-A, a parameter  reflecting
                                   9-129

-------
membrane organization.  A decrease in agglutination of rat AMs was found after
                            o
exposure to 980 or 1960 ug/m  (0.5 or 1.0 ppm) of 03.   A decre>ase in concanavalin-
A agglutinability of  trypsinized  red blood cells obtained from  rats exposed
                     3
for 2 hr to 1960 ug/m  (1 ppm) was also noted.  Hadley et al.  (1977) investigated
                                                        o
AM membrane receptors  from rabbits exposed to 980 ug/m   (0.5 ppm)  of 0, for
3 hr.  Following 0_ exposure, lectin-treated AMs have increased rosette forma-
tion with rabbit red blood cells.   The authors hypothesized that the O^induced
response indicates alterations of  macrophage  membrane receptors  for the wheat
germ agglutinin that  may lead to  changes  in  the recognitive  ability of the
cell.
     Ehrlich et al.  (1979) studied the effects of 0, and N02  mixtures on the
activity of AMs isolated from the  lungs of mice exposed for 1, 2, and 3 months.
                                                        3
Only after  a 3-month  exposure to the  mixture  of  196 ug/m   (0.1 ppm) of 03 and
0.5 ppm  of  NOp (3 hr/day, 5 days/week) did viability in macrophages decrease
significantly.   Iji vitro phagocytic activity was also not affected by a 1-month
exposure to this  level  of pollutants, but after 2- and 3-month exposures the
percentage of macrophages that had phagocytic activity decreased significantly.
     It  has been  reported that the acellular fluid that lines the lungs also
plays an important  role in defense of the lung  through its interaction with
pulmonary macrophages (Gardner et al., 1971; Gardner and Graham, 1977).   These
studies  demonstrated  that  the  protective components of this  acellular fluid
                                                                          o
can  be  inactivated by a  2.5-hr  exposure to 03  levels  as  low as 196 ug/m
(0.1 ppm).   When  normal AM's  are  placed  in  fluid  lavaged from 03-exposed
animals, they showed an increase in lysis (10 percent over control).  A similar
effect was seen when normal AM's were placed in protective fluid that had been
exposed  ui  vitro  to  03.  The data indicate that some of the effects of 03 on
lung cells  may be  mediated through this  lung lining  fluid.   Schwartz and
Christman (1979)  provided  evidence that normal lung lining material enhanced
macrophage migration,  but  the  macrophages obtained from rhesus monkeys after
                                 o
exposure for 7 days to 1568  ug/m  (0.8 ppm) of 03 demonstrated both a decrease
in the  number  of  cells  that migrated  (28 percent of control value)  and in the
total distance  they  traveled (71 percent  of  control  value).   Adding normal
lining fluid to isolated 03~exposed macrophages did enhance the migration, but
it was still significantly less than controls.
     Macrophages  are  rich in lysosomal  enzymes.  Because  these  enzymes  are
crucial  in  the functioning of the macrophage, perturbation of the  metabolic
                                   9-130

-------
or enzymatic mechanisms of these cells may have Important consequences on the
abilities of the  lung  to  defend itself against disease.   Enzymes that have
been identified include acid phosphatase,  acid ribonuclease, beta galactosidase,
beta glucuronidase, cytochrome oxidase, lipase, lysozyme, and protease.   Ozone
decreased significantly the activity  levels of lysozyme, p-glucuronidase, and
acid phosphatase in macrophages after a 3-hr exposure of rabbits to concentra-
                        q
tions as low as 490 |jg/m  (0.25 ppm) of 0- (Hurst et al,, 1970).  Such enzymatic
reductions were also observed in AMs exposed if} vitro (Hurst and Coffin,  1971).
The ability of  0, to alter macrophages1  enzyme activity was also  studied by
means of  unilateral  lung  exposure  of rabbits (Alpert et af., 1971b).   A sig-
nificant  reduction  in  these  same three intracellular enzymes was found to be
specific  to the lung that breathed 0,, rather than a generalized systemic re-
                                                                        3
sponse.   These effects were concentration-related, beginning at 980 ug/m  (0.5
ppm) of  0™.   The extracellular release of such enzymes may occur either as a
result of direct cytotoxic  damage  and leakage of  intracellular contents, or
they may be selectively released without any cell  injury.  Hurst and Coffin (1971)
showed that the reduction in  intracellular lysosomal enzyme  activity  observed
after ID vitro exposure coincides with the release of the enzyme into the sur-
rounding  medium.   In these  studies, the sum of the intra- plus extracellular
enzyme activity did  not equal the total activity,  indicating that the pollutant
itself can  inactivate  the hydrolytic enzyme as well as alter the cell mem-
brane.    Recently,  Witz et al. (1983) and Amoruso et al. (1981) reported that
in vivo  CU  exposure affected  the production of superoxide anion radicals  (CL)
          4                                                                  t
by  rat  AMs. This  oxygen  radical  is  important in antibacterial  activity.
                                           3
Exposure  to concentrations  above 3136 ug/m  (1.6 ppm) of 0™ for 2 hr appears
to result in a progressive decrease in (}„ production.  No statistical evaluation
of the  data was performed.  The  type  of membrane  damage as  well  as the mecha-
nisms by which  this damage is incurred are  not  well understood.  It is not
known .whether the  0,-induced inhibition  of CL... production  arises from the
direct oxidative  damage of  the membrane enzyme involved in the metabolism of
Oy  to  02, or whether  it  is a result  of  oxidative degradation of membrane
lipids that may  serve  a cofactor function.
     Shingu et  al.  (1980)  reported the effects of 0™  on the ability of two
cell types, macrophages and tonsiliar lymphocytes,  to  produce interferon,  a
substance that  aids in defending the  host organism against viral  infections.
                                              3
Macrophages from rabbits  exposed to  1960 ug/m  (1.0 ppm ) of 0,  for  3 hr ex-
hibited  a depression  in  interferon  production.    Interferon production by
                                    9-131

-------
tonsillar lymphocytes was  not  significantly  depressed by 9800 (jg/m   (5.0 ppm)
of 0-.   The authors suggested  that  an impairment of interferon production
might play  an  important  role in the  ability  of the host to combat respiratory
viral i>fortiops.   In neither  of the studies were statistical analyses of the
data reported.   Ibrahim  et al.  (1976)  also exposed mice to 0- and illustrated
              3
that 1568 ug/m  (0.8 ppm) of 03 for a period of 11 days inhibited the in vitro
ability of  tracheal  epithelial  cells to produce  interferon, but  no effect was
observed with alveolar macrophages.
9.3.4.3  Interaction with  Infectious Agents.  In general, the consequences of
any toxic response depend on the particular cell  or organ affected,  the severity
of the damage,  and the  capability of the impaired cells or tissue to recover
from the assault.  , Do  small decrements in the functioning  of these various
host defense mechanisms  compromise the host so  that  it  is  unable to defend
itself against a wide variety of opportunistic pathogens?  It has been sugges-
ted from epidemiological data  that exposure  to ambient levels of oxidants can
enhance the development of respiratory  infection  in humans (Dohan  et al.,
1962; Thompson  et  al.,  1970).   Measurement of the competency of the host's
antimicrobial mechanisms can best be tested by  challenging both the toxic-
exposed animals  and the clean-air exposed control animals  to  an aerosol  of
viable microorganisms.   If the test substance,  such as 03,  had any adverse
influence on the  efficiency of any  of the host's many protective mechanisms
(i.e., physical  clearance via  the mucociliary escalator,  biocidal  activity
mediated through macrophages,  and associated cellular and humoral immunological
events) that would normally function in defense against a microbe, the microbe,
in its attempt to survive, would take advantage of these weaknesses.   A detailed
description of  the infectivity model commonly used  for 03  studies  has been
published elsewhere (Coffin and Gardner,  1972b;  Ehrlich et al.,  1979; Gardner,
1982a).   Briefly,  animals  are  randomly selected  to be exposed to either clean
air or 03.   After the exposure ends, the animals from both chambers are combined
and  exposed to an aerosol of  viable microorganisms.   The  vast  majority of
these studies  have been conducted with Streptococcus sp.   At the termination
of this 15- to 20-min exposure, the animals are housed in clean air, and the
rate of mortality  in the two groups  is  determined  during  a 15-day  holding
period.   In this system, the concentrations of 03 used do not cause any mortal-
ity.   The  mortality in  the control  group (clean air plus  exposure  to the
microorganism)  is  approximately  10  to 20 percent  and reflects  the  natural
                                   9-132

-------
resistance of the  host  to the infectious agent.   The difference in mortality
between the 0~-exposure group and the controls is concentration-related (Gardner,
1982a).  No studies  have  yet been conducted to determine the.lowest possible
number of viable microbes that when deposited in the lung will  cause a pulmon-
ary infection in  the host exposed to 0,.  Miller et al. (1978a) examined the
cumulative data  from nearly  3,000 control mice used  in these infectivity
studies and determined  that  within the  range of 200 to  4,000 colony-forming
units  per  lung,  there is no correlation between number of bacteria deposited
in the lung and the resulting increase in percent mortality.  It must be noted
that in these studies, the control animals are capable of maintaining pulmonary
sterility by inactivating the microbes that are deposited in the lung.
     If the test agent does not impair the host's defense mechanisms, there is
a rapid inactivation of inhaled microorganisms that have been deposited in the
respiratory system.   However,  if the chemical exposure alters the ability of
these  defense cells  to  function,  i.e.,  rate of bacterial killing, the number
of microbes in the lungs  could  increase  (Coffin and Gardner, 1972b;  Miller et
al., 1978; Gardner,  1982a).   This acceleration in  bacterial growth  has  been
attributed to  the pollutant's alteration  of  the  capability of the  lung  to
destroy the inhaled  bacteria, thus permitting those with pathogenic  potential
to multiply and produce respiratory pneumonia.  With this accelerating growth,
there  is an invasion of the blood, and death has been predicted from a positive
blood  culture (Coffin and Gardner, 1972b).
     Coffin et al. (1967) treated  mice with 03 (0.08 ppm and higher  levels for
3 hr)  arid  subsequently  exposed  them  to an aerosol  of  infectious  Streptococcus
sp.  In  this  study,  03 increased  the  animals'  susceptibility  to infection,
resulting  in a significant increase in mortality rate in the O.-treated group.
Ehrlich et al.  (1977),  when  using the  same bacteria but a different  strain of
mice,  found a  similar effect at  0.1 ppm for 3  hr.  When using  CD-I  mice and
Streptococcus sp., Miller et al.  (1978)  studied the effects of a 3-hr exposure
                  3
to 0-j  at  196  MQ/tn  (0.1 ppm) in which the bacterial aerosol was administered
either immediately or 2, 4, or 6  hr after cessation of  the 0- exposure.  For
                                                             «5
these  postexposure challenges,  only the  2-hr time  resulted in  a significant
increase  in mortality (6.7  percent)  over controls.   However, when the  animals
were infected  with streptococci during  the actual  03 exposure, a significant
increase in mortality of  21 percent was  observed.   When  this latter  experimental
                                       3
regimen was used,  exposure to 157  ug/m   (0.08 ppm)  of 0-  resulted in a signifi-
cant increase in  mortality of 5.4  percent.
                                   9-133

-------
     The differences  in  results  among these studies may  have  been due to a
variation in the  sensitivity of the method of 03 monitoring, a difference in
mouse strain,  changes  in the pathogenicity  of the bacteria,  or "differences  in
sample size.   The  results  from such studies that use this  infectivity model
indicate the  model's  sensitivity  for detecting biological  effects  at low
pollutant concentrations and its response to modifications in technique (i.e.,
using different mouse  strains or varying the time  of  bacterial  challenge).
The model  is  supported  by experimental evidence  showing that pollutants
(albeit at different concentrations) that cause an enhancement of mortality in
the  infectivity  system  also cause  reductions  in essential  host  defense
systems, such  as  pulmonary bactericidal capability, the  functioning of the
alveolar macrophage,  and  the  cytological  and  biochemical integrity of the
alveolar macrophage.
     The pulmonary  defenses  in the  (^-treatment  group were  significantly  less
effective in combating the infectious agent to  the  extent  that, even at  low
concentrations, there  were significant increases in mortality over controls.
As  the  03  concentration increased,  mortality  increased.   In some studies,
additive effects  were reported.   These effects, increases  in respiratory
infection,   are supported  by many  mechanistic  studies  discussed  in this
section.  They indicate that 03 does effectively cause a reduction in a number
of  essential host  defenses that would normally play a major role in fighting
pulmonary infections.  Since 1978,  a number of new studies have continued to
confirm  these  previous findings and  improve the existing data base  (Table
9-13).
                                                     o
     When mice were exposed 4 hr to 392 to 1372 pg/m  (0.2 to 0.7 ppm) of 03
and then challenged with  virulent  Klebsiella pneumoniae,  a significant in-
                                           3
crease  in  mortality was noted at  785 (jg/m  (0.4 ppm) 0,  (Bergers et al.,
1983).   Groups of  30 mice  inhaled  approximately 30,  100,  and 300 bacteria/
                    3
mouse.   At  392 ug/m   (0.2 ppm)  of  03,  the 03 group showed  an increase in
mortality,  but it was  not  significantly different from controls.   At 785 pg/m
(0.4 ppm) of 03 a nearly threefold  lowering of the bacterial ID™ value compu-
ted from the  three challenge doses of  bacteria  was  found for  the  03~exposed
group, indicating a significant increase in mortality.   In the same study, the
                                  o
authors  also found that 1372 ug/m   (0.7 ppm)  of 03 for 3  hr  resulted  in  a
decreased ability  of the  lung to clear (bactericidal) inhaled staphylococci.
                                   9-134

-------
                              TABLE 9-13.  EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:  INTERACTIONS WITH INFECTIOUS AGENTS
Ozone
concentration Measurement Exposure
ug/n3 ppm method
157 0.08 NO
157-196 0.08, 0.1 CHEM
196 0.1 CHEM
196 0.1 UV
196, 588 0.1, 0.3 CHEM
T 392-1372 0.2-0.7 G
M
to
Ul
588 0.3 NO
1372-1764 0.7-0.9 NO
1960 1.0 CHEM
2940 1.5 NO
980 0.5 UV
1254 0.64 CHEM
duration and protocol
3 hr
3 hr
3 hr
5 hr/day,
5 days/wk for
103 days
3 hr
3-4 hr
3 hr/day for
2 days
3 hr
3 hr/day,
5 days/wk
for 8 weeks
4 hr/day,
5 days/wk
for 2 months
2-4 wks
4 wks
Observed effects
Increase in mortality to Streptococcus sp.
Significant increase in mortality during 0%
exposure (Streptococcus sp. ).
Increased mortality to Streptococcus sj>.
Increased susceptibility to bacterial infection
(Streptococcus sp. ).
Exercise enhances mortality in
infectivity model system.
Significant increase in mortality following
challenge with aerosol of Klebsiella pneumoniae.
Effect seen at 0.4 ppm.
Enhancement of severity of bacterial pneumonia
(Pasteurella haeinolytica).
Increased susceptibility to infection
(Streptococcus sp. ).
Increase in Mycobacterium tuberculosis lung titers.

No effect on resistance to Hycofaacterium
tuberculosis.

Reduced widespread viral infection of the lung,
resulting in a decrease in disease severity.
No enhancement of the severity of chronic
pulmonary infection with Pseudomonas aeruginosa.
Species
Mouse
Mouse
Mouse
Mouse
Mouse
Mouse
Sheep
Mouse •
Mouse
Mouse
Mouse
Rat
Reference
Coffin et'al. ,
Miller et al.,
Ehrlich et al.
Aranyi et al . ,
11 ling et al. ,
Bergers et al.
Abraham et al.

1967
1978a
!
,; 1977
1983
1980
, 1983
, 1982
Coffin and Blommer, 1970
Thomas et al . ,
Thienas et al.
Wolcott et al.
Sherwood et al
1981b
, 1965
, 1982
;, 1984
Measurement method:   ND = not described; CHEM = gas phase chemiluminescence; UV = UV photometry; G = galvanic meter.

-------
This finding  is consistent with that  of Goldstein et al.  (1971b) and confirms
previous studies  reported  by Miller and Ehrlich (1958)  and Ehrlich (1963).
     Sherwood et  al.  (1984)  established a chronic pulmonary infection  in rats
by inoculating agar  beads  containing viable Pseudomonas aeruginosa (PAO-381)
                                                              3
and then exposed these infected animals for 4 wks to 1254 ug/m  (0.64 ppm) 0^.
The exposure  to  ozone  did not affect the  pulmonary  antibacterial  defense
systems—i.e., no  increase in  number  of organisms cultured  from the lung—but
it did  cause  significant anatomical damages.  The lungs  of  rats exposed to 0~
were  larger  and  heavier when  compared with controls,  and had an increased
number  of  macrophages in  their terminal  bronchioles.   The authors state that
the reason why these results are  different  than  those  described above  for the
"infectivity  model"  is  that  in these later studies, the infective organisms
are given  by  inhalation and are  therefore  deposited throughout the lung and
the pollutant was able to interfere with the initial phase of the host-parasite
interaction.  In  this chronic  study,  the  infection  was isolated to the distal
area of the lung and  it was in its later stage of development when the animals
received the  0«  exposure.   Thus,  the timing of  the exposure to 0~ may be a
significant factor in the impairment of the  lung's antibacterial  defenses.
     Exposure to  ambient levels of  0, (0.5  ppm)  for 2  to 4  wks has been shown
to alter the  pathogenesis  of respiratory  infection  of  mice  with influenza A -
virus (Wolcott et al.,  1982).   The 03-exposed animals  showed a reduction in
severity of the  disease (less  mortality) and an increase in survival time.
The reduction of  disease  severity appears  to be dependent  on the continuous
presence of the DO during  the  infectious process.   This  effect did not corre-
late with  virus,  interferon, or  neutralizing antibody  titers recovered from
the lung,  or with  neutralizing antibody titers in the  sera.   The reduced
disease severity  in  the 0~-exposed animals appears to be due to significant
alterations in the distribution of viral  antigens within the pulmonary tissues,
i.e., less widespread infection of the lung.
     Chiappino and  Vigiani (1982) also reported  that 0^  potentiated pulmonary
infection  in  rats.   In  this study, the investigators  wanted to know  in what
way On  modified the  reactions  to  silica in  specific pathogen-free (SPF) rats.
                                                 3
Silica-treated animals  were  exposed to 1960 ug/m   (1.0  ppm) 0, (8 hr/day, 5
days a  week for up to 1 year)  and housed  in either  an  SPF environment  or  in a
conventional  animal  house, thus being exposed to the bacterial flora  normally
present there.  The  SPF-maintained, 0,-exposed rats showed  a complete  absence

                                   9-136

-------
of saprophytes and pathogens in the lungs, whereas the mlcrobial flora of the
lungs of the conventionally kept rats, also exposed,  consisted of staphylococci
(2500 to 4000  per  gram of tissue) and  streptococci  (300 to 800 per gram of
tissue).  The  lungs  of these rats showed bronchitis, purulent bronchiolitis,
and foci of pneumonia.   The exposure to 03 did not have any effect on particle
retention,  nor did it modify the lungs'  reaction to silica, but it did increase
the animals' susceptibility to respiratory infections.
     Changes in susceptibility  to infection  resulting from 03 exposure  have
also been tested in sheep.  In these studies, sheep were infected by an inocu-
lation  of  Pasteurella  haemolytica either 2 days before  being exposed to 588
    3
ug/m  (0.3  ppm) of 03  or  2 days after the  CL  exposure.   In  both cases, the Q3
exposure was for  3  hr/day for 2  days.   Ozone enhanced the severity of  the
disease (volume of  consolidated lung  tissues), with the greatest effect seen
when the 0, insult  followed the  exposure  to  the bacteria (Abraham et a!.,
1982).  Unfortunately,  only  a  small  number of  animals  were used in each Q3
treatment (n = 3).
     Thomas et al.  (l981b)  studied the effects  of single  and multiple expo-
sures to  0, on  the susceptibility of  mice  to  experimental tuberculosis.
                                3
Multiple exposures  to  1960 ug/m   (1.0 ppm) of 0~ 3 hr/day,  5 days/week for up
to  8 weeks, initiated  7  or 14 days  after the  infectious challenge with
Mycobacten'um  tuberculosis H37RV, resulted in significantly increased bacterial
lung titers, as  compared  with controls.   Exposure to lower concentrations of
03  did  not produce  any significant effects.   In an earlier study, Thienes et
al.  (1965)  reported that  exposure to 2940 ug/m3 (1.5 ppm)  of 03 4  hr/day, 5
days/week  for  2 months also did  not alter the resistance of  mice to M._
tuberculosis H37RV, but he did not measure lung  titers.
     Table  9-14 summarizes a number of studies that used mixtures of pollutants
in  their exposure regimes.   Ehrlich et  al. (1979)  and Ehrlich  (1983) expanded
the  earlier 3-hr-exposure studies to determine  the effects of longer periods
of  exposure to 0- and  N09 mixtures.  In the earlier  studies (Ehrlich et al.,
                                                                           3
1977),  they reported that exposures  to mixtures containing 196  to  980 ug/m
(0.1  to 0.5 ppm)  of-03 and 2920  to 9400 ug/m3  (1.5 to 5 ppm) of N02 produced
an  additive effect  expressed as  an increased susceptibility to streptococcal
pneumonia.   In the  later  studies, mice  were  exposed  3 hr/day,  5  days/week  for
                                        3                             3
up  to  6 months to mixtures  of  196 |jg/m   (0.1 ppm) of 03 and 940 ug/m   (0.5
ppm) of NOp and challenged with bacterial  aerosol.  The  1- or 2-month exposure

                                   9-137

-------
TABLE 9-14.  EFFECTS OF OZONE ON HOST DEFENSE HECBANISHS:   MIXTURES
Ozone
concentration
(jg/B3 pprn
98 0.05
3760

98-196 0.05-0.1
100-400
1500
196 0.1
241-483
V, 196 0.1
f 900
H
00
00
196 0.1
1090

196 0.1
940
196 0.1
940

196 0.1
13200
1040






Pollutant
03 +
MOj,

034
W2 +
ZnS04
03 +
H2S04
03 +
H2S04


03 +
H2S04

0,+
H02
03 +
N02

03 +
S02 +
(NH4)2SO«





Keasurenent3 Exposure
•ethod duration and protocol
CHEH 3 hr


CHEM 3 hr


CHEH 3 hr

CHEH 3 hr +
2 hr


CHEH 3 hr +
2 hr

CHEH 3 months

CHEH 3 hr/day for
1-6 months

UV 5 hr/day,
5 days/wk
for 103 days





h
Observed effects Species Reference
Exposure to Mixtures caused synergistic Mouse Ehrlich et al., 1977,
effect after multiple exposures; additive 1979, 1980
effect after single exposure.
Additive effect of pollutant nixtures House Ehrlich, 1983
with infecti vity node!.

Increased susceptibility to Streptococcus Mouse Grose et al., 1982
pyogenes.
Sequential exposure resulted in signifi- House Gardner et al., 1977
cant increase in respiratory infection.
Neither alone produced a significant
effect. .
Sequential exposure resulted in signifi- Hamster Grose et al., 1980
cant reduction in ciliary beating
activity over H2S04 alone.
Significant decrease in viability of Mouse Ehrlich et al., 1979
alveolar macrophages seen with mixtures.
At 3 and 6 months, susceptibility to Mouse Ehrlich, 1980, 1983
pulmonary infection increased sig-
nificantly. Delayed clearance rate.
Highly signifi cant -increase in Mouse Aranyi et al.s 1983
susceptibility to infection. Effects
attributed to 03. Increased 'bactericidal
rate over 03 alone. Mixture showed greater
growth inhibition in leukemia target cells
and an Increase in blastogenic response to
PHA, Con-A, and alloantigens.

-------
                                     TABLE  9-14.   EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:  MIXTURES   (continued)
10
H
U)
Ozone
concentration
ug/mj ppm
216-784 0.11-0.4
3760-13720 2-7.3
980 0.5
11-3000
1570 0.8
3500
3500
3500
Pol 1 utant
03 +
N02
03 +
Hj,S04
03 +
Fe2(S04)3 +
H2S04 +
(NH4)2S04
Measurement9 Exposure
method duration and protocol
A 17 hr before
bacteria or
4 hr after
bacterial
exposure
UV 3 and 14 days
UV 4 hr
Observed effects Species Reference
Physical removal of bacteria not Mouse Goldstein, E. , et al.,
affected. Bactericidal activity 1974
reduced at higher concentrations.
Significant increase in glycoprotein Rat Last and Cross, 1978
secretion; synergism reported;
effect reversible in clean air.
Exposure to mixtures produced same Rat Phalen et al., 1980
effect as exposure to 03 alone.
Measurement method:   CHEM = gas  phase  chemiluminescence; UV = UV photometry; A = amperometric.
 Abbreviations used:   PHA = phytohemagglutinin;  Con-A = concanavalin-A

-------
did  not  induce any  significant  changes  in susceptibility to streptococcal
infection.  After  3  and  6 months  of exposure, the  resistance to  infection was
significantly reduced.   If the mice were  re-exposed to the 03 and NO, mixture
after the infectious challenge, a significant increase in mortality rate could
be detected 1 month earlier.   The clearance rate of inhaled viable streptococci
from the  lungs  also  became significantly  slower after the 3-month exposure to
this oxidant mixture.
     In more  complex exposure studies,  mice were exposed to  a background
concentration of 188 ug/m  (0.1 ppm) of N09 for 24 hr/day, 7 days/week with a
                                                                            3
superimposed 3-hr  daily  peak (5  days/week) containing a mixture of 196 (jg/m
                            o
(0.1 ppm) of 0- and 940 ug/m  (0.5 ppm) of NOp.   Mortality rates from strepto-
coccal infection were not  altered by 1- and 2-month exposures, but a marked,
although  only  marginally significant  (p  <-0.1), increase was  seen  after a
6-month exposure.
     The  same  laboratory (Aranyi  et al.,  1983) recently reported a  study in
which mice were exposed  5 hr daily, 5 days/week up to  103  days to 0., and a
                                                                    3
mixture of 0,, SO,,, and  (NH4)2SO,.  The concentrations were 196 pg/m  (0.1 ppm)
of 03> 13.2 mg/m   of SOg, and 1.04 mg/m3 of (NH4)2S04>   Both groups showed a
highly significant overall  increase in mortality, compared  to  control  mice
exposed to filtered  air.  However,  the two exposure  groups  did not differ,
indicating that 0, was the major constituent of  the  mixture affecting  the
host's susceptibility to infection.   These  investigators also  measured a
number of specific host defenses to determine if host response td the mixtures
was significantly different from that of the controls or of the 0,-only groups.
Neither the  total  number, differential,  nor  the  ATP  levels of  macrophages
differed  from  controls in either  group, but the complex mixture  did  produce a
significant increase in bactericidal activity over the 03-alone and the control
animals.
     Previous studies (Gardner et a!., 1977) indicated that a sequential expo-
                3                            -3
sure to 196 ug/m   (0.1 ppm) 0~  followed  by  1000  ug/m  H^SO, significantly
increased streptococcal  pneumonia-induced mortality rates in mice.   Ozone and
HgSth  had an  additive effect when  exposed  in  this' sequence.  However,  the
reverse sequence did not affect incidence of mortality.   Sulfuric acid alone
caused no significant effect.  Grose et al. (1982) expanded  these studies and
                                                       3
showed that a  3-hr exposure to the mixture of  196 ug/m  (0.1 ppm) of 0, and
                                   9-140

-------
               3
483 or 241 pg/rn  HUSO, also significantly  increased the percent mortality, as
compared to control.
     The effects of exercise  on the response to  low  levels of 03 were also
                                                                      3
studied by using the infectivity model.  Mice were exposed to 196 |jg/m  (0.1 ppm)
                  3
of 0™  and 588 |jg/m   (0.3 ppm)  of ()„  for  3  hr while exercising.  Each  exposure
level yielded mortality rates that were significantly higher than those observed
in the 03  group  that was not exercised (Illing et al.f 1980).  Such activity
could  change pulmonary  dosimetry,  thus increasing the amount  of  0- reaching
the  respiratory  system.   Thus,  such studies clearly  demonstrate that the
activity level of  the exposed subjects  is  an important concomitant variable
influencing  the  determination of the  lowest effective concentration  of the
pollutant.
9.3.4.4  Immunology.   In addition  to  the  above  nonspecific, nonselective
mechanisms of pulmonary  defense, the respiratory  system also  is provided with
specific, immunologic mechanisms, which  can be  activated  by inhaled antigens.
There  are  two  types of  immune  mechanisms:  " antibody  (humoral)-mediated and
cell-mediated.   Both  serve  to protect the  respiratory tract  against  inhaled
pathogens.    Much  less  information  is available on how 0, reacts  with these
immunological defenses  than  is known about the macrophage  system (see  Table
9-15).   Most studies have involved  the  systemic  immune  system which, to  a
degree,  is compartmentalized from the pulmonary immune system.
     The effects  of 2900 ug/m3 (1.48  ppm)  of  03  for 3 hr  on  cell-mediated
immune  response  were studied by Thomas  et al.  (1981b),  who determined the
cutaneous  delayed   hypersensitivity  reaction to purified  protein derivative
(PPD), expressed  as the diameter of erythemas.   In the guinea pigs  infected
with  inhaled Mycobacterium  tuberculosis, the cutaneous sensitivity to PPD  was
significantly affected by 0,,.  The diameters of the erythemas  from the 0,-exposed
animals  were significantly smaller during  the 4 to 7 weeks  after  the  infectious
challenge, indicating  depressed cell-mediated  immune response.   Exposure  to
980 jjg/m3 (0.5 ppm) of Q3 had  no effect.
     The systemic  immune system was studied by Aranyi et al.  (1983), who  ob-
served the  blastogenic  response of  splenic lymphocytes to  mitogens and allo-
antigens and plaque-forming cells'   response to sheep red blood cells after a
chronic  exposure to 196  ug/m   (0.1 ppm)  of 03 5 hr/day, 5 days/week for 90 days.
No  alteration  in the response to alloantigens  or  the  B-ceTl mitogen  lipopoly-
saccharide  (IPS)  was noted,  but a  statistically  significant  suppression  in
                                    9-141

-------
                                         TABLE 9-15.   EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:   IMMUNOLOGY
Ozone
concentration Measurement Exposure
pi'/i3 ppl method duration and protocol
1% 0.1 UV 5 hr/day,
5 days/wk for
90 days
980, 2940 0,5, 1,5 NO 4 hr
,1 980, 1568 0.5, 0.8 Hast Continuous
*j 3-4 days
980-2900 0.5-1.48 CHEH 3 hr
1150 0.59 'ND 36 days
1568 0.8 NO 1, 3, 7, and
14 days
* Observed effects
Significant suppression in blastogenesis to
T-cell mitogen, PHA, and Con-A. No effect
on B-cell mitogen , LPS, or alloantigen of
splenic lymphocytes.
Attempt to increase immune activity with drug
Levandsole failed.
Increase in number of IgE- and IgA-containing
cells in the lung, resulting in an increase
in anaphylactic sensitivity.
Depressed cell -mediated immunity. No effect at
0.5 ppra for 5 days. Hemaggluti nation antibody
titers increased over control.
Impaired resistance to toxin stress.
Itnaunosuppression.
Depressed splenic lymphocyte response to
T-cell dependent antigen that correlated with
changes in thymus weights.
Species Reference
House Aranyi et al . , 1983
House Goldstein et al., 1978a
House Osebold et al., 1979, 1980
Gershwin et al. , 1981
Guinea pig Thomas et al,, 1981b
Mouse Campbell and Hilsenroth,
1976
Mouse Fujimaki et al., 1984
Measurement method:  ND = not described; CHEH = gas phase chemiluninescence; UV = UV photometry; Mast = Mast meter.
Abbreviations used:  PHA = phytohemagglutinin; con-A = concanavalin-A; LPS = lipopolysaccharide; IgE = imnunoglobulin-E', IgA = inraiunoglobulin-A.

-------
blastogenesis to the T-cell mitogens (PHA and Con-A) was detected.  The authors
suggested that  the  cellular  mechanisms for recognition and proliferation to
IPS and alloantigens were  intact, but  CL might  have been  interfering with the
cellular response to PHA and Con-A through alterations in cell surface receptors
or the binding of specific mitogens to them.   Ozone exposure enhances peritoneal
macrophage cytostasis  to  tumor  target cells.   There was no  effect on the
ability of splenic  lymphocytes to produce  antibodies  against  injected  antigen
(red blood cells).
     The effects  of 0™ on the humoral  immune  response  was  also  studied by
                                                                                3
Fujimaki et al. (1984).  BALB/c mice that were exposed continuously to 1568 pg/m
(0.8 ppm) had  an  increase  in  lung weight after  3,  7,  and  14 days  of exposure.
Spleen weights  were decreased only after  1 and 3 days  of exposure and then
returned  to  normal.   The  thymus weight was decreased  at all time periods
tested, i.e., 1, 3, 7, and 14 days.  Similar mice exposed to 0.4  ppm failed to
show any marked changes in the organ weights.   At 0.8 ppm 03, exposure depressed
the antibody response  to sheep red blood cells  (T-cell dependent  antigen), but
not at the antibody response to the T-cell independent antigen.   These changes
were correlated with the changes of thymus weights.  The authors  also concluded
that 03 affected mainly the T-cell population rather than the B-cell population.
     Campbell and Hilsenroth (1976) used a toxoid immunization-toxin challenge
approach  to  determine  if continuous exposure to  1150 |jg/m   (0.59 ppm)  of 0,
for 36 days  impaired  resistance  to a  toxin stress.  Mice were immunized with
tetanus toxoid  on the  fifth day of 0,,  exposure  and challenged with  the tetanus
toxin  on  day 27.   Compared with controls, the  0,,-exposed animals had greater
mortality and  morbidity following the challenge.  The authors suggested that
the effect was  due  to  immunosuppression.
     Because the  data  indicate that 0~ can alter the  functioning  of pulmonary
macrophages, Goldstein et  al. (1978a)  tried to  counteract this effect by using
a  known immunologic stimulant,  Levamisole,  which protects rodents against
systemic  infections of staphylococci  and  streptococci.   The  purpose  was to
determine if this drug might repair this  dysfunction  and  improve  the  bacteri-
cidal  activity of  the macrophage.   In this study, two concentrations  were
                          3
tested,  2940 and 980 pg/m  (0.5 and 1.5 ppm) of  03 for 4 hr.  In no case did
Levamisole  improve  the bactericidal activity of  the  0~-exposed macrophages.
The cells still  failed to  respond  normally.
                                    9-143

-------
     The possibility that 0~ may be responsible for the enhancement of allergic
sensitization has  important implications for human health effects.  Gershwin
et al.  (1981) reported  that 03  (0.8 and  0.5  ppm for 4 days)  exposure caused  a
thirty-fourfold increase in the number of IgE-containing cells in the lungs of
mice that  had been previously exposed to aerosolized ovalbumin.  In general,
the number of IgE-containing cells correlated positively with levels of anaphy-
lactic sensitivity,  Oxidant damage (0.5 to 0.8 ppm for 4 days) also causes an
Increase in  IgA-containing  cells  in the lungs,and  a  rise in IgA content in
respiratory  secretions  and  accumulation  of lymphoid tissue along the airways
(Osebold et  al.,  1979).   The number of  IgM  and IgG containing  cells did  not
increase.  These  authors  showed that a  significant increase in anaphylactic
sensitivity  occurred when  antigen-stimulated  and O^-exposed  animals  were
compared to  controls (Osebold et al., 1980).   Significantly greater numbers of
animals were allergic  in  experimental  groups when 0- exposure  ranged from 0.8
ppm to 0.5 ppm for 3 days.  The effects observed were most pronounced when the
allergen (ovalbumin) was  administered by I.V.  injection  than  by  the aerosol
route.   Further  studies  are needed to determine the threshold  level of these
effects,

9.3.5  To!erance
     Acclimatization, whether it be a long-term or a moment-to-moment response
of the  organism  to a changing environment, has been a phenomenon  of major in-
terest to  toxicologists for years.  Tolerance, in  the  broadest sense  of  the
word, may  be viewed as  a  special  form  of acclimatization  in  which exposure  to
a chemical  agent results  in increased  resistance, either  partial  or complete,
to the  toxicant  (Hammond and Bellies, 1980).  Often the  terms  tolerance  and
resistance  are used interchangeably.   The word, tolerance,  is primarily  used
when  the  observed decrease  in susceptibility occurs  in  an individual organism
as a  result  of its own previous or continuing exposure to the particular  toxi-
cant  or to some other related stimulus.  Resistance generally refers to relative
insusceptibility that is genetically determined (Hayes, 1975).
     A  third term, adaptation,  has been  widely used primarily to  describe the
diminution  of  response  seen in human subjects who have undergone  repeated 0,
exposure  (Chapter 10,  Section 3).  This adaptation might well  result  from  a
different  biologic process  than that referred to in the  various  animal tole-
rance studies.   It is not  yet  known  whether the laboratory animal develops
                                    9-144

-------
adaptive responses similar to those seen in humans (i.e., respiratory mechani-
cal  functions,  symptoms of respiratory  irritation,  and airway reactivity).
Thus, to  date,  the  precise  distinction or definition  of  these  two terms,
tolerance and  adaptation, are  not yet  fully  understood.   There are also
limited data regarding  the ability of cells in  these "adapted"  respiratory
systems to once again return to a pre-exposure condition following termination
of the  exposure.   However, Plopper et al. (1978)  found that  after  6 days  in
clean air, the  rat's lungs were almost  recovered  from the damage caused by
3 days  of exposure  to 0.8 ppm 0,,  and  that  these "new  recovered" cells have
approximately the  same  degree of  susceptibility to  a re-exposure of 0- as
their pre-exposed counterparts, i.e., fully susceptible to re-exposure.   These
morphological  findings  confirm  similar  conclusions  based on concomitant
biochemical studies of Chow et al.  (1976b).
     In animal  oxidant  toxicity studies, the  term tolerance  classically is
defined  as  the  phenomenon wherein  a  previous  exposure  to  a nonlethal
concentration of 0, will provide some protection against a subsequent exposure
to a concentration  of  0^ expected to be lethal.  The degree of tolerance
depends considerably  on the  duration of  the exposure and  the concentration.
Tolerance occurs  rapidly  and  can  persist  for several weeks  (Mustafa  and
Tierney,  1978).   The term tolerance should not be  considered to indicate
complete  or  absolute protection, because continuing  injury does  occur and  can
eventually  lead to  nonreversible  morphological  changes.   This  protective
phenomenon seen with oxidants was  originally  described by Laqueur and Magnus
(1921)  in cats undergoing  exposure to phosgene.
     In the  typical  experiment,  animals  are  pre-exposed to a  lower  concentra-
tion of 0, and  then challenged at  a  later time to  a  higher concentration.  As
early  as  1956, Stokinger  et  al. presented  data  clearly indicating that an
animal  could also become tolerant  to the  lethal effects of 0-.  Such tolerance
has  also  been  reported by many investigators,  including Matzen  (1957a),
Mendenhall and  Stokinger  (1959), Henschler  (1960), and Fairchild (1967).   The
observation of this tolerance phenomenon  in experimentally exposed animals has
led  to  the speculation  that it may also  be a mechanism  for protecting environ-
mentally  exposed  humans.   Tolerance  to  0,  also provides cross-protection
against the  pulmonary effects of  other  chemical agents, such  as N02, ketene,
phosgene, and  hydrogen  peroxide (Stokinger and Coffin, 1968)  and recently to
hyperoxia (Jackson and  Frank, 1984).
                                    9-145

-------
     The previous criteria  document for 0, and other photochemical oxidants
cited various  studies  that  examined 0- tolerance and presented some evidence
                                      «3       —
indicating possible mechanisms of action.  Review of these earlier data reveals
that pre-exposure to  a certain concentration of 0~ can protect test animals
from the acute lethal effects of a second exposure to 0-.   This protection has
been attributed to a significant reduction in pulmonary edema in the pre-exposed
animals.  Table 9-16 lists the key studies on 03 tolerance.
     Because 0,  has  a marked proclivity to  reduce  the ability of alveolar
macrophages to function, studies were conducted to determine how the pulmonary
defense  system in tolerant animals  compared  with  naive animals.   With the
bacterial  infectivity  model-  (Section 9.3.4.3),  the  pre-exposed  (tolerant)
animals were only partially protected from the aerosol infectious challenge
(Coffin and Gardner, 1972a;  Gardner and Graham, 1977).   The partial protection
was evident at 0, concentrations that had been shown to be edemagenic;  however,
                                     3
at the lowest concentration, 200 pg/m  (0.10 ppm) of 0-, there was no signifi-
cant difference  imparted  by the use of the tolerant-eliciting exposure.   The
data suggest  that at  the higher concentrations  (>  0.3 ppm),  pre-exposure
prevented  edema,  which prophylactically aided the animals'  defenses against
the inhaled microorganisms.   Because the protection was  only  fractional  and
did not  occur  at the  lowest level,  however, 03  still suppressed specific  body
defenses that were not protected by the phenomenon of tolerance.
     To  further  investigate this hypothesis (Alpert and Lewis, 1971; Gardner
et a!.,  1972), studies were conducted to evaluate the effects  of-tolerance at
the cellular level.  These  studies  indicated  that the initial  0,  exposure did
induce tolerance  against  pulmonary  edema in the exposed lung; however, there
was no protection afforded against the cytotoxic effects of Q3 at the cellular
level.    The cytological toxic injuries  measured in this study  (including  sig-
nificant reductions  in enzymatic activities of  macrophages and an increase in
inflammation,  as measured by the presence  of poJlymorphonuclear  leukocytes)
showed that there was  no protection against these cellular defense mechanisms.
     Frager et al. (1979) studied the possibility of tolerance to 0, in mucocil-
iary clearance.   Exposure of rats to 1.2 ppm of 0, following particle deposition
caused a substantial  delay  in mucociliary clearance.  The CL effect could be
                                          3       -
eliminated by  a  pre-exposure to 1600 pg/m  (0.80 ppm) of 03 for 4 hr,  3 days
before  the deposition  of the particles.   Thus,  the pre-exposure provided
complete protection  against the higher  03  level  that  lasted for about one

                                   9-146

-------
TABLE 9-16.   TOLERANCE TO OZONE
Ozone
(ug/m3)
p re-
exposure
196-1960
196
490
M3
jL 98°
*»
588
588
588-980
980-1960
Ozone
(ppm)
pre-
exposure
0.1-1.0
0.1
0.25
0.5
0.3
0.3
0.3-0.5
0.5, 1.0
Ozone Ozone
(Mg/m3) (ppm)
Length of after after
pre- latent latent
exposure period period
3 hr 196- 0.1-1.0
1960
30 min 196 0.1
6 hr 1966 1
6 hr
1 hr 39,200 20
3 hr 588 0.3
4 days 980 0.5
1372 0.7 1.0
1960
3 hr 43,120 22
Length of
exposure
after
latent
period
3 hr
30 min
6 hr
2 hr
3 hr
I, 2, 4
days
3 hr
Observed effect(s) Species Reference
Lower mortality for pre-exposed mice than Mouse Gardner and Graham, 1977
• mice receiving only one 03 dose. Complete
tolerance was not evident.
Tolerance exhibited in the lungs' periphery, Dog Gertner et a!., 1983b
as measured by collateral resistance.
Response < controls in tolerant animals.
No tolerance to edema unless pretreated Rat Alpert et a!,, 1971a
with methylprednisolone.
Edema as measured by recovery of 132I
in pulmonary lavage fluid.
Tolerance to edema effects of Oa did not House Gregory et al. , 1967
develop in thymectomized animals but
developed in sham-operated animals, in-
dicating the thymus may be involved in
tolerance.
20% lower mortality for pre-exposed mice than Mouse Coffin and Gardner, 1972a
mice receiving only one 03 dose. Partial
tolerance probably due to inhibition of edema-
genesis.
Lack of total protection indicated by increased Rat Evans et al., 1971, 1976a,b
numbers of type 2 cells.
With unilateral lung exposure technique, Rabbit • Alpert et al., 1971b
tolerance to edema occurred as a local Alpert and Lewis, 1971
effect (cellular) and was seen only in
the pre-exposed lung.

-------
                                                           TABLE 9-16.  TOLERANCE TO OZONE  (continued)
Ozone Ozone Length of
Ozone Ozone (tig/a3) (pp») exposure
((jg/n3) (pp») Length of after after after
pre- pre- pre- latent latent latent
exposure exposure exposure period period period
980 0.5 3 hr 5880 or 3 and 22 3 hr
1470 0.75 3 days 7840 4.0 8 hr
1490 0.76 3 day 6860-7840 3.5-4 8 hr
I
H
CO
1S70 0.8 3 days 1570 0.8 6 or 27
days
1600 0.8 4 hr 2352 1.2 4 hr
1960 1 " 1'hr NO KO ND
I960 1 1 hr 3920 2 1 hr
Observed effect(s) Species
With unilateral lung exposure technique, Rabbit
tolerance developed only to pulmonary edema.
No tolerance to the chemotaxis of polymorpho-
nuclear leukocytes or decreased lysosomal
hydrolase enzyme activity.
A smaller decrease in activities of glutathione Rat
peroxidase, glutathione reductase, glucose-6-
phosphate dehydrogenase and levels of reduced
glutathione in lungs of tolerant animals, as
compared to nontolerant animals.
When latent period was 11 days, no tolerance Rat
to decrease in 6SH peroxidase system immedi-
ately afttr challenge; 18 hr later, a smaller
decrease occurred. When latent period was
19 days, the decrease in enzyme activities
measured 16 hr post-challenge was less in pre-
exposed animals; 114 hr post-challenge, some
increases in the GSH peroxidase system were
observed.
After 6 days of recovery the lung is again Rat
fully susceptible to re-exposure. Adaptation
lasts only as long as the 0, exposure
continues.
Pre-exposure to 03 caused complete tolerance to Rat
delay in mucociliary clearance at 3 days, but
not 13 days.
All animals X-irradiated to 800 R. 60% of House
Qj-pre-exposed mice survived. 100% of
controls died.
Tolerance to allergic response to inhaled Guinea
acetylcholine. pig
Reference
Gardner et al . , 1972
Chow, 1976
Chow et al . , 1976b
Chow,- 1984
Plopper et al. , 1978
Frager et al. , 1979
Hattori et al., 1963
Matsumura et al . , 1972
ND = not described.

-------
week.  The possible mechanism for this protection could be a thickening of the
mucus layer, which would offer the epithelium an extra physical barrier against
Q~.  As the  secretion  returns to  normal, the protection  is  lost.  The  authors
suggested that  another possible mechanism for  this  protection involves the
ciliated cells  and  their cilia.   In this case,  the  protection could result
from either the formation of intermediate cilia (Hilding and Hilding, 1966) or
the  occurrence  of  some other temporary change  in  the regenerating  ciliated
cell.
     Evans et al. (1971, 1976b) also measured tolerance by studying the kinetics
of alveolar  cell  division in rats during a period of exposure to an elevated
                                    3
CL concentration of 980 or 1372 ug/m  (0.50 or 0.70 ppm, up to four days) that
                                                              3
followed initial exposure at a lower concentration of 686 ug/m  (0.35 ppm) for
four days.   Tolerance  in this case was the ability  of  type 1  cells  to with-
stand a second  exposure without any increase  in the number of type  2  cells,
which would  indicate a lack of complete tolerance.   Similar to the host defense
studies cited above, these investigations showed that tolerance to the  initial
concentration of  03 did not  ensure  complete protection  against re-exposure  to
the  higher 03 concentration.
     Attempts have  been made to explain tolerance by examining the morphological
changes that occur  due to repeated exposures  to  0^.  In these studies the
investigators  attempt  to assess  various  structural  responses with  various
exposure profiles  and  concentrations.   Dungworth et  al.  (1975b) and  Castleman
et al. (1980) studied  the repair  rate of 0,, damage as indicated by DNA  synthe-
sis.  These  effects are fully described in  Section 9.3.1.2, and they indicate
that with  continuous exposure to 0,,,  the  lung  attempts  to initiate the repair
of  the  0.,  lesion,  resulting  in  somewhat  reduced or  less than  expected total
damage.  These  authors suggest that this is an indication that although the
damage is continuing,  it  is at a  lower  rate, and they refer to this phenomenon
as adaptation.
     It has  been suggested that the tolerance  to edema  seen in animal  studies
can  be explained  through  the  indirect evidence  that  more  resistant cells, such
as  type  2 cells, may  replace the more  sensitive,  older  type 1 cells,  or  that
the  type  2 cells may  transform  to  younger,  more resistant cells  of  the same
type (Mustafa and Tierney,  1978).  A number of workers  have reported that the
younger type 1 cells are relatively more resistant to the subsequent toxicity
of 0-  (Evans et  al.,  1976a;  Dungworth  et al.,  1975a; Schwartz et al.,  1976).

                                    9-149

-------
Thus, there is also the possibility that this reparative-proliterative response
relines the airway epithelium with cells that have a biochemical armamentarium
more resistant to  oxidative stress (Mustafa et  a!.,  1977;  Mustafa and Lee,
1976),
     Another suggestion  is  that with  CL  exposure,  there  is  cellular accumula-
tion within the  airways  resulting  in  mounds  of cells  in  the terminal  bronchi-
oles that  may  cause  considerable narrowing  of the airways  (Berliner  et a!.,
1978).  As the airways become more obstructed, the 0- molecules are le'ss likely
to penetrate to  lumen.   This may result in a "filtering" system that removes
the 0~ before it reaches the sensitive tissue.
     Tolerance to  03  has also been studied by using a variety of biochemical
indicators to  measure the extent to  which a pre-exposure to 03 protects or
reduces the host response to  a subsequent exposure.   For example, Jackson and
Frank (1984) found that preexposure to 0, produced cross tolerance to hypoxia.
In these tolerant animals (0.8 ppm 03 x 7 days) there was a significant increase
in total lung superoxide dismutase, glutathione peroxidase, glucose-6-phosphate
dehydrogenase, and catalase.   Such an increase  in these antioxidant  enzymes
occurs with 03 exposure.  Chow et al.  (1976a,b) compared a variety of metabolic
activities of  the  lung immediately after an  initial 3-day continuous  exposure
            2
to 1600 |jg/m  (0.80 ppm) of 03 with the response after subsequent re-exposure.
At 6, 13,  and 27 days after  the  pre-exposure  ended,  the animals were once
again treated  to the  same exposure routine.   If tolerant, the animals should
have shown a  diminution of response.   However, the re-exposed rats responded
similarly  to those animals  tested after the initial  exposure.   The lungs of
the naive animals had equivalently higher activities of glutathione peroxidase,
glutathione reductase,  glucose-6-phosphate dehydrogenase and higher levels  of
nonprotein sulfhydryl  than  controls and were comparable  to the  animals that
were exposed and tested immediately  after the initial exposure.  The authors
state that this  indicated that by the time recovery from the pre-exposure is
complete,  the  lung is as susceptible  to  the  re-exposure  injury  as a lung that
has never been exposed.
     In a follow-up study, Chow (1984) pre-exposed rats for 3 days (apparently
                                 3
continuously) to air  or 1490 ug/m  (0.76 ppm) 03 and challenged them at various
times with an 8-hr exposure to a higher level of 0~.   As expected, the pre-expo-
                                                                    o
sure protected the rats  from  the  lethal  effects  of 6860  to  7840  ug/m   (3.5  to
4.0 ppm) 0-, whether  the challenge was  8, 11, or  19  days later.  Generally,

                                   9-150

-------
all 0-, exposures decreased the GSH peroxldase system.  When rats were challenged
              3
with 7644 ug/m   (3.9 ppm)  0~  11 days after the 3-day pre-exposure, there was .
no tolerance immediately after the challenge exposure; 18 hr after the challenge
exposure, a dampening of the decrease in the GSH peroxidase system was observed.
                                               o
When the pre-exposure  and challenge (6860 (jg/m , 3.5 ppm)  were  separated  by
19 days, the decrease in the GSH peroxidase,system measured 16 hr post-challenge
was less in  the  animals  receiving  pre-exposure  (except  for  GSH  reductase,  for
which values were equivalent); 114 hr post-challenge, no tolerance was observed,
but some increases in enzyme activity were observed.
     Gertner et  al.  (1983b) presented data showing  that  the development of
adaptation and tolerance to pulmonary function changes is rapid and mediated
through the vagus nerve.  These investigators used a bronchoscope to expose an
isolated segmental airway of the  lung to  03  and  study  changes  in  collateral
resistance (Rcoll).  During a 30-min exposure to 0.1 ppm, the Rcoll increased
31.5 percent within  2  min and then  gradually decreased to control level in
spite  of continual  exposure to 0~.   Fifteen  minutes after the  03  exposure
ceased,  the  Rcoll  returned to normal.   Subsequent  exposure to  0.1 ppm of  0«
did not  increase Rcolls  indicating that  some protection existed.  These investi-
gators  have  tried  to distinguish  between  the  terms  adaptation  and tolerance
based on these studies.  They used  adaptation to describe the pattern of changes
that occur during continuous exposure to 0., and the term tolerance to  describe
resistance to subsequent 0., exposure.
     Thus, the  available'evidence  from  animal  studies suggests  that tolerance
does not develop to all  forms of lung injury.  The protection described against
edemagenic  effects  of 03  does not appear to offer complete protection, as
illustrated by the following examples.

     1.   There  is  no  tolerance (i.e.,  no protection occurs)  on the part  of
          the specific pulmonary defense mechanisms  against bacterial  infection
          below  the edemagenic concentration; whereas above the  edema-inducing
          concentration  the effect of tolerance (i.e.,  inhibition  of pulmonary
          edema), can  lower the expected mortality rate because  the animals do
          not  have to cope with the additional  burden of  the  edema fluid.
     2.   Specific cellular functions of the  alveolar macrophage (i.e., enzyme
          activity) are  incapable  of being protected  by tolerance.
                                    9-151

-------
     3.   Various biochemical responses were found in both naive and pre-exposed
          animals.
     4.   Tolerance fails to inhibit the influx of polymorphonuclear leukocytes
          into the airway.

     This last finding  is  interesting  considering the effective tolerance for
edema  production.   This suggests that the chemotactic  effect  of  0, may be
separable from  the edemagenic  effect.   This may also  explain  why chronic
morphological changes in  the lung may occur after  long-term  exposure,  even
though there may not be any edema.
     The possible explanations for this tolerance phenomenon have been proposed
by Mustafa  and Tierney  (1978).   The primary mechanism of  tolerance may not be
due to  hormonal  or neurogenic pathways, because  unilateral  lung exposure does
not result  in tolerance of  the  nonpre-exposed  lung (Gardner et a!., 1972;
Alpert  and  Lewis,  1971).   But it should be noted that Gertner et al. (1983b)
have evidence that local  tolerance may involve  a neural  reflex.  Changes in
Rcoll   may  be mediated  through  the vagus  nerve.  After bilateral  cervical
vagotomy, the resistance  did not increase during Q~  exposure  but  did after
challenging  with histamine,  indicating that the parasympathetic  system  may
play a  role in  response to  Q~  in the  periphery of the lung.   There is some
evidence that 03 may  cause a decrease  in  cellular  sensitivity, an increased
capacity to  destroy the test chemical, or the  repair of the injured tissue
(Mustafa and Tierney, 1978).   In addition, 03 could possibly cause  anatomic
changes, such as an increase in mucus thickness, that may, in effect, reduce
the dose of  0- reaching the gas-exchange areas of the lung.
     It should be mentioned that the term tolerance carries with it the conno-
tation  that  some form of an insult and/or damage has occurred and there has
been an overt response at the structural and/or  functional level.   The response
may be  attenuated or  undetectable, but the basis for the  establishment of the
tolerance still  persists.   It  is possible that  the cost for tolerance may be
minor,  such  as a slight increase  in mucus secretion; however, one must also be
aware that changes in response to diverse kinds  of insults to a host's system,
such as the  immune system, are adaptations that  might even suggest an undesirable
effect  of ambient oxidant air pollution.
                                   9-152

-------
9.4  EXTRAPULMONARY EFFECTS OF OZONE
9.4.1  Central Nervous System and Behavioral Effects
     Despite  reports  of headache,  dizziness,  and  irritation  of the nose,
throat, and chest  in  humans exposed to 03 (see Chapter 10), and the possible
implications of these and other symptoms as indications of low- level Q3 effects,
few recent  reports  were found on behavioral and other central nervous system
(CNS) effects  of  CL exposure in animals.   Table  9-17 summarizes studies on
avoidance and conditioned behavior, motor activity, and CNS effects.
     Early  investigations  have  reported  effects of 0- on  behavior patterns  in
animals.  Peterson  and  Andrews  (1963)  attempted to characterize  the avoidance
behavior of mice to 0- by measuring "their reaction to a 30-min exposure on one
                     . O
side of an  annular plastic mouse chamber.  A concentration-related avoidance
                                                   3
of the  0- side was  reported at  1176  to 16,660  |jg/m  (0.60 to  8.50 ppm) of 0-.
However, the  study  had  serious shortcomings,  including a lack of positionT
reversal controls  (Wood,  1979), considerable  intersubject variability,  and
other design  flaws  (Doty,  1975).   Tepper et al. (1983)  expanded  on  the design
by using inhalant  escape behavior to  assess, directly the aversive properties
of Og.  Mice were individually expo.sed to 03 for a maximum of 60 sec, followed
by a  chamber  washout period of 60 sec.  The animals  could terminate exposure
                                                   *
by poking their noses into  only one of two brass conical  recesses containing a
                                    3
photobeam.   The delivery of 980 |jg/m   (0.50 ppm) of 03 was reliably turned off
for a  greater proportion of experimental trials,  compared to control trials
                                   3
with  filtered air.   At  19,600 (jg/m  (10  ppm)  of  O, all  animals turned off
100 percent of  the  trials with an average  latency of approximately 10 sec.
     Studies by Murphy et al.  (1964)  demonstrated that wheel-running activity
decreased by approximately 50 percent when mice were exposed to 392 to 980 pg/m
(0.20 to  0.50 ppm)  of 0-  for 6  hr and decreased to 60 percent  of  pre-exposure
                                                                  3
values during the first 2 days  of continuous exposure to 588 ug/m  (0.30  ppm)
of 0,,.   Running activity  gradually returned to pre-exposure values during the
                                               3
next 5 days of  continuous exposure to 588 jag/m  (0.30 ppm)  of  0.,.  If the  same
                                           3
mice were subsequently  exposed to  1372 ug/m   (0.70 ppm)  of 0,  for  an additional
7 days,  running activity  was depressed to 20  percent  of pre-exposure values.
                                                                              3
Partial  recovery was  described during the final days of  exposure to 1372 yg/m
(0.70 ppm), and complete  recovery occurred  several  days after exposure was
terminated.  However, partial tolerance was  seen when the air-control mice
                                      3
were subsequently exposed to  392 |jg/m  (0.20  ppm)  of 03  for 7  days.  Konigsberg
and  Bachman  (1970)  used  a capacitance-sensing  device to record the motor
                                   9-153

-------
                                            TABLE 9-17.  CENTRAL NERVOUS SYSTEM ABO BEHAVIORAL EFFECTS OF
M
cn
Ozone
concentration
u§/«3 ppi
110 0.056
98-1960 0.05-1.0
196-3920 0.1-2.0
235-1960 0.12-1.0
392-980 0.2-0.5
588-1372 0.3, 0.7
980-19600 0.5-10.0
980 0.5
Measurement3 * Exposure
method duration and protocol Observed effects(s) Species
d 93 days, continuous No overt behavioral changes. Cholines- Rat
terase activity inhibited at 75 days of
exposure, returning to control levels
12 days after termination of exposure.
HAST 45 min Motor activity progressively decreased Rat
with increasing 03 concentrations up to
0,5 ppa. Slight Increase in frequency
of 3-nin intervals without aotor activity.
CHEH G hr Linear and/or nonotonic decreases in Rat
operant behavior during exposure.
CHEH 6 hr Wheel running activity decreased Rat
iwmotonically with increasing Oa con-
centration. Components of running
were differentially affected at low
vs. high Os concentrations.
NBKI 6 hr Wheel running activity decreased 50%. House
MAST 7 days, continuous Running activity decreased 60% during
first 2 days, returning to control
levels during the next 5 days of expo-
sure; running activity decreased 20X
when 0.3 ppm exposure was followed
immediately by 0.7 ppm 03 exposure.
Adaptation with continued exposure
was apparent.
CHEM 60 s Exposure terminated by nose pokes with Mouse
increasing frequency as Os concentra-
tion increased.
NO 30 rain Elevation of simple and choice reactive Nonhunan
time, priaate
Reference
Eglite, 1968
Konigsberg and
Bachflian, 1970
Weiss et al.,
1981
Tepper et al . ,
1982
Murphy et al . ,
1964
Tepper et al . ,
1983
Reynolds and
Chaffee, 1970

-------
                                    TABLE 9-17.   CENTRAL NERVOUS SYSTEM AND  BEHAVIORAL EFFECTS OF  OZONE   (continued)
V£>
H
Ul
Ul
Ozone .
concentration Measurement '
(jg/m-1 ppm method
980-19SO 0.5, 1.0 NBKI
1176- 0.6-8.5 e
16,660
1960 1.0 HAST
NBKI
1960- 1-3 ND
5880
1960 1
1960- 1-3 ND
5880
Exposure
duration and protocol
1 hr
30 min
7 days, continuous
18 months,
8 hr/day
18 months;
8, 6, 24 hr/day
18 months,
8 hr/day
Observed effects(s)c Species
Evoked response to light flashes in the Rat
visual cortex and superior collicus
decreased after exposure.
Avoidance behavior increased with Mouse
increasing 03 concentration.
Reduction in wheel running activity; no Rat
effect of Vitamin E deficiency or
supplementation.
COMT activity decreased at 2 ppm, MAO Dog
activity increased at 1 ppm only.
COMT activity decreased as the daily
exposure increased from 8 to 24 hr.
MAO activity increased at 8 and 16
hr/day and decreased at 24 hr/day.
Alterations in EEG patterns after Dog
9 months, but not after 18 months of
exposure.
Reference
Xintaras et al. ,
1966
Peterson and
Andrews, 1963
Fletcher and
Tappel, 1973
Trains et al . ,
1972
Johnson et al . ,
1976
Measurement method:   MAST = Kl-coulometric (Mast meter);  CHEM
 described.
 Calibration method:  NBKI = neutral  buffered potassium  iodide.
^Abbreviations used:   COMT = catechol-o-methyltransferase;  MAO
 Spectrophotometric method with dihydroacridine.
eKI titration with sodium thiosulfate.
= gas phase chemiluminescence; NBKI = neutral buffered potassium iodide; ND = not
= monamine oxidase; EEG = electroencephalogram

-------
activity of rats during a 45-min exposure to 98, 196, 392, 980, and 1960 ug/m
(0.05,  0.10,  0.20,  0.50, and 1.0 ppm)  of  03<   Compared with control rats,
motor activity  following 0, exposure progressively decreased with increasing
                                3
0, concentrations up to 980 ug/m  (0.50 ppm).  No greater reduction was obtained
            3
at 1960 ug/m  (1.0 ppm).  In addition, the frequency of 3-min intervals without
measurable motor activity tended  to increase slightly (from 1  to  1.25 to
approximately 3) with increasing 03 concentration.
     A  detailed raicroanalysis of  motor activity was  undertaken by Tepper
et al. (1982), who exposed rats for 6-hr periods during the nocturnal phase of
their light-dark cycle  to 235, 490,  980, and 1960 pg/m3 (0.12, 0.25, 0.50  and
1.0 ppm) of Og.  The 3 days preceding an exposure were used for control obser-
vations to measure  running  activity for each  rat in a wheel attached to the
                                                                     3
home cage.  Decreases in wheel running activity occurred at  235 ug/m  (0.12 ppm)
and progressively greater  decreases in wheel  running activity occurred with
increasing 0-  concentration.  An analysis  of the  running  behavior showed that
the components  of running were differentially  affected by On.  An increase in
the time interval between running bursts primarily accounted for the decreased
motor activity at the low (235 ug/m  , 0.12 ppm) 03 concentration.  Postexposure
increases in  wheel  running  were seen following this  low 0, concentration.   At
                                    3
higher  03 concentrations (>490 pg/m, 0.25 ppm),  an increase in the time per
wheel revolution, a decrease  in the burst  length  as well  as the  extended time
interval  between bursts contributed  to the reduced motor activity.  These
higher concentrations also caused a decrease in performance compared to control
for several hours after exposure was terminated.
     Effects of 03 on behavior were further  investigated by Weiss et al. (1981)
in their studies on the operant behavior of  rats during 03 exposure.  The term
operant refers  to  learned behaviors that are controlled by subsequent events
such as food  or shock delivery.  In this case,  rats were  trained to  perform a
bar-pressing  response maintained by  a reward  with  food  pellets delivered
according to  a 5-min fixed-interval  reinforcement schedule.   The  rats were
                                                           3
exposed for 6 hr to 03 concentrations from 196 to 3920 ug/m  (0.10 to 2.0 ppm),
with at least 6  days separating successive exposures.  Two groups of rats were
tested, one beginning in the morning and the other in the mid afternoon.  Ozone-
                                                   3
induced decreases were linear from 196 to 2744 ug/m  (0.10 to 1.40 ppm) for the
first group; for the second, the decreases were generally monotonic from 196 to
         o
3920 ug/m  (0.10 to 2.0 ppm).  Analysis of the distribution of responses during
                                   9-156

-------
the  various  0- exposures  indicated concentration-related  decreases  arose
mainly from  the later portions of  the  sessions and that the  onset of the
decline in response occurred earlier at the higher 0- concentrations.   In con-
trast to  other  types  of toxicants,  0-  did  not disrupt the temporal pattern
that characterized response during each fixed-interval presentation.  Based on
the sedentary nature  of the task, the authors suggested that the inclination
to respond rather  than the physiological capacity  to  respond  was impaired.
     Kulle and  Cooper (1975)  studied  the effects  of  0-  on the electrical
                                                                           3
activity  of  the nasopalatine  nerve  in  rats.   Ozone exposure to  9800  ug/m
(5 ppm) for  1 hr  produced an  increase in nasopalatine nerve response (action
potential frequency) to amyl alcohol, suggesting that the nerve receptors were
made more sensitive by prior exposure to 03.  One-hour air perfusion following
the Og exposure reduced the neural  response to amyl alcohol, but not to pre-
exposure  levels.   The nasopalatine  nerve is a branch,of the trigeminal nerve
which responds  to  airborne chemical irritants.  Because  most  irritants,  in-
cluding  Qg,  also  have  odorant properties  and,  therefore,   stimulate  both
trigeminal and  olfactory  receptors  in the  nasal mucosa,  it is difficult to
distinguish  an  irritant response  from an odor  response as  the  mechanism  for
behavioral effects in laboratory animals.
     Effects of 0_ on the CNS  have been reported.   Trams et al. (1972) measured
biochemical  changes  in the cerebral cortex of  dogs exposed for 18  months to
                         •3
1960, 3920,  or  5880 ug/m  (1, 2, or 3  ppm) of 0,.   In 8 hr/day  exposures,
                                    ,             *3
reported  decreases (35 percent)  in the  catecholamines  norepinephrine and
epinephrine  were  not  statistically  significant, although 0- exposure at 3920
    3
ug/m  (2  ppm) caused  a statistically significant decrease in catechol-o-methyl-
transferase  (COMT) activity.   In contrast, monamine  oxidase  (MAO) activity was
                                    •3
significantly elevated  at 1960 ug/m  (1 ppm) of 0.,,  but  not at 3920  or 5880
    3                                                               3
pg/m   (2  or  3 ppm) of  03.   Increasing  daily  exposures to 1960  pg/m  (1 ppm)
from 8  to 24 hr/day  caused  a  significant decrease  in COMT  activity,  but  MAO
activity  increased at 8 and  16 hr/day but decreased at 24 hr/day.   Concurrently,
Johnson  et al.  (1976) measured electroencephalographic (EEC) patterns in the
same dogs and noted alterations  in  EEC  patterns after 9 months  of exposure to
                  3
1960 to  5800 ug/m  (1 to 3  ppm)  of 0-, but not after 18  months of exposure.
The  authors  noted that it was difficult to correlate the  observed EEG changes
with the  alterations  of metabolic balance described.   Furthermore,  it was  even
more difficult  to assess the  metabolic  and physiologic  significance of  the
changes without more  information abcut  chronic  0, exposure.
                                    9-157

-------
9.4.2  Cardiovascular Effects
     Very few reports  on the cardiovascular effects  of  03  and other photo-
chemical oxidants  in animals have been  published.   Brinkman et al.  (1964)
studied  structural  changes in the cell  membranes  and nuclei of myocardial
muscle fibers in  adult mice exposed to  0-.  After  a 3-week exposure  to  392
    3
ug/m  (0.20 ppm) of  0~  for 5  hr/day, structural changes  were  noted, but  these
effects were reversible about 1 month  after exposure.  However, because this
study had severe  design and methodology limitations,  the  results  should be
considered questionable until independently verified.
     Bloch et al.  (1971)  studied  the  effects of 03 on  pulmonary arterial
pressure in dogs.  They exposed 31 dogs to  1.0 ppm of Og daily for various
hours for 17 months.   Ten  percent  (3 dogs) of  the animals developed pulmonary
arterial hypertension,  and approximately 30 percent (9  dogs) had excessive
systolic pressure, but there was no proportional relationship between pulmonary
arterial hypertension  and  0,  exposure.   Unless sample sizes were too  small to
find  adequate dose-response effects,  the authors attributed  the  results to
genetic susceptibility.
     Revis et al.  (1981)  studied the effects  of 0« and  cadmium,  singly and
                                                          3
combined, in  rats.  The  rats were exposed  to 1176  ug/m  (0.60  ppm) 0.,,
                                             3
5 hr/day for 3  consecutive days or to 3 mg/m   cadmium for 1 hr or to both
pollutants.   All exposure treatments resulted  in increases in systolic pressure
and heart rate.   Neither diastolic pressure or mean pressure was affected.  No
additive or antagonistic effects were  seen  with the pollutant  combinations.
     Costa et al.  (1983) measured heart  rate and standard intervals of cardiac
electrical activity  from  the  electrocardiographic   (EKG) tracings  of rats
exposed  to 392,  1568,  or 3920 ug/m  (0.2, 0.8,  or  2 ppm) of Og 6 hr/day, 5
days/week for 62  exposure  days  as  part of a  more extensive evaluation of lung
function (Section  9.3.2).   Heart rate  was not altered by 0» exposure.   The
                                                                          3
predominant effects  occurred at the  highest  0, concentration  (3920  ug/m ,
2 ppm)  at  which there  was evidence of  partial A-V blockade and distorted
ventricular activity,  often  associated  with  repolarization  abnormalities.
     Friedman et  al.  (1983) evaluated the effects  of a  4-hr  exposure to 588
             2
and 1960 ug/m  (0.3 and 1.0 ppm) of 03 on the  pulmonary gas-exchange region of
dogs  ventilated through an endotracheal  tube.   Pulmonary capillary  blood flow
and arterial 0?  pressure (PaO-) were decreased 30  min following  exposure to
both 0., concentrations, and arterial pH  (p.H0)  was decreased following exposure
      J                                    d

                                   9-158

-------
            3
to 1960 |jg/m   (1.0  ppm)  of 0,.  Decreases  in pulmonary  capillary blood  flow
                                                       3
persisted 24 hr following  exposure to  588  and 1960  jjg/m   (0,3  and 1.0 ppm)  of
                                                         3
DO and as  long as 48 hr following exposure to  1960 pg/m  (1.0 ppm) of  0~.
Persistent decreases in pH  and Pa09 were observed 24 hr following exposure to
         3                a        L
1960 ug/m  (1-0 ppm) of 0~.  Pulmonary edema, determined histologically and by
increased lung water  content and tissue volume, was observed 24 hr following
                      3
exposure to 1960  ug/m  (1.0 ppm)  of 0_.  The data  indicate  that 03 exposure
can cause both acute and delayed changes in cardiopulmonary function.

9.4.3  Hematological andSerum Chemistry Effects
     Hematological effects reported in laboratory animals and man after  inhala-
                                                   3
tion of near-ambient 0., concentrations (£ 1960 ug/m ; £  1.0 ppm) indicate that
03 or some reaction product of 0~ can cross the blood-gas barrier.   In addition
to reports of  morphological  and biochemical effects  of  0,  on erythrocytes,
chemical changes  have also been detected in serum after iji vitro and in vivo
03  exposure.   Hematologieal parameters  are frequently  used to evaluate 0^
toxicity, because red blood cells (RBCs)  are structurally and metabolically
simple  and well   understood, and because the relatively noninvasive methods
involved in obtaining blood samples from  animals  and man make blood samples
available for study.
9.4.3.1   Animal  Studies - In Vivo Exposures.   The  effects  of jji  vivo  Q~
exposure in animals,  including studies reviewed in the  previous 03 criteria
document (U.S. Environmental Protection Agency, 1978), are summarized in Table
9-18.                                               •'-'•••'
     Effects of  03  on the blood were  first reported by Christiansen and  Giese
(1954)  after  they detected an  increased resistance to hemolysis of RBCs from
                           o
mice  exposed  to  1960 ug/m  (1.0 ppm)  for  30 min.   Goldstein  et al.  (1968)
reported a significant decrease in RBC  acetylcholinesterase (AChE) activity
                                      3
after  exposure of mice to 15,680 ug/m  (8  ppm)  of  0, for 4  hr.  Menzel  et al.
(1975a)  observed  the presence of Heinz bodies  in approximately 50 percent  of
                                                3
RBCs  in  the  blood of mice  exposed  to 1666 ug/m  (0.85  ppm) of 03 for 4 hr.
About  25 percent of  RBCs  contained Heinz  bodies after continuous exposure  of
mice to  0.85 ppm  of 0,  for  3 days.  Heinz  bodies are  polymers  of methemoglobin
formed  by  oxidant stress; they appear to  attach to the  inner  membrane of the
RBC.   However,  Chow et al. (1975) detected no  significant  changes in GSHs,
G-6-PDH,  oxidized glutathione reductase, or GSH peroxidase  in RBCs  of  rats or
monkeys  exposed to  the  same  0~ concentration 8  hr/day for 7  days.
                                   9-159

-------
                                                        TAfllE 9-18.   HEHATOLOGY:  ANIMAL ~ IK VIVO EXPOSURE
cn
o
Ozone
concentration
ug/w3
110
118
235
470
941
392
392
392
392-
1960
490
980
1372
588
588
ppm
0.056
0.6
0.12
0.24
0.48
0.2
0.2
0.2
0.2-
1.0
0.25
0.50
0.70
0.3
0.3
Keasurensent3
method
c
UV
ND
UV
UV
UV
UV
UV
UV
Exposure
duration and
protocol
93 days
2.75 hr
4 hr
8 hours/day,
5 days/week,
3 weeks
60 «1n
1-4 hr
2.75 hr
3 hr/day
until death
(2-3 wk)
3 hr
Observed effect(s)
Decreased whole blood cholintsterase,
which returned to normal 12 days after
exposure ceased.
RBC survival decreased at 0.06,
0.12, and 0.48 ppm; no concentra-
tion-response relationship.
Increased osiiotic fragility and
spherocytosis of RBC's.
Increased serum glutamic pyruvic
transaai nase and hepatic ascorbic
acid, No change in blood catalase.
Small decrease in total blood sero-
tinin.
Plasma creatine phosphokinase
activity altered immediately and
15 Bin postexposure; no effect
30 min postexposure. No change
in plasma histamine or plasma
lactic acid dehydrogenase.
RBC survival decreased at 0.25 ppm
only.
Increased mortality in mice
parasitized with Plasmodium
berghei . Increased number of
parasitized red blood cells.
No effect on RBC reduced glutathione,
L-ascorbic acid, hemoglobin, red
blood cell counts. Slight increase
(p = 0.08) in % methemoglobin.
Decreased hematocrit in 03-low
vitamin C group. Generally, vitamin
C deficiency did not increase sensi-
tivity to 03.
Species Reference
Rat Eglite, 1968
Rabbit Calabrese et al . ,
1983a
Rabbit Brinkman et al.,
1964
Mouse Veninga, 1970
Rabbit Veninga, 1967
Mouse Veninga et al.,
1981
Sheep Moore et al . ,
1981a
Mouse Moore et al., 1984
Guinea pig Ballew et al., 1983

-------
TABLE 9-18.  HEMATOLOGY:  ANIMAL — IN VIVO EXPOSURE  (continued)
Ozone Exposure
concentration Measurement duration and
ug/M3 ppi method protocol
627 0.32 UV 6 hr
± dietary
vitamin E
784 0.4 NO 6 hr/day,
5 days/week,
6 months
784 0.4 NO 6 hr/day,
5 days/week,
10 months
784 0.4 NO 10 months
 784 0.4 NO 6 hr/day,
fii • 5 days/week,
en 10 months
M
980 0,5 UV 2.75 hr
980 0.5 MAST Continuous,
23 days
980 0.5 NBKI 8 hr/day,
7 days
1254 0.64 UV 8 hr/day »
lyr
1470 0.75 ND 4 hr/day,
4 days
Observed effect(s)
Increased erythrocyte G-6-PD and
decreased AChE (both diets).
Increased plasma vitamin E
(both diets).
No change serum trypsin inhibitor
capacity.
Increase in serum protein esterase.
Increase in serum protein esterase.
Decreased serum albumin concentra-
tion. Increased concentration of
a- and 6-globulins. Not much change
in p-globulin. No change in total
serum proteins.
Decreased erythrocyte GSH.
Increased hemolysis of erythrocytes
of animals depleted of vitamin E.
No such change when rats received
vitamin E supplements.
No change in GSH level or activ-
ities of GSH peroxidase. GSH
reductase, or G-6-PD in erythro-
cytes.
Altered RBC morphology: decreased
number of discocytes, increased
number of knizocytes, stomatocytes,
and spherocytes. No effect on RBC
FA composition.
RBC's: Increased fragility;
decreased GSH, AChE; no effect
on LDH, G-6-PD.
Species
Mouse
Rabbit
Rabbit
Rabbit
Rabbit
Sheep
Rat
Monkey ,
rat
Monkey
Monkey
Reference
Moore et al . , 1980
P'an and Jegier,
1971
Jegier, 1973
P'an and Jegier,
1972
P'an and Jegier,
1976
Moore et al . ,
1981b
Menzel et al., 1972
Chow et al . , 1975
Larkin et al., 1983
Clark et al., 1978

-------
                                               TABLE 9-18.   HUWTQLQGY:  AHIHAl — IN VIVO EXPOSURE  (continued)
H
G\
to
Ozone
concentration
pg/n3 ppa
1568 0.8
1568 0.8
1568 0.8
1666 0.8S
1686 0.86
1960 1.0 .
1960 1.0
1960- 1.0
3920 2
Exposure
Heasureaent duration and
nethod protocol
mm 7 days
NBKI 8 hr/day,
7 days
NBKI Continuous,
29 days
HAST 4 hr
NO 8 hr/day,
§ days/week,
6 months
UV 4 hr ± vitamin E
ND 30 Bin
CHEM 2 or 7 days
Observed effect(s)
Increased activity of OSH pero-
oxidase, pyruvate kinase, and
lactate dehydrogenase; and
decrease in red cell level of
GSH of vitamin E-deficient animals.
Animals in both vitamin E-deficient
and supplemented diet groups exhibited
no change in activities of 6-6-DP,
catalase, and superoxide disnutase
and in levels of thiobarbituric acid
reactants, met hemoglobin, hemoglobin,
and reticulocytes.
No change in total lactate dehydro-
genase activity or isoeniyae pattern
in plasma or erythrocytes.
Increased lysozyne activity by
day 3.
Increased Heinz bodies in RBC's
(decreased with continual exposure).
Increased infestation and mor-
tality after infection with
Plasnodiun berghei. Increased
acid resistance of erythrocytes.
Decreased filterability. No pro-
tection by vitamin E. No lipid
peroxidation.
Increased resistance to erythrocyte
hemolysis.
No changes.
Species Reference
Rat Chow and Kaneko,
1979
Monkey Chow et al . , 1977
Rat Chow et al., 1974
House Menzel et al . , 1975a
House Schlipkoter and
Bruch, 1973
Mouse Dorsey et al., 1983
Mouse Mizoguchi et al.,
1973;
Christiansen and
Giese, 1954
Rat, Cavender et al.,
guinea pig 1977

-------
                                                TABLE 9-18.   HEMATOLOGY:  ANIMAL —  IN VIVO EXPOSURE   (continued)
Ozone
concentration
ug/nd ppro
1960 Ja.O
5880 3.0
1960 1
Exposure
Measurement duration and
method protocol
CHEH 4 hr
UV continuously,
2 wk
Observed effect(s)
No effects on oxyhemoglobin affinity,
2,3-DPG concentrations, heme-02
binding.
Increased serum cholesterol , low
density lipoproteins and very
low density lipoproteins. Males
apparently more affected than
females. No effect on trigly
cerides.
Species Reference
Rabbit Ross et al., 1979
Guinea pig Vaughan et al. ,
1984
M
1960
3430
5880
1960
2940
11,760
15,680
1
1.75
3
1
1.5
6.0
8.0
CHEM, UV 5 hr/day,
10 days within
14- day period
CHEM, UV 5 hr/day,
15 days
within 19-day
period
UV 3 days
4 days
4 days
No effect on serum lipids and Rat
lipoproteins at 1 ppm. Concentration
related linear increase in total
lipoprotein-free cholesterol
and high-density lipoprotein
total cholesterol; decrease in
triglycerides.
Increased serum total cholesterol
(p = 0.1) high density lipoprotein-
cholesterol (p = 0.08) and high
density lipoprotein-free cholesterol
(p = 0.006); decrease in trigly-
cerides (p = 0.06).
No effect on SOD, GPx, K* influx Rat
ratios (all levels). Increased Hb,
Hct, echinocytes II & III (6 & 8 ppm);
echinocytes correlated with petechiae
in lungs, indicative of vascular
endothelial damage.
Mole et al., 1985
Larkin et al., 1978
           Measurement method:   ND = not described;  CHEH = gas phase cherailuminescence; UV - UV photometry;  NBKI  = neutral  buffered potassium iodide;
            MAST = KI - coulometric (Mast neter);  I = iodometric.

            Abbreviations used:   RBC = red blood cell; G-6-PD = glucose-6-phosphate dehydrogenase; AChE =  acetylcholinesterase;  GSH = reduced
            glutathione;  GSH peroxidase = glutathione peroxidase; GSH reductase = glutathione reductase; FA = fatty acid;  LDH =  lactic dehydrogenase;
             '?!  E-r 2«3*'»PhosP|wfllycei'ate;  SOD = superoxide dismutase; GPx = glutathione peroxidase; K = potassium,  Hb  =  hemoglobin; Hct = henato- •
            cm; P6F2a = prostaglandin F2a; PGE2 = prostaglandin E2.

            Spectrophotometric method using dihydroacridine.

-------
     In more recent studies, Clark et al.  (1978)  investigated the biochemical
                                                          3
changes in RBCs  of squirrel  monkeys exposed to  1410  yg/m  (0.75 ppm) of Q3
4 hr/day for 4 days.   They observed an increase in RBC fragility with decreases
in GSH and AChE  activities.   No changes were  detected in G-6-PDH or lactic
dehydrogenase (LDH) activities.  After a 4-day recovery period,  RBC  fragility
was still significantly increased,  although to a lesser degree.   AChE activity
returned to control  levels  at 4 days postexposure; however, RBC GSH remained
significantly lowered.
     Ross et al.   (1979)  investigated the effects  of 0-, on the oxygen-delivery
                                                                         3
capacity of erythrocytes.  After exposure  of rabbits  to 1960 or  7880 pg/m   (1
or 3 ppm) of 03  for 4 hr, no  changes were detected in RBC 2,3-diphosphogly-
cerate concentration, oxyhemoglobin dissociation curve, or heme-oxygen binding
of RBCs.   Analysis of blood parameters 24  hr after exposure revealed no delayed
effects of 03.
     Alterations in RBC morphology have been previously observed in CK-exposed
laboratory animals and man  (Brinkman  et  al., 1964;  Larkin et  al., 1978).
                                                                             o
Similar observations  have recently been made in  monkeys exposed  to  1254 pg/m
(0.64 ppm) of  Oo  for  8  hr/day over a  1-yr period (Larkin et al.,  1983).
Ultrastructural  SEM  studies  of RBC1s following  exposure  to 0,  demonstrated
reduced  numbers  of normal discocytes  and  increased  numbers of  knizoeytes,
stomatocytes,  and spherocytes, which were either absent  or found  in  small
numbers in the blood of air-exposed controls.  Despite changes in shape, there
were no  differences  in the fatty acid  composition of the erythrocyte total
lipids.  Values  for  hematocrit,  hemoglobin, mean corpuscular volume, and red
cell and  reticulocyte count  were the same  in control  and 0,,-exposed animals.
                                                                               3
Moore et al.  (1981a) reported reduced RBC  survival in  sheep exposed  to 490 ug/m
(0.25 ppra) of 0,, for 2.75 hr.  Similar reductions in  RBC  survival were reported
                                                                            3
following 2.75-hr  exposures  to 0-  concentrations as  low  as 118 and  235 pg/m
(0.06 and 0.12 ppm) in rabbits (Calabrese  et al., 1983a).
     Vitamin E deficiency has been associated with an increased hemolysis in
rats and  other animal species (Scott, 1970; Gross and Melhorn,  1972).  Chow
and Kaneko (1979)  reported significant  increases in  RBC GSH peroxidase, pyru-
vate kinase, and LDH activities, and a decrease  in RBC GSH after exposure of
                                     3
vitamin E-deficient rats  to 1568 ug/m  (0.8 ppm)  of 03 continuously  for 7 days.
These  effects  were not observed in  vitamin E-supplemented rats (45 ppm  of
vitamin  E for 4 months).  The activities of  G-6-PD, catalase, superoxide
                                   9-164

-------
dismutase, and levels  of  TBA reactants, methemoglobin and reticulocytes were
not altered by 0- exposure or by vitamin E status.
     Moore et al.  (1980)  investigated the effects  of dietary vitamin E on
                                                      3
blood of  9-month-old  C57L/J  mice exposed to 627  (jg/m  (0.32  ppm)  of 03 for
6 hr.  Animals were maintained on vitamin E-deficient, or supplemented (3.9 mg
tocopherol/100 Ib., twice  the minimal daily requirement) diets  for  6 weeks
before 0™  exposure.   Mice  on  the vitamin  E-deficient  diet showed a 24-percent
increase  in G-6-PD activity  over controls after  03 exposure, and mice  fed a
supplemented diet exhibited a 19-percent increase.  Decreases in AChE activity
were observed in  both vitamin E-deficient (19-percent  decrease) and vitamin
E-supplemented (12-percent decrease) groups.
     Dorsey et al.  (1983)  evaluated the effects  of 03  on  RBC deformability
after exposure of vitamin E-deficient and supplemented (105 mg of tocopherol
                                           3                     3
per  kg of chow)  male  CD-I mice to 588 |jg/m  (0.3 ppm), 1372 ug/m  (0.7 ppm),
or 1960 ug/m  (1.0 ppm)  of 03 for 4  hr.  After incubation  of RBCs in buffer
(0.9 percent RBCs)  for up to 6 hr at 25°C, the time required for 2.0 ml of RBC
suspension to pass through a 3-[jm pore size filter was determined.  Exposure
                     3                       3
of mice  to 1960  ug/m  (1.0 ppm) or 1372 ug/m   (0.7 ppm) of 03 and incubation
of RBCs for 6 hr  resulted  in a significant increase in  filtration time of RBCs
from 0~-exposed  mice, and a  lack of  protection by dietary vitamin  E.   The
                                                            3
hematocrit of vitamin E-deficient  mice exposed to 1960 (jg/m  (1.0 ppm) of 0~
was  significantly greater than that of nonexposed vitamin E-supplemented mice.
The  increased hematocrit  was attributed to a  loss  of RBC deformability, and
sphering  resulting in decreased packing  of cells during centrifugation for
hematocrit  determination.   No TBA  reactants were detected  in the blood of
exposed animals,  with  or without vitamin E.
     The  influence of vitamin  C deficiency  on erythrocytes of  guinea  pigs
                    3                                '   '     '
exposed to 588 ug/m  (0.3 ppm) for 3  hr and examined  0.5 or 3 hr post-exposure
was  studied by Ballew et al.  (1983).  Ozone caused no effect on  reduced gluta-
thione  levels in erythrocytes.  There  was  a slight  increase (p = 0.08)  in
percent methemoglobin in 0.~-exposed animals.  Vitamin C levels did not  signifi-
cantly influence  these results.  Plasma L-ascorbic acid levels were  not affec-
ted  by 0, exposure.   Hematocrits were decreased in animals  on the low vitamin C
diet that were  exposed  to 03-   Hemoglobin and red  blood cell  counts were
unaffected by 03.
     Moore et al. (1984)  infected  mice with  Plasmodium berghei  (a blood-borne
                                                        3
malarial  parasite) 1 day prior to  exposure to  588 ug/m  (0.3 ppm)  Q3.  The
                                    9-165

-------
exposure lasted for  3 hr/day until death, or  approximately 2 to 3 wk.  Mice
exposed to 0-,  did not live as long as the controls.  Ozone-exposed mice also
had an  increase  in the number of  parasitized  red blood cells.   The authors
hypothesize that  there  are 2 potential mechanisms  responsible:   03 may have
altered the erythrocyte  membrane,  making it more permeable to IP.  berghei, or
Og increased the reticulocyte count, reticulocytes possibly being more sensitive
^•° £•  berghei  infestation.  These results are consistent with those of Schlipkoter
and Bruch  (1973),  who reported,  without statistical analysis, an increase in
infestation with  £.  berghei and higher  mortality in mice exposed for 6  mo
(8 hr/day, 5 days/wk) to 1686 Mfl/m  (0.86 ppm) O.
9.4.3.2  In Vitro Studies.   The effects of  in  vitro 0, exposure of animal
           ...........                               — »         ^
blood( have been  studied by a number of  investigators,  and these  reports are
summarized in Table 9-19.
     The effects of  i_n  vitro  03 exposure on  human  RBCs  have been  evaluated by
using a number of different end points, such as increases in complement-mediated
cell damage  (Goldstein  et  al.,  1974a), formation  of  Heinz  bodies (Menzel
et al., 1975b),  decreases  in RBC native protein fluorescence  (Goldstein and
McDonagh,  1975), and decreases  in concanavalin A  agglutinability (Hamburger
et al . , 1979).   Exposure of RBCs  or their membranes to  0- has  also  been shown
to inhibit (Na+ - K-*-) ATPase (Kindya and Chan, 1976; Chan et al . , 1977; Koontz
and Heath, 1979;  Freeman et al.,  1979;  Freeman and Mudd, 1981).   Kindya and
Chan (1976) proposed that  inhibition of  ATPase by  0,, caused spherocytosis and
increased fragility  of  RBCs after 03 exposure.  (See Tab-le 9-20  for a  summary
of the human in vitro studies.)
     Kesner et al.  (1979) demonstrated that Og-treated phospholipids inhibited
RBC membrane  ATPase.   Addition  of semicarbazide to  Og-exposed phospholipids
before mixing with RBC  membranes  substantially reduced  the inhibitory  effect,
suggesting that  the  inhibitors may be  carbonyl  compounds.   In addition, a
slower- forming semicarbazide-insensitive inhibitor was formed.
     Verweij and Steveninck (1980, 1981) reported  that  semicarbazide and also
p-aminobenzoic acid  (PABA) might protect by acting as 03 scavengers.  Spectrin
(a major glycoprotein component of the RBC membrane) solution was treated by
bubbling 0»-containing  02  through the  solution at  4 ml/min (2.5  uM/min of 03)
for 1 or  9  min.   Semicarbazide (40 pM)  or PABA (40 pM)  inhibited the  cross-
linking of 03-exposed spectrin.   The inhibition of AChE and hexokinase activi-
ties of RBC  ghosts  exposed to 0~ was  also  partially prevented by these two
                +
agents,  as was K  influx into whole RBCs.  The authors attributed the inhibition
                                   9-166

-------
                                             TABLE  9-19.  HEMATOLOGY:  ANIMAL — IN VITRO EXPOSURE
Exposure
Ozone Measurement duration and
concentration method protocol
W3
M
cr\
-o
980-
3920
1960-
13,132
2156
4508
0.5 CHEM 2 hr
2.0
1.0 NBKI 90 min-
6.7 4 hr
1.1 UV 16 hr
2.3
Observed effect(s) Species
Decrease in agglutination of erythro- Rat
cytes by concanavalin A.
Decreased erythrocyte catalase levels Rat,
at > 5 ppm when animals were pretreated mouse
with atninotriazole.
No effect on hemoglobin. No change Mouse
in organic free radicals as measured
by EPR spectra*. No statistics.
Reference
Hamburger and
Goldstein, 1979
Goldstein, 1973
Case et al., 1979
Measurement method:   CHEM = gas phase chemiluminescence; UV = UV photometry; NBKI = neutral buffered potassium iodide.

-------
                                                          TABLE 9-20.  HEHATOLOGY:  HUMAN - IN VITRO EXPOSURE
H
a\
co
Ozona3
concentration
980 pg/ra3 (0.5 ppm)
1960 jjg/ra3 (1.0 ppm)
Oa-treated phospholipids
4 uM/ndn
Methyl ozonide
10-4-2xlQ-3 M
750 nH/nin
106 nM/min
300 nM/min
0-9.8 uM/g of Hb
0.84 |iM/nin
78400 ug/m3 (40 ppm)
1,960 ug/m3 (1.0 ppn)
Measurement
method
CHEM
NO
I
NO
NBKI
NBKI
NBKI
NBKI
NBKI
NBKI
Exposure
duration and
protocol
0.5-2 hr
5, 10, 15, and
20 Bin
1 min
30 rain
14.3 or 43,0
nMol of 03 per 10s
cell equivalent
5, 10, 20, 30,
40 and 50 Bin
NO
0-2 hr
2 hr
20 and 60 min
Observed effect(s)
Decreased agglutination of RBCs by
concanavalin A.
Decreased ATPast activity.
Decreased ATPase activity.
Heinz body formation. Prevented
by dietary vitamin E.
RBC — No effect on ATPase.
Decreased cation transport.
RBC ghosts ~ decreased ATPase
activity.
Decreased activity of purified
Oj-proteinase inhibitor.
Decreased glyceraldehyde-3-PD.
Decreased ATPase.
No statistics.
Decreased 6SH. No effect on
Hb or on glucose uptake.
Increased complement-mediated cell
damage.
Decreased native protein fluore-
Species
Hunan
Human
(RBC ghosts)
Hunan
(RBC ghosts)
Human
Human
Human
Human
(RBC ghosts)
Hunan
(RiCs,
RBC ghosts)
Human
Human
Reference
Hamburger et al
Kesner et a1_, ,
Kindya and Chan
Henzel et al . ,

. , 1979
1979
, 1976
1975b
Koontz and Heath,
1979
Johnson, 1980
Freeman et al , ,
1979


Freeman and Hudd,
1981
Goldstein et al
1974a
Goldstein et al
* »
. ,
                                                                            scence.  No statistics.
(RBC ghosts)    1975

-------
                                                TABLE 9-20.  HEMATOLOGY:  HUMAN - IN VITRO EXPOSURE   (continued)
(£>
Ozone8 Measurement
concentration method
40 nM/min I
2.5 (jM/nrin I
2.5 jjM/min I
?Not ranked by concentration; listed by
Measurement method: NO = not described
Exposure
duration and
protocol
4 min
20, 40, and 60
min
20, 40, and
60 min
reported values.
; CHEM = gas phase
Observed effect(s)
Decreased ATPase activity; lost 40%
membrane sulfhydryls. Lipid per-
oxidation and protein crossl inking
detected.
Pretreatment with semicarbazide
prevented crosslinking.
Cross-linking of membrane proteins
inactivation of glyceraldehyde-3-
phosphate dehydrogenase.
Crosslinking of spectrin. Decreased
ACHase activity. Increased K+ leak-
age from RBCs. Semicarbazide and
p-amino benzole acid prevented
these 03 effects.
cheroi luminescence; NBKI = neutral buffered
Species
Human
(RBC ghosts)
Human
Human
(RBC ghosts)
potassium iodide;
Reference
Chan et al.
Verweij and
Steveninck,
Verweij and
Steveninck,

, 1977
1980
Van
1981
I = iodometric.

-------
of ATPase  to oxidation  of phospholipids  with subsequent cross-linking of
membrane protein  by  lipid  peroxidation products.   Because the  reaction  of
ozonolysis products  with semicarbazide and PABA during  0,  treatment of RBCs
was not  directly  measured in these studies, the protective mechanism remains
unclear.
     In  a  recent  study, Freeman and  Mudd (1981) investigated the  iji vitro
reaction of  0,,  with sulfhydryl groups  of human RBC membrane, proteins, and
cytoplasmic  contents.   After exposure of  RBCs  to 0, in 02 at 20 ml/min (0.84
p:Mol/min of  03) for up to 2 hr, oxidation of  intracellular GSH was  observed.
Ozone exposure  produced membrane disulfide cross-links in RBC ghosts but not
in intact RBCs.  Neither oxyhemoglobin  content  nor glucose uptake was affected
by 03 exposure  of RBCs.  These data  support earlier studies of Menzel et al.
(1972) that  reported decreased RBC GSH  levels following exposure  of rats to
        3
980 ug/m  (0.5 ppm)  of 03 continuously  for 23  days.
     Although jm  vitro  studies  using  animal  and  human  RBCs have provided
information  on the possible mechanism by  which  0, may react with cell membranes
and RBCs,  extrapolation of these data to in vivo 0- toxicity in  man is  diffi-
                                          """""       O
cult.  In most jm vitro  studies, RBCs were exposed by bubbling high  03 concen-
trations (>  1 ppm)  through cell  suspensions.   Not only were the  03 concentra-
tions unrealistic and the method of  exposure  nonphysiological, but  the toxic
species  causing RBC injury may be  different  during in vitro and  in vivo 0,,
                                                     ™"rm~   v ::::::'"::n     •—— -—.—,—„.  ^
exposures.   Because  of its reactivity,  it is  uncertain that 03 per se reaches
the RBCs after  inhalation but may instead appear in blood in the form of less
reactive products (e.g.,  lipid, peroxides).  However, during in vitro exposure
of RBC suspensions,  03 or highly reactive free-radical products  (e.g.,  hydroxyl
radical, superoxide  anion, singlet  oxygen) may be the cause of injury.
9,4.3.3  Changes  In  Serum.   In addition to O3's effects  on  RBCs,  changes have
been detected in  the serum of animals  exposed to 0,.  P'an and Jegier  (1971)
                                      3
investigated the  effects of 784 ug/m  (0.4 ppm)  of 03 6 hr/day,  5 days/week
for 6 months on the  serum trypsin  inhibitor capacity (TIC) of rabbits.   With
the exception of  a  sharp rise  after  the first day of exposure,  TIC values-
remained within normal  limits.  However,  after exposure for 10 months,  the TIC
had progressively  increased to about three times  the normal  level (P'an and
Jegier, 1972).  Microscopic evaluation  suggested that the rise in  TIC may have
been due to  the thickening of small pulmonary  arteries.  The results from this
study'are questionable,  however, because  the rabbits may  have had  intercurrent
infectious disease,  which was more  severe in the exposed  animals (Section 9.3.1).
                                    9-170

-------
     P'an and  Jegier (1976) also reported  changes  in serum proteins after
                                3                         3
exposure of rabbits  to  784 [jg/m  (0.4 ppm) and  1960  [jg/m  (1.0 ppm) of 0~.
                               3
Following exposure to 784  [jg/m  (0.4 ppm) of 03  for  105 days, the  albumin
concentrations began to  decrease,  and a- and 6-globulin concentrations began
to increase.  At the end of 210  days  of  exposure, the mean  albumin  level  fell
16 percent,  the crglobulin  level  rose 78 percent,  and  the  6-globulin levels
fell  46 percent.   No significant changes were observed in total protein concen-
tration.
     Chow et al.  (1974) observed that the  serum lysozyme activity  of  rats
increased significantly during continuous (24 hr/day) but not during intermit-
                                      2
tent (8  hr/day)  exposure  to 1568 |jg/m   (0.8 ppm) of  0~  for 7  days.  The  in-
creased release of lysozyme into the plasma was suggested to be a result of 0-
damage to alveolar macrophages.
     Veninga et al.  (1981)  reported that short-term exposures of mice to low
03 concentrations  induced  changes  in serum  creatine phosphokinase (CPK)
activity.   Ozone  doses  were expressed as  the  product of concentration and
                                                 3
time; the maximum  03 concentration was 1600 pg/m   (0.8 ppm), and the maximum
exposure time  was  4 hr.  Alterations  in CPK  were  detected immediately and
15 min  after  termination  of the exposure.  By  30 min postexposure, the CPK
activities had returned to control levels.  Neither plasma histamine nor plasma
LDH was altered by the range of 03 doses employed.  The authors concluded that
these  responses  may  represent adaptation of the  animals to 03 toxicity  by
enhanced metabolic processes.
     Serum  lipids  and  lipoproteins  of rats exposed continuously  for 2  wk to
         3
1960 |jg/m   (1 ppm) were  determined  (Vaughan et al., 1984).   Serum  from each
guinea  pig  was  sampled before and'  immediately  after  the 2-wk exposure and
30 days .after exposure  ceased.   Thus, each animal  served as  its  own control
and there was no air-exposure group.  Immediately after exposure, cholesterol,
low density lipoproteins-,  and ve-ry low density lipoproteins were elevated in
males.   Generally, females  had similar effects,  but  no changes  ,in  very low
density  lipoproteins.  Triglycerides were not affected  in either sex immediately
after exposure.  Although  statistical comparisons  of sex susceptibility were
not performed, it appears that males were more affected than females.  Statis-
tical  tests of the  post-exposure group were unclear,  but  it appears  that
levels of cholesterol, low density lipoproteins, and very low density lipopro-
teins had returned to pre-exposure values.
                                   9-171

-------
     Serum lipids and lipoproteins have also been evaluated in male rats after
repeated 0-  exposure (Mole  et  al.,  1985).   A concentration response  study
                                                    o
involved exposure to  air;  1960, 3430, and 5880 ug/m  (1, 1.75, and 3 ppm) 03
for 5 hr/day  for  10 exposure days within a  14-day period.  For a  given  rat,
serum samples taken 2 days prior to the first exposure were compared to samples
taken 20 hr after the last exposure;, each animal served  as its own control.
Another group of animals was exposed to air and sampled pre- and postexposure.
Shifts in the 0--exposed rats (pre-  vs. postexposure) were compared statisti-
cally to shifts in the air-exposed animals.  Ozone caused a concentration-related
linear increase  in  total  lipoprotein-free cholesterol and high density lipo-
protein total  cholesterol  (both the  free and esterified components)  and  a
decrease in  total  lipoprotein  triglycerides.  There  was  no effect on  high
density lipoprotein-tryglycerides  or on total lipoprotein-free fatty  acids.
                 2
At the 1960 ug/m   (1 ppm)  03 level, none of the values was elevated signifi-
cantly over controls.  In a sampling time study (Mole et al.,  1985), rats were
                     2
exposed to 1960  ug/m   (1  ppm)  03 for 5 hr/day for 15 days  in  5-day segments
within a 19-day  period.   Serum samples were taken from each rat 4 days prior
to the first  exposure and at 7 times  (0  to  44 hr)  after the last exposure.
Ozone increased  serum total  chlolesterol  (p  = 0.1), high  density lipoprotein-
cholesterol (p = 0.08), and  high density lipoprotein-free cholesterol  (p  =
0.006), and decreased triglycerides (p = 0.06).   The changes  appeared to  be
maintained over  the 44-hr  post-exposure period and were  greater  than those
                          2
observed at  the 1960 ug/m   (1  ppm)  03 level of the  concentration-response
study described  above that used fewer  days  of exposure.   Thus,  03 caused a
mild hypercholesterolemia and hypotriglyceridemia.
     Both the Vaughan et al. (1984)  and  Mole et al.  (1985) studies  report
                                                          /
increases in  serum  cholesterol.  There is some disparity between results for
other lipoproteins.    The Vaughan et al. study did not account for changes that
can be produced  by  exposure  stress  alone, as indicated  by Mole._et_al.  (19.85).
The Mole  et  al.  (1985) studies  showed  only  marginally  significant effects.
Thus, a possible conclusion from the rat and guinea pig studies is that short-
term exposure to 1960 ug/m  (1  ppm) has the potential of elevating cholesterol
in animals.   Elevation  in  human serum  cholesterol is a risk factor  in human
coronary heart disease (Gottov  1979; Dawber, 1980).
9.4.3.4   Interspecies Variations.   The use  of animal  models  to investigate
the effects of  03 on the blood is  complicated,  because few species respond

                                   9-172

-------
like humans.  The  rodent  model  has been most  commonly used to predict the
effects of 0- on human RBCs (Calabrese et al., 1979). However, the reliability
of this model was  challenged by Calabrese  and Moore  (1980).on the following
grounds:    (1) ascorbic  acid synthesis  was  significantly increased in mice
following 03 exposure (Veninga and Lemstra,  1975), (2) ascorbic acid protected
human G-6-PD-deficient  RBCs jji vitro  from the  oxidant stress  of acetylphenyl-
hydrazine  (Winterbourn, 1979), and (3) humans lack the ability to synthesize
ascorbic acid.  Although  Calabrese and Moore (1980) stressed that this hypo-
thesis is  based on a  very limited  data base, they,point  out the importance of
developing animal  models  that  can accurately predict  the  response  of human
G-6-PD-deficient humans to oxidant stressor agents.   In  another report, Moore
et al. (1980)  suggested that C57L/J  mice may present an acceptable  animal
                                                            3
model, because these  mice responded to 0, exposure  (627  ug/m  , 0.32 ppm for  6
                                         «3
hr)  in a  manner  similar to that of humans,  with  increases  in serum vitamin E
and  G-6-PD activity.   Unlike many other mouse strains, the C57L/J strain has
low  G-6-PD activity,  which is  similar  to that found in human RBCs.   Moore
et al. (1981b) also followed up on the proposed  use of Dorset sheep  as  an
animal model  for  RBC G-6-PD deficiency  in  humans (National Research Council,
1977).  However,  Dorset sheep  were found to be no more sensitive than normal
humans with respect to 03-induced  changes in GSH  and also differed from humans
in the formation  of methemoglobin.   Further studies  (Calabrese et al., 1982,
1983b,d; Williams  et al., 1983a,b,c)  demonstrated that the  responses  of sheep
and  normal human erythrocytes were very  similar when  separately incubated with
potentially toxic  Q-> intermediates, but G-6-.PD-deficient human erythrocytes
were considerably  more susceptible.   Consequently, the authors also questioned
the  value  of  the  sheep erythrocyte as a quantitatively accurate predictive
model.

9.4..4  .Reproductive and .Teratpge.m'c Effects          ,
     Pregnant animals and developing  fetuses may  be  at greater risk to effects
from photochemical oxidants,  because the volume  of  air inspired by females
generally  increases  from  15 to 50  percent during  pregnancy  (Altman and Dittmer,
1971).   Before  1978,  experiments  designed  to investigate  the  reproductive
effects  of photochemical  oxidants often used  complex mixtures of gases,  such
as  irradiated auto exhaust (see  Section 9.5),  or  they used  oxidant  concentra-
tions  greater than  those typically found  in ambient air.  Brinkman  et al.
                                    9-173

-------
(1964) exposed pregnant mice to lower concentrations of 03, but the results of
their experiments are difficult to  interpret, because the  time of 0, exposure
during gestation and  postparturition was not specified.   They reported that
                                3
mice exposed  to  196  or 392 ug/m  (0.1 or  0.2 ppm)  of 03 for 7 hr/day and 5
days/week  over  3 weeks had normal  litter  sizes,  compared with air-exposed
controls.  However, there  was  greater neonatal  mortality  in  the  litters  of
                                                         3
0--exposed mice,  even at  the  exposure  level  of 196 [jg/m  (0.1 ppm)  of  03
(Table 9-21).  Unfortunately,  without more details  on the  period of exposure,
it is impossible to  ascertain  whether the decreased infant survival  rate was
due to development  interference  j_n  utero,  to a direct effect on the pups, or
to a nutritional deficiency caused  by parental anorexia  or reduced lactation,
or a combination of these effects.  When using a similar experimental  protocol,
                                                  o
Veninga (1967) found that mice exposed to 392 ng/m  (0.2 ppm) of 03 for 7 hr/day,
5 days/week  during  embryo!ogical  development and  the 3 weeks  after  birth
(total exposure  time  not  reported)  had  an increased  incidence  of excessive
tooth growth, although no statistical evaluation was provided.
     In more  recent  experiments,  Kavlock et al.  (1979) exposed pregnant rats
to 0, for  precise  periods  during organogenesis.  No  significant teratogenic
effects were  found  in rats exposed 8 hr/day to concentrations of Oq varying
from 863 to  3861 ug/m3  (0.44  to  1.97 ppm) during  early  (days 6 to 9), mid
(days 9  to 12),  or late (days 17 to 20) gestation, or the entire period of
organogenesis (Days 6 to 15).   Continuous exposure of pregnant rats to 2920 ug/
m  (1.49 ppm) of 0, in midgestation resulted in increased resorption of embryos.
                                                                  q
A single dose of 150  mg/kg sodium salicylate followed by 1960 ug/m  (1.0  ppm)
of 0~ during midterm produced a significant synergistic increase in the resorp-
tion rate, a decrease in maternal weight change, and a decrease in average fetal
                                                         3
weight.   Exposure of  pregnant  rats  8 hr/day to 862 pg/m  (0.44 ppm) of  03
throughout the period of organogenesis also resulted in a significant decrease
in average maternal weight gain.-
     In a follow-up study,  Kavlock et al. (1980) investigated whether i_n utero
exposure to 0, can affect postnatal  growth or behavioral development.   In con-
trast to the results of Brinkman et al.   (1964),  neonatal mortality of rats was
                                      3
not increased by exposure  to 2940 [jg/m   (1.5 ppm) of  0,  for  periods of 4  days
                                                                      3
during gestation.  Pups  from  litters of females exposed to  1960  yg/m  (1-0
ppm) of 03 during mid- (days 9 to 12) or late (days 17 to 20) gestation exhibi-
ted significant  dose-related reductions  in weight  6  days  after birth.  Pups
                                   9-174

-------
                                           TABLE 9-21.   REPRODUCTIVE AND TERATOGENIC  EFFECTS OF OZONE
vo
1
l_l
~J
in
Ozone
concentration
ng/B3 ppm
196 0, 1
392 0.2
862 0.44
2920 1.49
3
Measurement Exposure
method duration and protocol
ND 7 hr/day, 5 days/week
for 3 weeks
ND 7 hr/day, 5 days/week
for 3 weeks
I" 8 hr/day over entire
period of organogenesis
(days 6 to 15)
Continuous during mid-
gestation
Observed effect(s) Species
Increased neonatal mortality £4.9 to 6.8% House
vs. 1.6 to 1.9% for controls) .
Unlimited growth of incisors (5.43. incidence Mouse
vs. 0.9% in controls) .
Decreased average maternal weight gain. Rat
Increased fetal resorption rate (50% vs. 9%
for controls).
Reference
Brinkman et al. , 1964
Veninga, 1967
Kavlock et al. , 1979
1960     1.0
1960     1.0
2940     1.5
Continuous during late
gestation
Continuous during mid-
(day 9 to 12) or late
(days 17 to 20) gestation

Continuous during late
gestation (days 17 to 20)
Slower development of righting, eye opening,
and horizontal movement; delayed grooming
and rearing behavior.

Average weight reduced 6 days after birth.
3 males (14.3%) were permanently runted.
Rat
Kavlock et al., 1980
 Measurement method:   ND = not described,  I = iodoroetric (Saltzntan  and Gilbert,  1959).
 No statistical evaluation.

-------
from the  late  gestation  exposure group were affected to a greater extent and
for a longer period of time after parturition.  In fact, several males exposed
            3
to 2940 |jg/m   (1,5 ppm)  of 03 during late gestation were  also significantly
slower  in the development  of  early movement reflexes and  in  the onset of
grooming and rearing behaviors.  The authors pointed out that it is impossible
to distinguish between prenatal  and postnatal contributions to  the behavioral
effects,  because  foster  parent procedures were not  used  to raise the pups.

9.4.5  Chromosomal and Mutational Effects
9.4.5.1   Chromosomal Effects of  Ozone.   A large portion of  the  data available
on the chromosomal and mutational effects of 0, was derived from investigations
                          3
conducted above 1,960 ug/m  (1 ppm) of 03, and their relevance to human health
is questionable.   However,  for completeness of the review of the literature,
and for possible insight into the mechanisms by which 0™ may produce genotoxi-
city, this discussion will  not be limited to data derived from research conducted
                       3
at or below  1,960 ug/m  (1 ppm)  of 0-,-   Data derived predominantly from In
                                                                         3
vitro experiments  conducted at CL concentrations in excess of 1,960 (jg/m  (1
ppm) of 0- will  be discussed .first (Table 9-22), followed by a discussion of
the genotoxicity  data  from  both  i_n  vitro and In vivo  research  conducted at  or
below 1 ppm of 03  (Table 9-23).
     The potential for genotoxic effects  relating to 0- exposure was predicted
from the  radiomimetic  properties of 03.   The decomposition of 03 in  water
produces  OH  and  HOp radicals, the same  species that are generally considered
to be the biologically active products  of ionizing radiation.  Fetner (1962)
reported  that  chromatid  deletions were  induced in a time-dependent manner in
                                      3
human KB  cells  exposed to  15,680 M9/m   (8 ppm)  of 03 for 5 to 25 min.  The
chromatid breaks  were  apparently identical to those  produced  by x-rays.   A
10-min exposure  to 8 ppm of 0., was slightly more efficient in the production
of chromatid breaks than 50  rad of x-rays'.   Significant mitotic delay was
measured  in  neuroblasts  from grasshoppers (Chortophaga  viridifaciata)  exposed
to 3500 to 4500 ug/L of Og  in a  closed system (Fetner, 1963).
     Scott and Lesher (1963) measured a  sharp loss of viability with Escherichia
coli as the  0™ concentration was increased.   Viability was reduced to zero
   111          *5
when cells were exposed to  1 pg/ml  of Q~.  Damage 'to cell membranes was evident
by the  leakage  of nucleic acids and other  cellular components from  cells
exposed to 0.18 ug/ml of 0-
                                   9-176

-------
                               TABLE 9-22.   CHROMOSOMAL  EFFECTS FROM IN VITRO EXPOSURE TO HIGH OZONE CONCENTRATIONS
                  a
     concentration
Measurement
    method
     Exposure
duration and protocol
          Observed effect(s)
  Species
   Reference
15,680 MO/M8 (8ppm)
   UKI
    5-25 min
Chroinatid deletions.
Humans KB
cells
Fetner, 1962
98,000 ug/m3 (50 ppro)
   MAST"
    30 min
Tex mutants deficient in
repair of x-ray-induced
DMA strand breaks were more
sensitive to lethal effects
of Oa than were the wild-type
repair-proficient parental strains
Escherichia
coTi
Hamelin and
Chung, 1974
98,000 M9/B3 (50 ppm)
   MASF
    30 Bin
DNA Polymerase I mutant strains
(KMBL 1787, 1789, 1791) were more
sensitive to the cytotbxic effects
of 03, and OKA was degraded to a
greater extent in the first 3 hr
after 03 exposure than strain KMBL
1788, which contains a normal DNA
Polymerase I.
E. coli
Hamelin et al.,
1977a
98,000 pg/m3 (50 ppa)
   MAST"
    30 min
Hucoid mutant strains (MQ 100 &
105) obtained by treating MQ 259
with 03 yet having full complement
of DNA repair enzymes were shown
to be more sensitive to Oa and
degraded DNA to a greater extent
than the Ion + (MQ259) strain.
E. coll
Hamelin et al.,
1977b
98,000 pg/m3 (50 ppra)
   MAST"
    up to 3 tirs         15 different DNA repair-deficient
                        strains were tested for sensitivity
                        to the cytotoxic effects of 03; DNA
                        Polytnerase I was involved in DNA
                        repair but Polymerases I and II
                        and DNA synthetic genes dna A,  B,
                        and C were not;  recombinational
                        repair pathways, assayed with rec
                        A and rec B strains,  were only
                        partially involved in the repair
                        of 03-induced DNA damage.
                                          E. coli
                  Hamelin and Chung,
                  1978

-------
                                TABLE 9-22.  CMROHQS0HM. EFFECTS FROM IN VITRO EXPOSURE TO HIGH       COHCINTRATIONS  (continued)
M
~J
oo
Oa Measurement
concentration nethod
0.1 ug/»l [51 pp«] UV
0.5 ug/ml [255 ppin] UV
0.18 fig/ml [92 ppn] UKI
1.0 |ig/ml [510 ppm]
0.5-6 ug/ml [255- NBKI
3061 ppn]
1-10 ug/nl [510- NBKI
5100 ppm]
5X [SO, 000 ppm] NBKI
3.5-4.5 Mi/«1 [1786- UKI
2296 ppm]
2% [20,000 ppm] e
8% [80,000 ppm] 'GPT
Exposure
duration and protocol Observed effect(s)
60 nln (70 8l/«1n) Preferential degradation of yeast
RNA at the N-glycosyl linkage;
sugar-phosphate linkage was
Oa stable.
30 rain (330 ml/rnin) 5-ribonucleotide guanosine
monophosphate was degraded
most rapidly.
NO Release of nucleic acids;
cell lethality.
0-5 nin Os reacts with pyrinndine
bases-froin nucleic acids
(thyinidine > cytosine >
uraci 1 ) .
30 nin Cell death and nonspecific
chromosomal aberrations:
shrunken and fragmented nuclei,
clumped metaphase chromosomes
and chromosome bridges.
0-40 min Rapid loss of glycolytic and
respiratory capacity; loss of
tumorigenicity after 20 min.
exposure.
ND Hitotic delay.
3 min Abnormal nuclei; fragmentation.
5, 15, 60 s Rapid degradation of nucleic
acid bases, nucleosides or
nucleotides in 0.05M phosphate
buffer, pH 7.2.
Species
Yeast
E. coli

E. coli

Chick
embryo
fibroblasts
Mouse
as cites
cells
Chortophaga
viridifaciata
Am. oyster
ND
Reference
Shiniki et al.,
1983
Scott and Lesher,
1963
Prat et al., 1968
Sachsenmaier
et al,, 1965
Fetner, 1963
Maclean et al . ,
Ii73
Christensen and
Giese, 1954
        Not ranked by air concentration;  listed by reported  exposure  values and [approximate ppm conversions].
        Measurement method:   MAST = Kl-coulometric (Mast meter);  NBKI =  neutral buffered potassium iodide; UV = UV photometry; GPT = gas phase titration;
        UKI = unbuffered potassium iodide.
       C03 flov* rate given in (ml/min):, when available.   ND  = not described.
        Concentrations of 03 were not measured in the cell suspensions.
       eOa analyzer (Fisher and Porter, Warminster,  PA).

-------
                                            TABLE 9-23.   CHROMOSOMAL EFFECTS FROM OZONE CONCENTRATIONS AT OR BELOW 1960 ug/m3 (1 Ppm)
H
»-J
VD
03
Concentration
M8/"i'i
294
412
1940
451
2548-
14,700
3234-
27,832
3920
392
470-
588
ppm
0.15
0.21
0.99
0.23
1.3-
7.5
1.65-
14.2
2.0
0.2
0.24-
0.3
Measurement3'
method
NBKI
NBKI
NBKI
NBKI
NBKI
MAST
UKI
MAST
UKI
Exposure0
duration and protocol
5 hr
5 hr
2 hr
5 hr
(in vitro)
ND
(in vitro)
ND
(in vitro)
5-90 rain
(in vitro)

5 hr
(in vivo)
5 hr
CLE vivo)
Observed effect(s) Species Reference
No effect induced by 03 treatment Mouse Gooch et a!,,
on the frequency of chromosome or 1976
chromatid aberrations in Chinese
hamster or mouse peripheral Hamster
blood lymphocytes stimulated with
PHA; no effect on spermatocytes
in mice 8 wk following exposure.
Peripheral blood lymphocytes exposed Human
to 03 in culture 12 hr after stlnr lymphocytes
ulation with PHA showed no increase
in chromosome or chromatid aberrations.
Peripheral blood lymphocytes exposed Human
to Oa in culture, 36 hr after PHA lymphocytes
stimulation, showed no change in the
frequency of chromosome or chromatid
aberrations at any concentration
except 7.23 ppm of Og.
No apparent increase in the fre- Human
quency of chromosome or chromatid lymphocytes
aberrations 12 or 36 hr after
PHA stimulation.
Combined exposure to 03 and radi- Hamster Zelac et al.,
ation (227-233 rad) produced an 1971b
additive effect on the number
of chromosome breaks measured in
peripheral blood lymphocytes.
A significant increase in chro- Hamster Zelac et al.,
mosorae aberrations (deletions, 1971a
ring dicentrics) in the peri-
pheral blood lymphocytes; in-
creased break frequency was
still apparent at 6 and 16 days
following exposure.

-------
                              TABLE 9-23.  CBRQMOSOHAL EFFECTS FROM OZONE CONCENTRATIONS AT OR BELOW 1860 (jg/ra3  (1 ppo)   (continued)
f
(-•
CO
o
03
Concentration
ug/P ppi
490- 0.2S-
1960 1.0
588- 0.3-
1568 0.8
843 0.43
3920 2.0
1960- 1.0-
9800 5.0
Beasurenent3' Exposure
method duration and protocol
UV 1 hr
(in vitro)

UV 8 days,
continuous
Cifi vitro)
UV, 5 hr
HBKI (in vivo)
6 hr
ijn v|yo)
CHEM 24 hr
NBKI (in vivo)
Observed effect(s) Species Reference
Dose- related increase in SCE fre- Hunan Guerrero et
quency in WI-38 diploid fibroblasts fibroblasts al., 1979
exposed in culture.
Growth of cells from lung, breast, Hunan Sweet et al.,
and uterine tumors were inhibited tumor 1980
to a greater degree than IHR-90, cells
a nontunor diploid fibroblast.
Increase in chromatid-type Hamster Tice et al.,
aberrations in peripheral blood 1978
lymphocytes of 03-exposed hamsters;
Increase in deletions at 7 days
and increase in achromatic lesions
at 14 days after exposure; ehronosoae-
type lesions were not significantly
different; no chromosomal aberrations
in bone Marrow lymphocytes; no change
in SCE frequency in peripheral blood
lymphocytes.
No change in SCE frequency in peri- Mouse
pheral blood lymphocytes.
Variable decrease in the molecular Mouse Chaney, 1981
weight of DNA from peritoneal exu-
date cells of 03 exposed mice
becoming significant at 5 ppm;
significant induction of single-
strand breaks at 5 ppm.
          Measurement method:   HAST = Kl-coulometric (Mast meter);  NBKI = neutral  buffered potassium iodide; UV = UV photometry

          Calibration method:   UKI = unbuffered potassium iodide;  NBKI  = neutral buffered potassiua iodide.

         CND = not described.

          Abbreviations used:   PHA = phytohemagglutinin;  SCE = sister chronatid exchange.

-------
     The molecular mechanism  for the clastogenic and  lethal  effects  resulting
from Qg exposure  are  not precisely known.  Bubbling  8 percent CL through a
phosphate buffer  solution  (Q.05M,  pH 7.2) containing DNA caused an immediate
loss in absorption at 260 nm and an increase in the absorption of the solution
at wavelengths  shorter  than  240 nm (Christensen and Giese, 1954).  A similar
rate of degradation was observed with RNA and the individual purine and pyrinr
idine bases,  nucleosides,  and nucleotides.   In a more recent report, Shiniki
et al.  (1983) examined  the degradation of a mixture of 5' nucleotides, yeast
t-RNA or  tobacco mosaic virus  RNA with 03 (0.1 to  0.5 mg/L).  The guanine
moiety was  found  to be  the most 0,.-labile among the  four  nucleotides, whether
the guanine was present as free guanosine monophosphate or incorporated into
RNA.  The  sensitivity to degradation by  03  among  the four  nucleotides was
found to be,  in decreasing order, GMP  >  UMP  > CMP >  AMP (GMP = guanosine
monophosphate,  U  =  uridine,  C = cytidine, A = adenosine).  Even after exten-
sive ozonolysis of yeast t-RNA  (0.5 |jg/ml,. 30 min) and substantial degradation
of the guanine  moieties,  the RNA migrated as a single band on polyacrylamide
gels.  The  band exhibited  the  same-mobility  as the  intact t-RNA,  indicating
that although the glycosidic bond between the sugar and the base is 03-labile,
the sugar-phosphate backbone was intact and extremely stable against 0-.  Prat
et al.  (1968) investigated the reactivity of  the  pyrimidines  in  E._  coll  DNA
with 0^  (0.5 to  6  mg/L, 0 to  5 min)  and  radiation.   Ozone preferentially
reacted with  thymidine, then  with  cytosine and uracil,  in decreasing order  of
reactivity.   The  results are slightly different from those reported by Shiniki
et al.  (1983) in  that the  reactivity with  uracil and cytidine" are in reversed
order.
     There  is evidence  that  single-strand  breaks in  DNA may contribute  to the
genotoxic  effects of  0-j.   Radiosensitive lex  mutants  of E.  coli,  which were
                       O                                  .~  '~—
known to be defective  in  the repair of x-ray-induced single-strand breaks in
DNA, were-found to be significantly more sensitive to the  cytotoxic effects  of
Qv than  the  repair-proficient  parental strain  (Hamelin and Chung,  1974).
     In an  effort to investigate the nature of the 0.,-induced lesion in DNA,
Hamelin  and co-workers investigated the  survival  of bacterial strains with
known defects in  DNA repair.   Closely  related strains of £_._  coli K-12 with
mutations  in  DNA  polymerase I were shown to be more sensitive  to the cytotoxic
effects of 0~ than the DNA  polymerase  proficient  (pol  +) strain  (Hamelin et
al.,  1977a).    Polymerase  I-deficient  strains also  exhibited an  extensive
                                   9-181

-------
degradation of DNA in response to 03 or x-ray treatment.  The authors concluded
that DNA polymerase I plays a key role in the repair of lesions produced in E.
coli DNA by 0_ and that the unrepaired damage was responsible for the enhanced
™nmmr::r~--n         £
degradation of DNA and the enhanced cell killing observed in the pol- mutants.
This interpretation of the data may not be entirely correct, because an enhanced
degradation of  DNA and an increased sensitivity  to cell  killing were  also
observed in a Ion  mutant  strain  of  £._  coli  K-12  (Hamelin  et  al., 1977b).   The
Ion mutant appears to have a full  complement  of DNA repair enzymes.   With
these mutants,  there  may  be an enhanced DNA repair  activity  (evidenced  by  the
extensive degradation of DNA),  and the enhanced activity of  the  Ion gene
products was thought to be responsible for the increased cell killing observed
with these strains when they were exposed to 03.
     Although DNA  polymerase I  was shown  to be involved  in the  repair of
0,-induced DNA  damage (Hamelin  et  al.,  1977a),  JL_ coli  cell  strains with
mutations in DNA  polymerase  II  or  III were not found to be more sensitive to
0, than  the wild-type,  suggesting that these enzymes are not involved in the
repair of DNA damaged by 0, (Hamelin and Chung,  1978).  Mutant strains of g._
coli with defects  in  DNA  synthesis  (DNA A,  B,  C, D,  and G) showed no  enhanced
sensitivity to CU.  Therefore, the DNA gene products are probably not involved
in the repair of  (U damage.   Recombinational repair mutants, rec A  and  rec B,
only showed a slightly increased sensitivity to 0~ than the wild-type, suggest-
ing that the  rec  gene products  are  only partially  involved  in  the  repair  of
Og-induced DNA lesions (Hamelin and Chung,  1978).
     Other effects  have been  observed  than  those described above on bacteria.
In the commercial  American oyster exposed to CU-treated sea  water (Maclean et
al., 1973),  fertilization occurred  less readily and abnormal nuclei  (degenera-
tion, fragmentation) were observed approximately twice as frequently.   Sachsen-
maier et al.  (1965) observed a rapid loss of glycolytic and respiratory capacity
and subsequent  loss of tumorigenicity  in mouse ascites  cells treated  with  CL.
These authors also reported  that chicken embryo fibroblasts exposed to CU (1
to 10 ul/ml) for  30 min exhibited nonspecific  alterations  in cells  resembling
those seen after  x-ray  damage,  including shrunken  nuclei, clumped  metaphase
chromosomes,   arrested mitosis,   chromosome  bridges  and  fragmented  nuclei.
     In the studies described up to this point, the investigators have predomi-
nantly examined the  in  vitro effects  of extremely  high Og concentrations  on
biological  systems  or biologically  important cellular  components.  Although
                                   9-182

-------
these investigations  may  be important for the  elucidation  of the types of
damage or responses that might be expected to occur at lower CL concentrations,
the most  relevant data on  the  genotoxicity  of CL should be  obtained from
                                                                   3
investigations where  the  03 concentration did not exceed 1960 (jg/m  (1 ppm).
Research, conducted at or  below  1 ppm  of 0, will be presented  below  (See Table
9-24).
     Several investigators have examined the j_n vivo cytogenetic effects of 03
in rodents  and  human  subjects.   Until the reports of Zelac et al. (1971a,b),
the toxic effects  of  Q» were generally assumed to be confined to the tissues
directly in  contact with  the  gas,  such as the  respiratory epithelium.  Due to
the highly  reactive  nature  of CU,  little systemic absorption was predicted.
Zelac, however,  reported  a  significant increase in chromosome aberrations in
                                                                       3
peripheral  blood  lymphocytes  from  Chinese hamsters exposed to 392 pg/m  (0.2
ppm) for 5 hr. Chromosome breaks, defined as the sum of the number of deletions,
rings and dicentrics, were  scored  in lymphocytes collected immediately after
0» exposure  and  at 6 and 15 days postexposure.  At all sampling times, there
was an increase in the break frequency (breaks/ cell) in the_03-exposed animals
when  compared with nonexposed control  animals.  Zelac  et al.  (1971b)  reported
that  Q3 was additive  with  radiation  in the  production  of chromosome  breaks.
Both  0» and radiation produced  chromosome breaks  independently  of each other;
Simultaneous  exposure to 0.2 ppm  of  03  for  5 hr and  230  rad of radiation
resulted  in  the  production  of 40 percent more breaks than were expected, from
either agent alone and  70 percent  of  the  total  number  of.breaks  expected  from
the combined effects  of the two agents,  if  it was assumed that the effects
were .additive.                                                  .
      Chaney  (1981) investigated the effects of 03 exposure on mouse peritoneal
exudate cells (peritoneal  macrophages) stimulated  by an i.p.  injection  of
                                                                            3
glycogen.    Mice  were  subsequently  exposed by inhalation to  1960  or  9800 ug/m
(1 or 5 ppm) of 0., for 24 hr.  A significant reduction  in the average molecular
weight  of the DNA was  observed in the peritoneal exudate  cells from mice
                     3                         '                             3
exposed to  9800  ug/m  (5 ppm) of  03  but not in animals exposed to 1960 ug/m
(1 ppm) of OT for  24  hr.  The reduction in the average molecular weight of DNA
in 03~exposed animals indicated the  induction of single-strand breaks in the
DNA.  It  should  be noted,  however, that the alkaline sucrose gradient method
of determining  the average  molecular weight of the DNA does  not discriminate
between the frank strand-breaks in DNA, produced  as  a direct  effect of the 03
treatment,  and  the induction of alkaline-labile  lesions  in DNA by  03,  which
                                   9-183

-------
                                                            TABLE 9-24.  MUTAT10NAL EFFECTS OF OZOHE
     concentration       Measurement
ug/i3     pjw
                           method
                        Exposure      .
                 duration and protocol
      Observed effect(s)
Species
Reference
t->
00
196 0.1 HAST
58,800 30 NBKI
98,000 50 UV
60 rain
(2.1 nl/min)
3 hr
30 rain
(2.1 ml/min)
Various mutated, growth factor auto-
trophic states, were recovered; mutant
strains differed from parental strains
in sensitivity to UV light and excessive
production of capsular polysaccharide.
Induction of a dominant lethal muta-
tion during stages of spermatogenesis;
spern were found to be twice as sen-
sitive as earlier stages.
Radiation-sensitive mutant strains
defective in repair of single strand
Escherichia coli
(HQ 259)
Drosophila
virilis

Saccharomyces
cerevisae
Hamelin and
Chung, 1975a
Erdman and
Hernandez, 1982
Oubeau and
Chung, 1979
                                              (rad 51)  and  double  strand (rad  52)
                                              DMA breaks were more sensitive to
                                              03  cell killing than either the
                                              wild-type or  the UV  light  repair
                                              deficient strain (rad 3);  recom-
                                              binational repair-deficient strain
                                              (rad 6) was moderately sensitive
                                              to  03.
     98,000    50
  UV
                         30 Bin
Induction of forward mutations at 2
loci of adenine biosynthesis (strain
C16-11C); induction of reversion
nutation at 6 genetic loci (strain
XV 185-14C); induction of intra-
genic and intergenic recombinational
mutants (strain 07); 03 was 2Q-20Gx
less mutagenic than equitoxic expo-
sures to UV light, x-rays, or MNNG.
Saccharomyces
cerevisae
Dubeau and
Chung, 1982
      Measurement method:

      QS flow rates  given  in  (ml/min), where available.
HAST = Kl-coulometric (Mast meter);  NBKI  = neutral - buffered potassium iodide;  UV =  UV  photometry

-------
would be converted to strand breaks under the alkaline condition of the assay.
Although these  experiments  do not prove that  0~  exposure can cause strand
                                                                 3
breaks in DNA,  they do indicate an 03 effect on DNA at 9800 ug/m  (5 ppm) of
03 for 24 hr.
     Because of the importance  of the reports by Zelac et al. (1971a,b) that
indicated that significant levels of chromosome aberrations in Chinese hamster
peripheral  blood lymphocytes collected as late as 15 days after 0~ exposure by
inhalation,  Tice et al.  (1978)  tried to repeat the  experiments  of Zelac as
                                                               3
closely as possible.  Chinese hamsters were exposed to 843 ug/m  (0.43 ppm) of
0- by  inhalation for 5  hr.  The  authors  investigated  chfomatid and chromosome
aberrations  in  peripheral  blood  lymphocytes and bone  marrow  of  control and
0~-exposed animals  immediately after exposure and at  7  and 14 days  after  0-,
 o              '         -  •                           •                   t *5
exposure.  They also  investigated the sister chromatid exchange (SCE) frequency
in peripheral blood  lymphocytes of Chinese hamsters.   In separate experiments,
SCE  frequencies in  C57/B1 mice exposed to 2 ppm of Q3 for 6 hr were examined
in peripheral  blood  cells collected  from  the animals immediately after  0-,
exposure and at 7 and 14 days after 03 exposure.
     The authors  reported  no  significant increase  in  the SCE  frequency  of  the
Q3-exposed hamsters or mice at any  sampling  time,   nor did they observe  a
significant  increase  in the number of chromosome aberrations of phytohemagglu-
tinin  (PHA)-stimulated peripheral blood or bone marrow cells.   The only report-
ed statistically  significant differences were  observed in peripheral  blood
lymphocytes, in which there was an increase in the number of chromatid deletions
and  achromatic  lesions in the 7- and 14-day samples, respectively.  Both types
of chromatid aberration were observed at consistently higher frequencies  in
the  blood  samples  of the 03-exposed  animals,  frequently in the  range of 50 to
100  percent  increases over the control values.  Statistically significant  dif-
ferences were  assigned at the 1 percent level  of significance.   It is  not
clear  how a  slightly more rigorous evaluation of significance (e.g., p < 0.05)
would  have influenced the interpretation of the data.
     Although both  Zelac et al.  (1971a,b) and Tice et  al.  (1978) reported  sig-
nificant increases  in chromosome aberrations in peripheral blood lymphocytes
                                           3
following 5-hr  exposures to 392  to 843 ug/m   (0.2 to 0.4 ppm) of 03, the types
of lesions  observed  in the two  studies  were clearly  different. Tice et al.
observed chromatid-type  lesions  and  no increase in the chromosome aberrations,
whereas Zelac et al.  reported a  significant  increase  in  the number of chromosome
                                    9-185

-------
aberrations.  There were a number of differences in the experimental protocols
that may have produced the seemingly different results:

     1.   The animals  were exposed to  different  concentrations  of CL.
          Zelac et al.  (1971a) administered 470 to 590 ug/m3 (0.24 to 0.3
          ppm) of  Go to  Chinese hamsters for  5 hr, whereas Tice  et al.
                   o                                         ,,
          (1978) exposed animals to an atmosphere of 840 \ig/m  (0.43 ppm)
          of Do for 5 hr.
     2.   Zelac stimulated peripheral blood lymphocytes into DNA synthesis
          with pokeweed  mitogen, which  is mainly  a B-lymphocyte mitogen.
          In  the  experiments of Tice  et al.  (1978),  lymphocytes were
          stimulated with  PHA, which is  a T-lymphocyte mitogen (Ling and
          Kay, 1975).
     3.   Zelac cultured lymphocytes with the mitogen jn vitro for 3 days
          (72 hr), whereas Tice et  al.  cultured lymphocytes with  mitogen
          for 52 hours.

Because of  the  longer  incubation  time with the mitogen in the experiments of
Zelac  et al.  (1971a),  lymphocytes  may have converted  chromatid type aberra-
tions, like those  reported by Tice  et al. (1978), into chromosome aberrations
with another  round of  DNA  synthesis.  Because  the experiments of  Zelac et al.
and Tice et al.  were conducted with peripheral blood lymphocytes stimulated by
                *
two different mitogens (pokeweed vs. PHA), the cytogenetic consequences of 03
exposure were examined in different populations of lymphocytes.   If one of the
populations of lymphocytes was more sensitive  to  03 than the other, different
cytogenetic responses  could  be  expected when PHA was used as a mitogen,  com-
pared with the results with the use of polkweed mitogen.
     There  is evidence that  the B-lymphocyte may  be more sensitive to  0~ than
the T-lymphocyte.  Savino  et al.  (1978) measured the  effects of  03 on human
cellular and humoral  immunity by measuring rosette formation with human lympho-
cytes  (See  Chapter 10).  Rosette formation measures the reaction  of antigenie
red cells  with  surface membrane sites  on  lymphocytes.   Different antigenie
RBCs are used to distinguish T-lymphocytes from B-lymphocytes.   Rosette forma-
tion with  B-lymphocytes  was  significantly depressed in eight human subjects
                   3
exposed to 784 yg/m  (0.4 ppm) of 03 by inhalation for 4 hr,   A similar inhibi-
tion of rosette  formation  was not  observed with  T-lymphocytes from the same
                                   9-186

-------
subjects.   The  depressed B-cell  responses  persisted for 2 weeks  after 0^
exposure,  although partial  recovery to the pre-exposure  level  was evident.
     It cannot be  stated with any certainty how  the  differences in the CU
exposure,  the choice  of  mitogen,  and the length  of the mitogen  exposure may
have contributed to the  differences in the results reported by  Zelac et al.
(1971a,b)  and by Tice et al.  (1978).  There are sufficient differences  in the
experimental  protocols of the two reports so that the results need not be con-
sidered directly contradictory.
     An assumption that  is made in all of the reports in which lymphocytes are
stimulated with mitogens is that the lymphocytes from the CL-exposed animals
and the control  animals ara equally sensitive to the mitogenic stimulus.  This
assumption is probably  not correct, because in investigations by Peterson et
al. (1978a,b) the proliferation of human lymphocytes exposed to PHA was signi-
ficantly suppressed in blood samples taken immediately after the subjects were
                   3
exposed to 784 ug/m   (0.4 ppm)  of 0,  for 4  hr  (See 10.7).  Other reports have
suggested that 0^  might  inhibit or  inactivate  the PHA receptor  on  lymphocytes
(see Gooch et al., 1976).  Because the ability  to  measure chromosome aber-
rations in mitotically arrested cells is absolutely dependent on the induction
of Gg  or  G-^  cells  into the cycling  state, cells exposed to sufficient concen-
trations  of  Og would  not be stimulated to divide, and hence no  O^-induced cy-
togenic effects would be observed in activated cells.  In their report, Tice
et al.  (1978) stated  that the  lymphocytes of the Q~-exposed animals in their
experiments  "did tend to be worse than those  from  controls."  Only a small
difference in the  number  of responding lymphocytes could make large differences
in  the results  of the experiments  if Q~~damaged lymphocytes were  selected
against in the cytogenetic  investigations.
     In other investigations  with rodents,  Gooch et al. (1976) analyzed bone
                                                        3
marrow samples from Chinese hamsters exposed to 451 ug/m  (0.23 ppm) of 0^ for
5  hr.   Marrow  samples were taken at  2,  6,  and 12 hr  following  0,  exposures.
                                                                      3
In  separate  experiments, male C3H mice were exposed to  294 or 412  ug/m   (0.15
or  0.21 ppm) of 03 for  5 hr, or to 1940 ug/m3 (0.99 ppm) of 03 for 2 hr.
Blood  samples were drawn from these animals at various times for up to 2 weeks
following 0, exposure.   The mice were killed 8 weeks following 0~ exposure, and
spermatocyte preparations were  made and analyzed  for reciprocal translocations.
Data from the Chinese hamster bone marrow samples and the mouse  leukocytes in-
dicated that there was  no  effect induced by 0~ treatment on the frequency of

                                    9-187

-------
chromatid or chromosome aberrations, nor were there any recognizable reciprocal
trans! ocations in the primary spermatocytes.
     Several investigators  have  examined the effects of 0., on human cells in
                                                          C3                —
vitro.  Fetner  (1962)  observed the induction of chromatid deletions in human
KB cells exposed  to  15,680  pg/m3  (8 ppm)  of Q3  for  5  to 25 min.   Sweet et al.
(1980) reported that the  growth of human  cells  from breast,  lung,  and  uterine
                                                     a
tumors was  inhibited by exposure to 588 to 1568 (jg/m  (0.3 to 0.8 ppm) of 0,,
                                                                            «3
for 8 days in culture.
     Guerrero et al.  (1979) performed SCE analysis on diploid human fetal lung
cells (WI-38) exposed to 0, 490, 980, 1470, or 1960 yg/m3 (0, 0.25, 0.5, 0.75,
or 1.0 ppm)  Oo  for 1 hour in vitro.  A dose-related increase in the SCE fre-
quency was observed in the WI-38 human fibroblasts exposed to Og.  In the same
report, the  authors  stated that no significant increase in the SCE frequency
over control values was observed in peripheral blood lymphocytes from subjects
exposed to  03 by  inhalation (Chapter 10).   Unless the  lymphocyte  is  intrinsi-
cally less  sensitive to the induction of  SCE  by 03  than the  WI-38  human fetal
lung fibroblast,  the results indicate  that exposure of human subjects to 980
    2
ug/rn  (0.5  ppm) of 03  for  2  hr  did not result in  a  sufficiently  high  con-
centration of 03  or  Oo reaction products in the circulation to induce an in-
crease in the SCE frequency in the lymphocytes.   From the authors' data on the
induction of SCE  in  WI-38 cells, the concentration of 0« required to  induce
                                            3
SCE in human cells is approximately 490 (jg/m  (0.25 ppm) for 1 hour.
     Gooch et al.  (1976) also investigated the effects of 0- exposure on human
cells IB yjtrjo.   In  these experiments,  lymphocytes were stimulated with  PHA
for  12 or  36  hr  before the 03 exposure to obviate the potential problems of
CL inactivation of the PHA receptor.  Human leukocyte cultures v/ere exposed to
         3
3920 ug/m   (2 ppm)  of  03 for  various lengths of time to accumulate total  0~
exposure doses  of 3234 to 27,832 HS/m  per hour (1.65 to 14.2 ppm/hr).   The
results showed  no  increase  in the chromatid  and  chromosome  aberrations  at any
total  dose,  with the possible exception  of  an apparent spike in  chromatid
                                              2
aberrations at a total exposure of 14,170 \ig/m  (7.23 ppm/hr).  The significance
of this observation is unclear because the data showed no dose-response increase
in the number of chromatid aberrations at concentrations near 7.25 ppm/hr, and
the authors  did not report how, or  indeed  if, the data were  statistically
evaluated for differences in chromatid aberrations.
     In summary, in vitro 0» exposure  has been shown to  produce toxic effects
                 _"•» — — —   $
on cells and cellular components including the genetic material.   Cytogenetic
                                   9-188

-------
toxicity has been  reported  in cells in culture  and  in cells isolated from
animals if 0- exposure has occurred .at sufficiently high levels and for suffi-
ciently long periods,
9.4.5.2  Mutational Effects of Ozone.  The  mutagenic  effects of 0,, have been
investigated in surprisingly  few  instances  (Table 9-24).  No publication to
date has investigated the mutagenic effects of 03 in mammalian cells.
     Sparrow and Schairer  (1974)  measured an increase in the frequency,  over
the background  level,  in  the  induction of  pink  or colorless mutant cells or
groups of cells in petal  and/or stamen hairs  of mature flowers of various
blue-flowered Tradescantia.   No 03 concentration was reported in this publica-
tion.
     LL coli, strain MQ 259, were mutated to various growth factor auxotrophic
states,  including  requirements for  most  common  amino acids5 vitamins, and
                                                                           3
purines and  pyrimidines (Hamelin  and Chung,  1975a).   Ozonated air  (196 |jg/m  ,
0.1 ppm) was passed through the bacterial suspensions at a rate of 2.1 L/min
for 30  min.   Many  of the Q~-induced mutant strains were either more or less
sensitive to UV light than  the  parental strain.   Other  mutant strains, called
mucoid mutants, had apparent  defects in DNA repair pathways and were charac-
terized  (Hamelin and  Chung, 1975b)  as producing  excessive  amounts  of capsular
polysaccharide.
     Erdman  and Hernandez (1982) investigated the induction of dominant lethal
                                                       2
mutations in Drosophila virilis exposed to 58,800 |jg/m  (30 ppm) of 0, for 3
hr. 0™ induced dominant lethal mutations at various stages of spermatogenesis.
The sperm-sperm bundle stage  was the most  sensitive  to 0-,,  and the meiotic
cells were the least  sensitive,
     Dubeau  and Chung (1979,  1982)  have investigated the mutagenic and cyto-
toxic  effects  of  0-  on Saccharomycescerevisae.   Several  different strains
were utilized  to investigate  forward, reverse, and  recombinational  mutations.
                   3
ozone  (98,000  yg/m ,  50  ppm;  30  to  90  min) induced a variety of  forward  and
reverse  mutations  as  well  as  gene conversion and mitotic crossing-over.   Both
base-substitution  and frame-shift mutations were induced  by 0,.   Ozone was
shown  to  be more recombinogenic than mutagenic in yeast, probably as a  result
of  the induction  of  strand breaks  in  DNA,  either  directly  or indirectly.
     In the  investigation of Dubeau  and Chung  (1982), the mutagenic potency of
                3
0~  (98,000 M9/m .  50  PP^; 30 to 90 min) was  compared with other known mutagens.
                                               2
The positive controls were UV  light (1.54  J/m  per second,  1-min exposure),
N-methyl-N'-nitrosoguanidine  (MNNG)  (50 yg/ml, 15 min), or x-rays  (2 kR/min,
                                   9-189

-------
40 min).   By comparing the  induced  mutation frequency at similar cellular
survival levels for Q3, MNNG, UV light, and x-rays, it was shown that 0- was a
very weak  mutagen.   Induced mutation  frequencies  were generally 20 to 200
times lower for 03 than for the other three mutagens.
     In  summary,  the mutagenic properties of  0~  have been demonstrated  in
procaryotic  and  eucaryotic cells.   Only  one  study,  however,  (Hamelin and
Chung, 1975a, with Ju_ coll) investigated the mutagenic effect of 0-, at concen-
trations of  less  than  1 ppm.  The results clearly indicate that if cells in
culture are exposed to sufficiently high concentrations of 0_ for sufficiently
long periods, mutations will result.   The relevance of the presently described
investigations to human or even other  mammalian mutagenicity is  not apparent.
Additional studies  with human  and  other mammalian cells  will  be required
before the mutagenic potency of 03 toward these  species  can be determined.

9.4.6  Qther Extrapulmonary' Effects
9.4.6.1   Liver.   A  series of studies reviewed by Graham et al. (1983a)  have
shown that 0,  increases drug-induced sleeping time in animals  (Table  9-25).
The animal was injected with  the  drug  (typically pentobarbital),  and the  time
to the  loss  of the righting  reflex  and the  sleeping  time  (time  between loss
and regaining  of  the righting reflex)  were measured.   Because the  time to the
loss of the righting reflex was very rarely altered in the experiments described
below,  it will  not be  discussed  further.  Animals awake from pentobarbital-
induced sleep, because liver xenobiotic metabolism transforms the drug into an
inactive form.  Therefore, this response is interpreted as an extrapulmonary
effect.
     Gardner et al. (1974) were the first to observe that 03 increases pentobar-
bital-induced sleeping  time.   Female CD-I mice were exposed for 3 hr/day for
                           3
up to  7 days to 1960 yg/m  (1.0  ppm)  of 03 and the  increase  was found  on
days 2 and 3 of exposure,  with the greatest response occurring on day  2.  Com-
                                                                    o
plete tolerance did not occur; when mice were pre-exposed (1960 (jg/m , 1.0 ppm;
3 hr/day  for 7 days) and  then  challenged with a 3-hr exposure  to 9800 pg/m
(5.0  ppm)  on the eighth day,  pentobarbital-induced sleeping time increased
greatly.
     A  series  of  follow-up studies was conducted  to  characterize the  effect
further.  To  evaluate  female  mouse strain sensitivity, one outbred (CD-I) and
two inbred (C57BL/6N and DBA/2N)  strains were  compared (Graham  et al., 1981).
                                   9-190

-------
                                                         TABLE 9-25,   EFFECTS OF OZONE ON THE LIVER
Ozone
concentration
ug/n3 ppm
196- 0.1-
9800 5






Measurement3' Exposure
method duration and protocol
CHEH 3 hr/day,
GPT 1-17/days







Observed effects(s)
Increase In pentobarbital- induced
sleeping time with following expo-
sure regimens; 0.1 ppm, IS or 16 days;
0,25 ppm, 6 or 7 days; 0.5 ppn,
2 or 3 days; 1 ppm, 1, 2, or 3 days;
S ppm, 1 day. No effects at days
before or after days given above.


Species Reference
House Graham et al . ,
(female) 1981





H
588
1470
5880
19,600
1600
(max)
1960
1960
0.3
0.75
3
10
0.82
(max)
1
1
UV
NBKI
UV
ND
CHEM
GPT
3 hr
3 hr
4 hr
(max)
3 hr/day
5 hr
No effect on liver reduced ascorbic
acid levels.
No effect on hepatic benzo(a)pyrene
hydroxylase activity.
Decrease in hepatic reduced ascorbic
acid content. Actual exposure regi-
mens not reported, only maximal
levels given.
Increase in pentobarbital -induced
sleeping time after 2 or 3 days,
but not other days (up to 7 days).
No tolerance to a challenge of 03
(9800 MS/1"3. 5 ppm x 3 hr) to mice
pre-exposed to 1 ppn for 7 days.
Increase in pentobarbital-induced
sleeping time in 3 strains of female
Rat
Hamster
Mouse
House
(female)
Mouse, rat,
hamster
Calabrese et al. ,
1983c
Palmer et al. ,
1971
Veninga et al. ,
1981
Gardner et al . ,
1974
Graham et al . ,
1981
                                                                              mice (CD-I,  C57BL,  and D1A),  female
                                                                              rats,  and male and  female hamsters.
                                                                              Male mice and rats  not affected,
                                                                              even when male mice received  3 days
                                                                              (5 hr/day) of exposure.   The  increase
                                                                              in male hamsters  was less than the
                                                                              increase in  female  hamsters..
(male and
female)

-------
                                                    TABLE 9-25,   EFFECTS OF  OZONE ON THE LIVER  (continued)
Ozone
concentration
jjgTS8 ppi
1960 1
I960 1
Measurement3'
method
CHEH
GPT
CHEH
GPT
Exposure
duration and protocol
5 hr/day,
1,2,3 or
4 days
5 hr
Observed effects(s)
Increase in pentobarbital- induced
sleeping time at 1,2, and 3 days,
decreased with increasing days of
exposure. 24-hr postexposure for
each group, no effects occurred.
Increase in hexobarbital- and thiopen-
tal-induced sleeping tine and zoxazo-
Species
House
(female)
House
(female)
Reference
Graham et al . ,
1981
Graham et al . ,
1982a
&
to
lamine-induced paralysis time.   Pre-
treatment with mixed function oxidase
inducers (phenobarbital, pregneolone-
I6a-carbon1trile, and p-naphthofla-
vone, but not pentobarbital} decreased
phenobarbital-induced sleeping time in
CD-I mice, and 03 increased the sleeping
tine in all groups.   Pretreatntent with
inhibitors (SF525A,  piperonyl butoxide)
reduced the sleeping time, but 03 increa-
sed the sleeping time, with the magnitude
of the increase becoming larger as the
dose of inhibitor was increased.
1960 1 CHEH 5 hr
9800 5 GPT 3 hr
No effect on hepatic cytochrome Mouse Graham et al . ,
p-450 concentration, aminopyrine (female) 1982b
N-demethylase, or p-nitroanisole
0-demethylase activities. Aniline
hydroxylase activity increased at
5 ppm (3 hr) and at 1 ppm (5 hr/day
x 2 days). No change in liver
to body weight ratios.

-------
                                                TABLE  9-25.   EFFECTS OF  OZONE ON THE LIVER   (continued)
UD
H
VO
U>
Ozone
concentration
ug/n3 ppm
1960 1
9800 5
1960 1
3920 2
Measurement*' Exposure
method duration and protocol Observed effects(s) Species
CHEH 5 hr At 1 ppm: 71% increase in plasma House
GPT half-life of pentobarbital , decrease (female)
(p = 0.06) in slope of clearance curve.
At 5 ppm: 106% increase in plasma
half-life of pentobarbital; decrease
in slope of clearance curve; no
effect on concentrations of pen-
tobarbital in brain at time of
awakening; no change in type of
pentobarbital metabolites in
serum/or brain.
ND 90 rain No effect on hepatic cytochrome Rabbit
P-450 concentration.
UV 8 hr/day Supplementing or depriving rats of Rat
vitamin E or selenium altered the
03 effect. 03 caused changes in
several |n vitro enzyme activities
in the iTver and kidney (see text).
Reference
Graham 1979;
Graham et al . ,
1983, 1985
Goldstein
et al., 1975
Reddy et al , ,
1983
aMeasureraent method:   CHEM = gas-phase cheniluminescence;  NBKI  - neutral  buffered potassium  iodide;  UV = UV photometry;  ND = not described
 Calibration nethod:   GPT = gas phase titration

-------
                3
Ozone (I960 jjg/m  , 1.0 ppm  for 5  hr)  increased pentobarbital-induced sleeping
time in all  strains.   To determine whether  this effect was sex- or species-
specific,  male and female CD-I mice,  rats, and hamsters were exposed for 5 hr
to 1960 |jg/m3 (1.0 ppm) of Og  (Graham et al., 1981).  The females  of all
species exhibited an increased pentobarbital-induced sleeping time.  Male mice
and rats were not affected.  Male hamsters had an increase in sleeping time,
but this  increase was less (p =  0.075) than the  increase  observed  in the
females.   Thus, the  effect  is not specific  to strain of mouse or to  three
                                  »
species of animals, but it is sex-specific, with females being more susceptible.
Female CD-I mice were exposed to  196  to 9800 yg/m   (0.1 to  5.0 ppm) of 0Q  for
                                                                           3
3 hr/day for  a varying  number of days  (Graham et  a!., 1981).  At  1960 |jg/m
(1.0 ppm), effects were  observed  after 1, 2, or 3 days of exposure,  with the
                                                 3
largest change  occurring on  day  2.   At 980  (jg/m   (0.5  ppm),  the greatest
increase in pentobarbital-induced sleeping time was  observed on day 3, but at
        o
490 |jg/m   (0.25 ppm),  6 days of exposure were required to cause an increase.
At the lowest concentration evaluated  (196 |jg/m3, 0.1 ppm), the increase was
only observed at days 15 and 16 of exposure.   Thus, as the concentration of Q~
was decreased, increasing  numbers of daily  3-hr exposures  were  required to
significantly increase pentobarbital-induced sleeping time.  Once  the  maximal
effect occurred,  increasing the  number of exposures resulted in a diminution
of the effect.  Generally,  effects were observed around an approximate C X T
(concentration, ppm  x  time, hr)  value  of 5.   Also, CD-I female mice were ex-
                                                              2
posed to four different concentrations  of 03  (1960 to 196 ug/m ,  1.0 to 0.1 ppm)
for 5 to  20  hr continuously in a fashion yielding a C x T value of 5 (Graham
et al., 1981).  All  regimens increased pentobarbital-induced  sleeping time.
                                                               3
The time  to  recovery was examined in mice exposed to 1960 ug/m  (1.0 ppm) of
0,~ for 5  hr/day.   Recovery was complete within 24 hr after exposure, whether
exposure was  for 1, 2, or 3 days.
     To determine whether  the previous responses were specific to pentobar-
bital, other  drugs with known and different mechanisms  for termination of
                                                         3
action were used in CD-I female mice exposed  to 1960 ug/m  (1.0 ppm) of Og for
5 hr  (Graham  et  al., 1982a).  Ozone increased sleeping time induced by hexo-
barbital and  thiopental  and paralysis  time  induced  by zoxazolamine.   Within
the liver, pentobarbital and  hexobarbital metabolism is more related to  cyto-
chrome P-450  than to  cytochrome  P-448-dependent  activities.  Zoxazolamine
metabolism is more  related to  cytochrome P-448  than to cytochrome P-450.
                                   9-194

-------
Other major differences  in  the nature of the metabolism (i.e., aliphatic vs.
aromatic) also exist  between  these three drugs.  In  contrast  to the above,
sedation from thiopental  is terminated because  of drug redistribution.  Thus,
it would appear that  03  might  affect  some aspects of  both drug redistribution
and metabolism.   Although there are different mechanisms involved in hexobarbital,
zoxazolamine,  and pentobarbital  metabolism,  it is possible that some common
component(s) of metabolism may have been altered.
     CD-I female mice were pretreated with mixed-function oxidase inducers and
inhibitors  with partially characterized mechanisms of action  to relate any
potential differences in the effect of 0, to differences in the actions of the
agents.  Mice were  exposed  to 1960 ug/m  (1.0  ppm)  of 03 or  air for 5 hr
before  measurement  of pentobarbital-induced  sleeping time (Graham et  al.,
1982a).  Again, the effect  of  0-  was  observed,  but mechanisms  were  not  eluci-
dated.
     The effect of  03 on hepatic mixed-function oxidases in CD-I female mice
was  evaluated  in  an attempt to relate to  the  sleeping time studies  (Graham  et
al. ,  1982b).  A  3-hr  exposure to  concentrations  of  03 as  high as 9800  [jg/m
(5.0  ppm)  did  not change the  concentration  of cytochrome  P-450 or the activi-
ties of related enzymes  (aminopyrine  N-demethylase or p-nitroanisole 0-demethy-
                                                               o
lase).   However,  this exposure regimen and another  (1960  [jg/m , 1.0 ppm, 5
hr/day  for  2  days)  increased  slightly the activity of another mixed-function
oxidase, aniline  hydroxylase.   Goldstein et al. (1975) also found no effect of
a 90-min  exposure to  1960  ug/m   (1.0 ppm)  of 0- on  liver cytochrome P-450
levels  in rabbits.  Hepatic benzo(a)pyrene hydroxylase (another  mixed-function
oxidase)  activity of  hamsters was unchanged  by  a  3-hr exposure to  up to
19,600  ug/m3 (10  ppm) of QS (Palmer et al., 1971).
      Pentobarbital pharmacokinetics in female CD-I mice were also examined.  A
3-hr exposure  to  9800 ug/m   (5.0  ppm)  of 03  did  not affect  brain concentra-
tions  of pentobarbital  at-time  of  awakening,  even  though sleeping time was
increased  (Graham et  al. , 1985).   Therefore,  it appears  that 03  did not alter
the  sensitivity  of brain receptors  to  pentobarbital.   A similar  exposure
regimen also  did not alter the pattern  of brain or plasma metabolites of
pentobarbital  at  various times  up to 90 min postexposure (Graham et  al.,
1985).   Following this exposure,  first-order  clearance kinetics  of  pentobarbital
were observed in both  the  air and 0, groups,  and  0, increased  the  plasma
half-life  by  106 percent (Graham  et  al. ,  1985).   Mice exposed to 1960 ug/m3
(1.0 ppm)  of 0-, for 5 hr had  a 71 percent increase in the plasma half-life of
                                    9-195

-------
pentobarbital.   This ozone  exposure  resulted in a decrease (p = 0.06) in the
slope of the clearance curve.  Clearance followed first-order  kinetics with a
one-compartment model in this experiment also.
     In summary, the mechanism(s) for the effect of 03 on pentobarbital-induced
sleeping time are not known definitively.  However,  it is hypothesized (Graham
et al., 1983) that  some common aspect(s) of,  drug metabolism is quantitatively
reduced, whether it is direct (e.g.,  enzymatic) or indirect (e.g., liver blood
flow) or a combination of both.   In addition, drug redistribution is apparently
slowed.  It is unlikely that ozone itself caused these effects at target sites
distant from the  lung  (see Section 9.2).  Because of the free-radical nature
of oxidation  initiated  by 0~, a myriad  of oxygenated products, several of
which  have  toxic potential,  may be  formed  in the  lung (Section 9.3.3).
However, the stability  of such  products in the blood and their reaction with
organs such as the liver are speculative at present.
     Reddy et al.  (1983) studied the  effects  of a 7-day (8 hr/day) exposure of
rats to 3920 ug/m3  (2.0 ppm)  of  03 on  liver  xenobiotic metabolism by  perform-
ing jm  vitro  enzyme assays.  Although lower 03 levels were not tested, this
study  is presented  because it indicates the  potential of 0-, to cause hepatic
and kidney  effects.  The  rats used were either supplemented or deficient  in
both vitamin E  and  selenium.   Ozone  exposure caused a decrease in microsomal
cytoehrome  P-450  hydroperoxidase activity  in livers  of rats deficient in both
substances, whereas  an  increase  resulted in  the supplemented  animals.  Rats
deficient in vitamin E and selenium experienced a decrease in liver microsomal
epoxide hydrolase activity after 03 exposure; no effect was observed in supple-
mented  rats.  Glutathione  S-transferase  activity  was increased in  the  liver
and kidney in both the supplemented and deficient groups.  Selenium-independent
glutathione peroxidase activity  was  not significantly affected in the livers
of the supplemented or deficient rats.   However, 0~ decreased selenium-dependent
glutathione peroxidase activityin the-livers  of supplemented  rats  and-caused
an increase  in  deficient  rats.   In the  kidney,, both these groups of  animals
had an  increase  in  this enzyme activity.  Other groups of rats (deficient in
vitamin E,  supplemented with  selenium; supplemented  with vitamin  E, deficient
in selenium) were examined also and in  some cases,  different results were
observed.   The authors interpreted these results (along with pulmonary effects)
as a compensatory mechanism to protect cells from oxidants.  A more extensive
interpretation of the effects depends on the  nutritional  status and the presence
                                   9-196

-------
of other compounds metabolized by the affected enzymes.   For example, epoxide
hydrolase,  which was decreased in the vitamin E-  and  selenium-deficient  rats,
metabolizes reactive epoxides to dihydrodiols.  The metabolism  of  a  substance
such as benzo(a)pyrene  would be  expected to be affected by such  a  change.
However, because of the complexity of the metabolism of a given chemical, such
as benzo(a)pyrene,  a precise interpretation is not  possible at this time.
     Veninga et al.  (1981)  exposed mice to  03 and evaluated hepatic reduced
ascorbic acid content.   The authors expressed the exposure regimen in the form
of a C  x  T value from  about 0.2  to 3.2.  Actual  exposure regimes cannot be
                                                                 3
determined.  They stated  that the maximal 03 level was 1600 ug/m  (0.82 ppm)
and the maximal  exposure  time was 4 hr, which would have resulted in a C x T
of about 3.2.   Animals  were studied at  0,  30, and 120  min  postexposure.   It
appeared that immediately after exposure, a C x T value < 0.4 caused a decrease
in the  reduced  ascorbic acid content of  the liver.   At a C x T value of 0.4
and 0.8, there  appeared to be an  increase  that  was not  observed  at higher
values.  For the  30-min postexposure groups, the  increase in reduced ascorbic
acid shifted, with  the  greatest increase being at about  a C x  T value of 1.2
and no  change  occurring at a C x  T value of 2.0.  The 120-min postexposure
group was  roughly similar to the immediate  post-exposure group.   No effects
occurred 24 hr postexposure.
     Hepatic reduced ascorbic acid  levels  were  also studied   by  Calabrese
                                                      3
et al.  (1983c)  in rats exposed  for 3 hr to 588  ug/m   (0.3 ppm)  of 03  and
examined at 5 postexposure periods up to 24  hr.   Rats had significantly  increased
ascorbic acid  levels in both the 03 and air .groups,  with the  greater change
taking place in the  air group.  Thus, there were  no changes due to 03.   Likewise,
there was  no 0  effect  on reduced ascorbic acid content  in  the  serum.
9.4.6.2  The Endocrine  System.  A summary of the  effects of 03 on the  endo-
crine  system,  gastrointestinal tract,  and  urine  is  given in  Table 9-26.
Fairchild  and  co-workers  were the  first to observe the  involvement of  the
                                                                3
endocrine  system  in  0,,  toxicology.  Miceiexposed  to 11,368  ug/m (5.8 ppm) for
4  hr  were  protected against mortality  by a-naphthylthiourea (ANTU) (Fairchild
et  al.,  1959).   Because ANTU has  antithyroidal  activity and can alter adrenal
cortical  function,  Fairchild and  Graham (1963)  hypothesized a  possible  inter-
action  of  03 with the  pituitary-thyroid-adrenal  axis.   In exploring the hypo-
thesis,  they exposed mice and rats  for 3 to 4 hr to unspecified lethal  con-
centrations  of  03.   Thyroid-blocking agents and  thyroidectomy  increased the

                                    9-197

-------
                                      TABLE 9-26.  EFFECTS OF OZONE ON THE ENDOCRINE SYSTEH, GASTROINTESTINAL TRACT, AND URINE
Ozone
concentration Measurement
yg/i* ppi aethod
5.4
21
no
490
2940
980
1960
0.003 c
0.01
O.OSi
0.25 d
1.5
0.5 I
1
Exposure
duration and
protocpl
93 days,
continuous
2 hr
30 nin
5 hr/day,
4 days
Observed effect(s) Species
From 6th wk to end of exposure, 0.056 pp« Rat
increased the urine concentration
of 17-ketosteroids. After 93 days of
exposure to 0.056 ppn, the ascorbic and
level of the adrenal glands was decreased.
No data were presented for these effects.
1.5 pp« of 03 (30 Bin) inhibited gastric Rat
mortality; recovery was rapid. The
lower level caused no effects.
No effects on thyroid release of 131I, Rat
96-384 hr post 131I injection.
Reference
Eglite, 1968
Roth and Tansy, 1972
Fairchild et al., 1964
           1470
H
vo
CO
0.7S
m
4-8 hr
LM and TEM changes in parathyroid glands.
Loss of "clusterlike" arrangement of
parenchyma.  Dilated capillaries.
Vacuolated chief cells.  Increased RER,
prominent Golgi, abundant secretory
granules.
                                                                                                                  Rabbit
Atwal and Wilson, 1974
1470 0.75 NO
1470 0. 75 ND
1470 0.75 ND
1568- 0.8- NBKI
2940 1.5
(range) (range)
48 hr
postexposure
1-20 days
48 hr
postexposure
12-18 days
48 hr,
3, 10-13, 18
days post-
exposure
6 hr/day,
4 days/wk
about 19 wk
Early postexposure parathyroid glands * Rabbit
enlarged and congested with focal
vasculitis. After 7 days postexposure,
parenchyma! atrophy, leukocyte infiltra-
tion and capillary proliferation. Authors
suggest lesions may be due to autoimmune
reactions.
Mi crovascular changes in the parathyroid Dog
f glands, including hemorrhage, endothelial
"proliferation, platelet aggregation, and
lynphocyte infiltration.
Ciliated cysts found in parathyroid gland. Dog
Hallory body-like inclusions found in chief
cells of parathyroid with highest incidence
being 10-13 days postexposure.
Lower titratable acidity of urine, Rat
with no changes in levels of creatine,
uric acid/creatinine, ami no acid
nitrogen/creatinine, or excretion of
12 ami no acids.
Atwal et al., 1975
Atwal and Pens ingh, 1981
Perns ingh and Atwal , 1983
Atwal and Pensingh, 1984
Hathaway and Terrill, 1962

-------
                               TABLE 9-26.   EFFECTS OF 010NE ON THE ENDOCRINE SYSTEM, GASTROINTESTINAL TRACT, AND URINE  (continued)
Ozone
concentration
Hg/na
1960
3920
7840
1960
ppm
1
2
4
1
Measurement8
method
I
ND
Exposure
duration and
protocol
5 hr •
24 hr
:• Observed effect(s)b
Decreased release of 131I from thyroid,
48-384 hr post mlnjection to all 03
levels above 1 ppm. -
Decreased serum level of thyroid-
stimulating hormone from anterior
Species
Rat
Rat
Reference
Fairchild et al. ,
demons and
Garcia, 1980a,b.

1964

f
                                 pituitary, thyroid hormones (Ta,  T4,
                                 and free T4), and protein-bound iodine;
                                 no change in unsaturated binding  capacity
                                 of thyroid-binding globulin in serum;  in-
                                 crease in prolactin levels; no change  in
                                 levels of corticotropin, growth hormone,
                                 luteinizing hormone,  follicle stimulating
                                 hormone from pituitary or insulin.   Thy-
                                 roidectomy prevented the effect on  TSH
                                 levels.   There was no effect 'on the
                                 circulating half-life of 131I-TSH,   The
                                 anterior pituitaries had fewer cells,  but
                                 more TSH/cell.   The thyroid gland was  also
                                 altered.   Exposures to between 0,2  to  2 ppm
                                 for unspecified lengths of tine up  to  a
                                 potential maximum of 500 hr also  caused
                                 a decrease in TSH levels.
          1960     1
   ND
24 hr
          9800
CHEM
 3 hr
                                                               Decreased serum levels of T3, T4,  and  TSH.
                                                               In thyroidectomized and" hypophy sectoral'zed
                                                               rats,  the decrease in T4 was greater when
                                                               rats were supplemented with T4 in  the
                                                               drinking  water.
                                                                                   Rat
                                                                demons and Wei, 1984
79,800 > 5.0 I
> 3 hr
< 8 hr
Anti-thyroid agents, thyroidectony,
hypophysectomy, and adrenal ectomy
protected against 03- induced mor-
tality. Injection of thyroid hormones
decreased survival after Q3 exposure.
Mouse,
rat
Fairchild et al. ,
1959; Fairchild and
Graham, 1963;
Fairchild, 1963.
Increased levels  of 5-hydroxytryptamine
in lung;  decreases  in brain;  no change
in kidney.
                                                                                                                 Rat
Suzuki, 1976.
           Measurement method:   CHEM = gas  phase chemiluminescence;  NBKI = neutral buffered potassium  iodide;  I =  iodometric
           (Byers and Saltzman,  1959); ND = not described.

           Abbreviations used:   LM = light  microscopy;  TEM = transmission  electron microscopy; RER = rough  endoplasmic  reticulum;
           T3 = triiodothyronine;  T4 = thyroxine.

           Spectrophotometric technique (dihydroacridine).
           Flow rates from ozonator.

-------
survival of mice and rats acutely exposed to 0~, and injections of the thyroid
hormones, thyroxine  (T*),  or triiodothyronine (I,) decreased their survival.
This response  in  animals with altered thyroid function was not specific to a
concomitant  altered  metabolic rate,  because another drug  (dinitrophenol),
which increases metabolism, had no effect on the CL response.
     Hypophysectomy  and  adrenalectomy also  protected against Ov-induced mor-
tality, presumably in rodents (Fairchild, 1963).  Hypophysectomy would prevent
the release of thyroid-stimulating hormone,  thereby causing a hypothyroid con-
dition as well as preventing the release of adrenocorticotropic hormone (ACTH),
which would  cause a  decreased stimulation of the  adrenal  cortex to release
hormones.  This confirms the above-mentioned finding of thyroid involvement in
0,  toxicity  and  suggests that a decrease in adrenocorticosteroids reduces 0.,
toxicity.  Rats  that have been adrenalectomized and treated with  adrenergic
blocking agents are more resistant to 03 than rats that have only been adrena-
lectomized,  indicating  that  decreases in catecholamines reduce  CL toxicity.
     Potential tolerance to  the effect of 03  on thyroid activity was  also
investigated by  Fairchild et al.  (1964).  -A  variety  of  exposure  regimens were
used for  the rats,  and the  release  of    I  was used as an index of thyroid
                                                      •3
function.  A 5-hr exposure to 1960, 3920, or 7840 pg/m  (1, 2, or 4 ppm) of 0~
                          131                                            131
inhibited the  release  of    I at several time periods post injection of    I
(48, 96,  192,  and 384 hr).  The rats were injected before the 5-hr exposure,
presumably shortly before exposure.   Twenty-four hr post-injection, only the
highest  0,  concentration  showed  an effect.   Rats were also  exposed for 5
hr/day  for 4 days to either  980  or  1960 ug/m3  (0.5 or 1.0  ppm) of  Ov   At  96,
                                    131
192, and 384  hr  post-injection  of    I, no  effects were  observed.   Thus,
tolerance appeared  to  have  occurred, a  finding consistent with lethality
studies with 0-  (Matzen,  1957a).   In another study,  rats were exposed  to 3920
    3
(jg/m   (2.0 ppm)  of 0-  5 hr/day for  2 days and challenged with a  5-hr exposure
             3
to  7840  yg/m  (4.0  ppm) on the third day.  These animals exhibited a greater
effect  than  rats  that  received only the 7840-ug/m  (4.0 ppm) challenge.  This
difference persisted for 48, 96, and 192 hr post-injection.  Thus, although it
appears that tolerance occurs, it results in a condition that leads to stimula-
tion of thyroid  activity  after  a subsequent exposure  to  an 03 challenge.
     demons and  Garcia (1980a,b) extended this area of research by investi-
gating  the effects  of 0_ on the  hypothalamo-pituitary-thyroid axis  of rats.
Generally, these  three endocrine organs regulate the function of  each other
                                   9-200

-------
through complex feedback  mechanisms.  Either stimulation or inhibition of the
hypothalamus  regulates  the release of  thyrotropin-releasing  hormone (TRH).
The thyroid  hormones  (1~3 and T.) can  stimulate TRH,   Thyrotropin-releasing
hormone and  circulating  thyroid hormones (T.3 and T.)  regulate secretion of
thyroid-stimulating hormone (TSH) from the anterior pituitary.  Stimulation of
                                                                      3
the thyroid  by  TSH releases T~ and T,.   A  24-hr exposure  to  1960 ug/m   (1.0
ppm) of CL  caused decreases in the serum concentrations of TSH, T3, T», free
T., and protein-bound  iodine.   There was no  change in the uptake of T3, and
thus no change in the unsaturated binding capacity of thyroid-binding globulin
in the  serum.   Prolactin levels were increased also,  but  no alterations were
observed in the concentrations of other hormones (corticotropin, growth hormone,
luteinizing  hormone, follicle-stimulating hormone,  and insulin).  Plasma TSH
                                                                            3
was also evaluated after continuous exposures  to between  392 and 3920  ug/m
(0.2 and 2.0 ppm) for  unspecified  lengths of  time in  a fashion to result in a
concentration x time  relationship  between about 2 and 100.  Plasma TSH was
decreased after a C x T exposure of about 6.  These data cannot be independently
interpreted,  because  the  specific  exposure regimens  were not given.   The
authors state,  without any supporting data,  that the decrease in TSH  levels
persisted  "beyond two weeks"  when exposure  was  continued.  Therefore, it
appears that tolerance may not have occurred.  Thyroidectomized rats exposed
to 1960, 3920,  5880 or 7840 ug/m3  (1.0,  2.0,  3.0, or  4.0 ppm) of DS  for 24 hr
did not exhibit a decrease in  the  levels of TSH.   Exposure  to (presumably)
         3
1960 ug/m   (1.0 ppm)  for  24 hr did not alter  the circulating half-life of
125
   I-labeled  TSH  injected  into  the rats,  and therefore, there  apparently is no
effect  on TSH once  it  is  released from the pituitary.
                                                                            3
     To evaluate  pituitary function further,  rats exposed  24  hr to  1960 ug/m
(1.0 ppm) of  ozone  were immediately subjected to a 45-min  exposure to the  cold
(5°C)  (demons  and Garcia,  1980a,b).   The  anterior  pituitary released an
increased  level^pf TSH,  indicating., that  the  hypothalamus  was still able  to
respond (via increased TRH) after 0,  exposure.   The  increase was greater  in
the 0-  group, which might  indicate  increased production of TRH or  increased
sensitivity  to TRH.  In addition, the anterior  pituitaries of the 0, group had
fewer  cells  than  the air  group.  The cells from the  0,,-exposed rats had more
TSH and prolactin per  1000 cells, irrespective  of whether  the cells  had  received
a  TRH  treatment.   The cells from the 0-  group  also released  a greater  amount
of TSH, but  not prolactin,  into the tissue  culture medium.
                                    9-201

-------
     The thyroid gland  itself was altered by the 00 exposure (apparently 1960
    3
     , 1.0 ppm,  for 24 hr).  Ozone increased thyroid weight without changing
protein content (e.g., edema) and decreased the release of T. per milligram of
tissue.  There was  no change in T. release per  gland,   demons and Garcia
(1980a,b) interpreted these findings  as an 0~-induced lowering of the hypo-
thalamic set point for the pituitary-thyroid axis and a simultaneous reduction
of prolactin-inhibiting-factor activity in the hypothalamus,
                                                 3
     The effect  of  a 24-hr exposure to 1960 ug/m  (1 ppm) 03 on exogenous T4
levels was examined in rats by demons and Wei (1984) to increase understanding
of their previous  studies.   Normal, thyroidectomized, and hypophysectomized
rats were exposed  to Qg and serum  T.  levels  were measured.   Ozone reduced
serum T^ levels  in normal  rats, but  not  the  other groups of rats which  had
reduced levels of  1. prior to exposure.  However, when the thyroidectomized
and hypophysectomized rats  received supplemental  T. in their drinking water
that  increased  their serum  T.  level,  03  caused a decrease in  serum  T*.
Generally,  the higher the  pre-exposure T* levels, the greater the 03~induced
reduction in 1,  levels.   Similar observations were made for thyroidectomized
rats when serum T3 levels were measured.  Exposure to CU also decreased plasma
TSH levels in normal rats and in thyroidectomized rats supplemented with  T* in
the drinking water.   These observations led the authors  to hypothesize that
the results are  not due to a reduction in the hypothalamus-pituitary-thyroid
axis as previously suggested (demons and Garcia, 1980a,b), but that 03 causes
decreases in serum  T,,  T*,  and TSH  by peripheral changess possibly changes in
serum binding.
     The susceptibility  of  the  parathyroid gland to 0~ exposure  was  inves-
tigated by Atwal  and co-workers.   In  the  initial  study (Atwal  and Wilson,
                                        3
1974), rabbits were exposed to 1470 ug/m  (0.75 ppm) of 03 for 4 to 8 hr, and
the parathyroid  gland was  examined with light and electron microscopy at 6,
18, 22, and 66  hr after exposure.   The parathyroid gland exhibited increased
activity after  03  exposure.  Changes  included  hyperplasia of chief cells;
hypertrophy and  proliferation of the rough endoplasmic reticulum, free ribo-
somes, mitochondria,  Gqlgi  complex,  and  lipid bodies; and  an  increase  of
secretion granules  within  the vascular endothelium and capillary  lumen.  Such
changes suggested  an increased  synthesis  and  release of parathormone, but
actual hormone levels were not measured.
                                   9-202

-------
     Atwal et al.  (1975) also  investigated possible autoimmune involvement in
                                                                  3
parathyroiditis of rabbits following a 48-hr exposure to 1470 (jg/m  (0.75 ppm)
of Oq-  The  authors  stated that there was both a "continuous" and an "inter-
mittent" 48-hr exposure, without  specifying which results were due to which
exposure regimen.  Animals  were examined between 1 and 20 days postexposure.
Hyperplastic parathyroiditis was  observed to be followed by capillary proli-
feration and leukocytic infiltration.  Cytologic changes included the presence
of eosinophilic  leukocytes,  reticuloendothelial  and  lymphocytic infiltration,
disaggregation of the parenchyma, and interstitial edema.  A variety of alter-
ations were  observed  by electron microscopy,  including  atrophy of the endo-
plasmic reticulum of the chief cells, atrophy of mitochondria, degeneration of
nuclei, and proliferation of the venous limb of the capillary bed.  The alter-
ations to  the parathyroid  gland were progressive during  postexposure periods.
Parathyroid-specific autoantibodies  were  detected in the serum of 0.,-exposed
rabbits,  suggesting  that  the  parathyroiditis  might  be due to inflammatory
injury with an autoimmune causation.  Microvascular changes in the parathyroid
were further studied by Atwal and Pemsingh (1981) in dogs exposed to 1470 ug/m3
(0.75 ppm) of 03 for 48 hr.  They reported focal hemorrhages, vascular endothe-
lial proliferation, intravascular platelet aggregation, and lymphocytic infiltra-
tion.  A  potential autoimmunity after 0,  exposure was  also observed  by Scheel
et al.  (1959), who showed  the  presence of circulating  antibodies  against  lung
tissue.
     In  another  study, Pemsingh and  Atwal  (1983)  studied cells   (APUD-type)
within ciliated  cysts  of the parathyroid  gland  of dogs  exposed for  48 hr to
         3
1470 ug/m  (0.75 ppm)  0~.   Examination was made  at several times  post-exposure
(3,  10-13,  and  28 days).   Ozone caused ciliated cysts in 5 out of 12 exposed
dogs  and  1 out of 4  control, dogs.   Cysts  in  the 03-exposed dogs had  different
cellular  content,  namely the presence of  APUD-type  cells  on the cyst wall.
The  interpretation of  these findings is  not- clear.   The ultrastructure  .of
chief  cells,  which are the hormone-producing cells,  was  also  examined  in  dogs
exposed  identically  to  those  described  above  (Atwal  and Pemsingh,  1984).
Mallory body-like  inclusions (i.e.,  accumulations of filaments in the perinuclear
area) were found within the chief cells of the parathyroid of 0.,-exposed  dogs.
Additional  alterations were observed.  The  highest  incidence  of  this  change
was  at 10  to 13 days post-exposure.
                                    9-203

-------
     Also classed  as a  hormone is 5-hydroxytryptamine  (5-HT),  and  it too
interacts with 0-  toxicity (Suzuki, 1976).  It  has a variety of activities,
including bronchoconstriction and  increased capillary permeability;  it can be
a neurotransmitter.  Although rats  were exposed  to a high concentration  (9800
    3
|jg/m , 5,0 ppm) of 0_ for 3 hr,  this study is discussed because extrapulmonary
effects were observed.   This  exposure caused an increase in the 5-HT content
of the  lung and spleen,  a  decrease  in  5-HT in the brain, and no  change in the
levels of 5-HT in the liver or kidney.
     Adrenal  cortex  function  after a 93-day (continuous) exposure to 0, was
                                                           3
investigated in rats (Egl.ite,  1968).   Exposure to 110 ug/m  (0.056 ppm), but
not lower levels,  increased the urine concentration of  17-ketosteroids  from
the 6th week of exposure to the end of exposure.   There was also a decrease in
ascorbic acid  in the adrenals.  The generation and monitoring methods for the
0, exposures  were not  sufficiently described.   The  authors  described the
effects as  statistically  significant  but did not specify  the statistical
methods.  No data  were  provided.   Therefore, these results  need to be con-
firmed before accurate interpretation is possible.
9.4.6.3  Other Effects.   Rats were exposed for  6  hr/day,  4 days/week,  for
about 19 weeks,  and  analyses were performed on urine collected for the 16 hr
following the  exposure  week (Hathaway and Terrill, 1962).   Ozone exposures
were uncontrolled, ranging from 1568 to 2940 (0.8 to 1.5 ppm).   All parameters
were not measured for each week of'exposure.   On days 91 and 112 after initial
exposure, there was  a lower titratable acidity and higher pH in the urine of
Qg-exposed animals.  Titratable acidity was  also lower on day 98.  Ozone did
not alter the  levels of creatinine, creatine, uric acid/  creatinine,  ami no
acid  nitrogen/creatinine excretions,  or  excretion  of 12 ami no acids.   The
lungs and kidneys  were  examined histologically,  and no consistent  differences
were observed.  The  authors interpreted the results as a reflection of respira-
tory alkalosis, assuming no kidney• toxic-i-ty.
     Gastric secreto-motor activities of the rat were investigated by Roth and
                                           3
Tansy (1972).  A  2-hr exposure to  490  ug/m   (0.25  ppm)  caused  no effects.
                                       •a
Thirty minutes of exposure to 2940 ug/m  (1.5 ppm)'inhibited gastric motility,
but  activity  tended to  return  towards normal  for the remaining 90  min of
exposure.  Recovery  had occurred  by 20  min  postexposure.   However, these
results are questionable because ozone was monitored only  by ozonator flow
rates.
                                   9-204

-------
9.5  EFFECTS OF OTHER PHOTOCHEMICAL OXIDANTS
9.5.1  Pe rpxyacety1 N i1rate
     Very little information on the toxicity of peroxyacetyl nitrate (PAN) has
appeared in the literature since the previous criteria document on photochemi-
cal oxidants  (U.S.  Environmental  Protection Agency, 1978).   The document
reviewed the  results  of  inhalation experiments with mice  that tested PAN's
lethal  concentration  (LCgp) (Campbell  et al. , 1967);  its  effects  on lung
structure (Dungworth  et  al. ,  1969); and  its influence  on  susceptibility to
pulmonary bacterial  infections  (Thomas  et al., 1979).  The concentration of
                                                3                      3
PAN used in  these studies ranged from 22.3 mg/m  (4.5 ppm) to 750 mg/m  (150
                                                        3
ppm).   They  are  considerably  higher than the  0.232  mg/m   (0.047 ppm) daily
maximum concentration  of PAN reported in  recent years for  ambient air samples
in areas having relatively high oxidant levels (Chapter 5) and are of question-
able relevance to the assessment of effects on human health.
     Campbell et  al.  (1967) estimated that the LC5Q for mice ranged from 500
to 750 mg/m3 (100 to 150 ppm) for a 2-hr exposure to PAN at 80°F  (25°C).  Mice
in the 60- to 70-day-old age group were more susceptible to PAN lethality than
mice  ranging  from 98 to 115 days  in  age.   Temperature  also influenced  the
lethal  toxicity  of PAN;  a higher LC5Q (less susceptibility) was  seen at 70°F
(125 ppm) than at 90°F (85 ppm).  In a follow-up study,  Campbell  et al.  (1970)
characterized the  behavioral  effects  of PAN by determining the depression of
voluntary wheel-running  activity in mice.1   Exposures  to  13.9, 18.3,  27.2,
31.7, and 42.5  mg/m3 (2.8, 3.7, 5.5, 6.4,  and 8.6 ppm) of PAN for 6 hr de-
pressed both the 6-hr and 24-hr activity when compared to similar pre-exposure
periods.  Depression  was more  complete and occurred more rapidly with higher
exposure levels.   However,  the  authors  indicated  that PAN  was  less  toxic than
0,, when  compared to similar behavioral data reported by Murphy et al. (1964)
(Section 9.4.1).
                                                                             3
     Dungworth et al,  (1969) reported that daily exposures  of mice to 75 mg/m
(15 ppm) of PAN 6 hr/day for 130 days caused a 30-percent weight  loss compared
to  sham controls,  18 percent mortality, and  pulmonary lesions.   The most
prevalent  lesions  were  chronic hyperplastic  bronchitis  and proliferative
peribronchiolitis.
     Thomas  et  al.  (1979) found that mice  exposed to 22.3 mg/m3  of PAN (4.5
ppm)  for 2 hr and  subsequently  challenged with a Streptococcus  sp. aerosol for
1  hr  showed a significant  increase  in mortality and a reduction in mean  survival
                                   9-205

-------
rate, compared  to  mice exposed to air.   No  effect  on the  incidence of  fatal
pulmonary  infection  or survival time  was observed  in mice challenged  with
                                                                •3
Streptococcus sp.  1  hr before  the  pollutant  exposure  (27.2 mg/m of PAN for '3
hr),  Thomas at al. (1981a) published additional data that extend observations
of  reduced resistance  of  mice to streptococcal pneumonia over a  range of
                                                                            o
exposures  to PAN.  A single 2- or 3-hr exposure to PAN at 14.8 to 28.4 mg/m
(3.0 to  5.7  ppm)  caused a significant increase in the susceptibility of mice
to  streptococcal pneumonia.   The mean excess mortality rate ranged from 8 to
                                              3
39  percent.  Mice  exposed  to 0~ at  0.98  mg/m  (0.5 ppm) and challenged with
the Streptoccus sp. aerosol resulted in a mean excess mortality (38 percent) that
was almost equivalent to  the excess mortality  for  the group'exposed  to 28.4
    o
mg/m  of  PAN.   The results  agreed  with earlier  reports that  PAN is less toxic
                                                                       2
than Og to mice exposed under ambient conditions.   Exposure to 7.4 mg/m   (1.5
ppm) of  PAN 3  hr/day,  5 days/week for 2 weeks had no  appreciable effect,
although  no  statistics were provided.   Neither exposure routine altered  the
morphology, viability, or phagocytic activity of isolated macrophages, although
there was  a  decrease in ATP levels.   The other noticeable effect was  that
macrophages  isolated  from  the  animals that were repeatedly exposed failed to
attach themselves  to  a glass  substrate.   These investigators  also studied
whether a  chronic  infection initiated with an exposure to Mycobacteriurn tubercu-
losis (RIRv) was  influenced by subsequent exposure to PAN.  The exposure to
this oxidant (25  mg/m  for i
growth in  the lungs of mice.
                      3
this oxidant  (25  mg/m  for 6 days) did  not alter the pattern of bacterial
9.5.2  Hydrogen Peroxide
     Hydrogen peroxide  (H«0?)  has been reported to occur in-trace amounts in
urban air  samples  (Chapter 5), but very little is known about the effects of
HpOn from  inhalation  exposure.   Most of the  early work  on  H202  toxicity  in-
volved exposure to very high concentrations.  Oberst et al.  Q954) investigated
the inhalation toxicity of 90 percent H909, vapor in rats, dogs, and rabbits at
                                    3
concentrations ranging  from  10 mg/m  (7 ppm) daily for six months to an 8-hr
                    o
exposure to 338 mg/m  (243 ppm).  After autopsy, all animals showed abnormali-
ties of the lung.  In a recent experiment, Last et al. (1982) exposed rats for
                                                                  3
7 days to  > 95-percent H^O- gas with a concentration of 0.71 mg/m  (0.5 ppm)
in  the  presence of respirable ammonium sulfate particles.   No  significant
effects were  observed in body weight,  lung  lobe weights, and  protein or  DNA
                                   9-206

-------
content of lung homogenates.   The authors suggested that because HpQp is highly
soluble, it  is  not expected  to penetrate to the deep lung, which may account
for the absence of observed effects.
     The majority  of studies  on H?0p explore  possible  mechanisms  for the
effects of HpOp.   These  include direct cellular effects  associated with i_n
vitro exposure  to  H000  and biochemical reactions  to  H000  generated in  vivo.
  :::::::::::::::::               £ £»                                £-  £            —  	
     Hydrogen peroxide may affect lung function by the alteration of pulmonary
surfactant.  Wilkins  and  Fettissoff (1981) found  that lo"2  to  lO'^N of H202
increased the surface tension of saline extracts of dog lung homogenates.  The
                                                 3
authors estimated  that  a  dog breathing 1.4 mg/m  (1 ppm) of H9Q~  for  30 hr
                                               -2
would build  up  a pollutant concentration  of 10  M  in  the  surfactant, assuming
that all the HpOp  was retained by the lungs.   However, as stated above, this
estimate of  tissue dose  is not  realistic, because most  of the  HpOp would  be
absorbed in the upper airway.
     Another mechanism by which H909 may affect ventilation  is by changing the
                                 ^ *•                                   -A
tone of airway  smooth muscle.   Stewart et al.  (1981) reported  that 10   M  of
HpQp caused  significant  constriction  of strips of subpleural  canine  lung
parenchyma and  of  bovine  trachealis muscle.   In the distal airway preparation
(canine),  this  contraction was  reversed  by catalase.   Pretreatment of both
proximal (bovine)  and distal muscle strips with meclofenamate or iodomethacin
markedly reduced the response  to HpOp.   This  suggested  that the increase  in
airway  smooth muscle tone produced by HpOp involves  prostaglandin-like sub-
stances.
     The potential  genotoxic effects from in  vitro HpOp exposure  have been
evaluated  in isolated cell systems.   Bradley  et al. (1979) reported that HpOp
produced both  toxicity  and single-strand DMA  breaks  but was not mutagenic at
concentrations  up  to 530  pM.   They observed  a significant  increase in the
frequency  of reciprocal  sister  chromatid  exchanges in V-79  Chinese hamster
cells  at a concentration  of  353 |jM of HpQp.   However, the authors  pointed  out
that  sister chromatid exchange  frequency was  not  necessarily equivalent  to
increased  mutant  frequency.   In subsequent experiments, Bradley and Erickson
(1981)  confirmed  these  observations in V-79  Chinese  hamster lung  cells and
were unable  to  detect any DNA-protein or DNA-DNA  crosslinks with  353  uM  of
HpOp for 3 hr at 37°C.  Increases  in the  frequency of sister chromatid  exchanges
in  Chinese hamster cells  have  also  been  found by  MacRae  and Stich (1979),
Spelt  and  Vogel (1982),  and  Speit et al.  (1982).  Wilmer and Natarajan (1981)
                                    9-207

-------
reported only a  slight enhancement in the  frequency of  sister chromatid ex-
changes in Chinese hamster ovary cells following treatment with up to 10  M of
HgOg.   In comparison, cells were killed with a concentration of 10  M of H2Q2.
Similarly, HpQp  (10  mg/ml) was  negative  in the  Ames mutagenicity assay
(Ishidate and Yoshikawa, 1980), and several other investigators have confirmed
the lack of mutagenicity  for H202  (Stich et a!,, 1978; Kawachi et a!.,  1980).
     Johnson et  al.  (1981)  reported that the intrapulmonary  instillation of
glucose oxidase,  a  generator of HpO,,,  increases lung permeability in rats.  A
greater increase  in lung  permeability was achieved by the addition of a com-
bination of glucose oxidase and lactoperoxidase than by the glucose oxidase-
HoCU-generating  system alone.   Horseradish peroxidase did not effectively
substitute for  lactoperoxidase in  the potentiation of damage.  Injury was
blocked by catalase but not by superoxide  dismutase  (SOD),  suggesting that
HjjQ,  or  its  metabolites,  rather  than  superoxide,  were  involved.   Because
horseradish peroxidase did not potentiate  the  glucose oxidase damage, the
authors speculated that the mechanism of injury occurs through the action of a
hall de-dependent  pathway  described for cell injury produced by H202 and lac-
toperoxidase/myeloperoxidase (MPO)  (Klebanoff and Clark,  1975).  Any source of
H^O*  plus MPO  and a halide cofactor  is capable of  catalyzing many  oxidation
and halogenation reactions  (Clark  and Klebanoff, 1975),  but  other possible
mechanisms of oxygen radical production have been proposed  (Halliwell,  1982).
     Carp and Janoff (1980) have  shown that a H202  generating system with MPO
and Cl  will  suppress the elastase  inhibitory capacity of the protease inhibi-
tor (BMPL) present  in bronchial mucus.  This antiprotease is capable of inhibit-
ing the  potentially dangerous  proteases  found in  human polymorphonuclear
leukocytes (PMNs),  including  elastase and cathepsin-G.  Inactivation of BMPL
could make the  respiratory  mucosa  more susceptible to attack by inflammatory
cell  proteases.   Human PMN  have not been shown  to contain MPO but macrophages
may contain analogous forms of peroxidase.
      In other i_n  vitro experiments, Suttorp and Simon (1982) demonstrated that
H202  generated  by glucose oxidase was cytotoxic to cultured  lung epithelial
cells  (L9  cells) in a concentration-dependent fashion.   Cytotoxicity was
                          51
measured  by  determining    Cr  release from target cells.  Cytotoxicity was
prevented by  the addition  of  catalase.   It was stressed that there  is no
established identity between  the  L? cell line and the jji situ type 2 pneumo-
cytes from which  they were derived.
                                   9-208

-------
9.5.3  Formic Acid
     lexicological interest  in  formic  acid has focused primarily on its role
as the metabolic  end product of methanol  and  on  any similarity the effects
from inhalation exposure to formic acid vapor may have with methanol toxicity.
Unfortunately, the concentrations  (20  ppm) used in the two studies discussed
below are over three orders of magnitude greater than the trace concentrations
(<0,015 ppm) reported in the highest oxidant areas of Southern California (see
Chapter 5).
     Zitting and  Savolainen  (1980) exposed rats to  20 ppm  (0.8  |jmol/L)  formic
acid vapor,  6 hr/day for  3  and 8 days.   Tissue  samples were analyzed for
neurochemical effects as  well  as effects on drug-metabolizing enzymes in the
liver and  kidneys.   Enzyme profiles  were variable, depending on the sampling
period during  exposure.   However, a general  pattern  developed.   Cerebral
glutathione  concentration  and  acid proteinase activity increased after 3 and
8 days  of exposure,  respectively,  indicating possible  lipid peroxidation
associated with  cerebral  hypoxia.  Decreased kidney ethoxycoumarin  deethylase
activity,  cytochrome P-450 content,  and glutathione concentrations were also
consistent with  changes due to lipid peroxidation.  The liver, which is less
sensitive to tissue  hypoxia, showed only small increases in deethylase activity
although associated  with a reduction in glutathione.  Since prolonged low-level
cerebral  hypoxia  can potentially lead to  demyelination  and  subsequent  nerve
degeneration, the authors  repeated the study to look specifically at effects
on glial  cells  (Savolainen and Zitting, 1980).   Rats were exposed  to 20  ppm
formic acid  vapor,  6 hr/day, 5 days/week  for  2  or 3 weeks.   Again, enzyme
profiles  were  indicative  of metabolic responses to tissue hypoxia, providing
evidence  for potential  central  nervous system toxicity  at high formic acid
vapor concentrations.

9.5.4  Complex Pollutant Mixtures                        •    ' '    •
     Additional  toxicological  studies have  been  conducted on the potential
action of complex mixtures of oxidants and other pollutants.  Animals have
been  exposed under  laboratory  conditions  to ambient air  from  high  oxidant
areas,  to UV-irradiated  and nonirradiated reaction mixtures of  automobile
exhaust and  air,  and to other combinations  of  interactive  pollutants.  Although
these mixtures  attempt  to simulate the photochemical  reactions  produced under
actual atmospheric conditions, they  are extremely difficult to  analyze because
                                    9-209

-------
of their  chemical  complexity.   Variable  concentrations  of total oxidants,
carbon monoxide,  hydrocarbons,   nitrogen  oxides,  sulfur oxides,  and  other
unidentified complex  pollutants have been  reported.   For  this reason, the
studies presented in this section differ from the more simplified combinations
of 0~ and  one  or two other nonreactive pollutants  discussed  under previous
sections of  the  chapter.   The effects described  in animals exposed to UV-
irradiated exhaust mixtures are not necessarily uniquely characteristic of CU,
but most of  them could have been produced by 0,,.  In most cases,  however, the
biological  effects  presented  would be difficult  to associate with any one
pollutant.
     Research on ambient  air  and UV-irradiated or nonirradiated exhaust mix-
tures is summarized  in  Table  9-27.  Long-term exposure of various species of
animals to ambient  California atmospheres have produced changes  in the pul-
monary function of  guinea pigs   (Swann and Balchum,  1966; Wayne and Chambers,
1968) and  have produced a number of biochemical,  pathological, and behavioral
effects in mice, rats, and rabbits (Wayne and  Chambers,  1968;  Emik and Plata,
1969; Emik et  al.,  1971).  Exposure to UV-irradiated automobile exhaust con-
taining oxidant  levels  of 0.2,to  1.0  ppm produced histopathologic changes
(Nakajima et al.,  1972)  and  increased susceptibility to infection (Hueter et
al., 1966) in  mice.   Both UV-irradiated and nonirradiated  mixtures produced
decreased spontaneous running activity (Hueter et al., 1966; Boche and Quilligan,
1960) and  decreased infant survival rate and  fertility (Kotin and Thomas,
1957; Hueter et  al.,  1966; Lewis,  et al.,  1967)  in a  number  of experimental
animals.  Pulmonary changes were demonstrated  in  guinea  pigs  after short-term
exposure to  irradiated automobile  exhaust (Murphy et  al., 1963; Murphy, 1964)
and  in  dogs  after long-term  exposure  to  both  irradiated and  nonirradiated
automobile exhaust  (Lewis et  al.,  1974; Orthoefer et al., 1976).   Irradiation
of the  air-exhaust  mixtures  led to the formation of  photochemical reaction
products that were  biologically more active than  nonirradiated mixtures.  The
concentration of total  oxidant  as  expressed by CU  ranged  from 588 to 1568
    3
|jg/m  (0.30  to 0.80 ppm)  in the irradiated  exhaust  mixtures,  compared to  only
a trace or no oxidant detected in the nonirradiated mixtures.
     The description  of  effects following exposure of dogs for 68 months to
automobile exhaust, simulated smog, oxides  of  nitrogen,  oxides of sulfur, and
their combination  has been expanded in a monograph by Stara  et al.  (1980).
The  dogs  were examined after 18 months (Vaughan et  al.,  1969),  36 months
                                   9-210

-------
                                                         TABLE 9-27.   EFFECTS OF  COMPLEX  POLLUTANT MIXTURES
ro
H

Concentration,
(ppra)


Pollutant
Exposure
duration and
protocol


Observed effect(s)c


Species


Reference
A. Ambient air
0.032 - 0.050
9.1 - 13.5
0.044 - 0.077
0.019 - 0.144



0.057
1.7
2.4
0.019
0.015
0.004


0.062 - 0.239
0.03 - 0.07
2.7 - 4.4
0.27 - 0.31
13 - 38
2-9
0.09 - 0.70


0.4 (max)

°x
cfi
N02
NO



0,,
c8
HC
N02
NO
PAN


ox
NU2
HC
°x
c8
HC
NOX


°x

Lifetime
study,
continuous




2.5 years,
continuous






13 months,
continuous

1 year,
continuous



19 weeks,
continuous
No clear chronic effects; "suggestive"
changes in pulmonary function, morphology,
and incidence of pulmonary adenomas in
aged animals. Increased 17-ketosteroid
excretion in guinea pigs. Decreased
glutamic oxalacetic transaminase in blood
serum of rabbits.
Reduced pulmonary alkaline phosphatase
(rats); reduced serum glutamic oxaloacetic
transaminase (rabbits); increased pneu-
monitis (mice); increased mortality (male
mice); reduced body weights (mice); de-
creased running activity (male mice); no
significant induction of lung adenomas
(mice).
Decreased spontaneous running activity.


Expiratory flow resistance increased on
days when 0 reached > 0.30 ppn or at com-
bined concentrations of > 40 ppm of CO,
16 ppm of HC, and 1.2 ppm of NO . Indi-
dividual sensitivity demonstrated. Tem-
perature was an important variable.
No consistent effects on conception rate,
litter rate, or newborn survival.
Mouse,
rat,
hamster,
guinea pig,
rabbit


Mouse,
rat,
rabbit





Mouse


Guinea pig




Mouse

Wayne and Chambers,
1968





Emik et a!., 1971







Enrik and Plata, 1969.


Swann and Balchum, 1966




Kotin and Thomas, 1957


-------
                                                     TABLE 9-27,  EFFECTS OF COMPLEX POLLUTANT MIXTURES  (continued)
V
NJ
Concentration,3 .
(pom) Pollutant0
Exposure
duration and
protocol
Observed effect(s)c
' Species
Reference
B. Automobile exhaust
0.012-
3.0
0.04 -
0.06 -
0.06 -
0.04 -
0.1S -
0.4 -
6
20 -
0.1 -
0.2 -
5
40 -
0.2 -
100
24 -
0.1 -
0.1 -
0.42 -
0.02 -
0.65
1.10
5.00
1.20
0.2
0.5
l.B
36
100
0.5
0.6
8
60
0.4
30
1.0
2.0
0.49
0.03
03 (max)
HC (propylene)
N02
NO
S02
03
N02
NO
HC (CH4)
CO
N0*
HC*
CO
Os
CO
HC (CH4)
N02
NO
S02
H2S04
0.5-6 hr
(diesel)
1.5-23 10
4 weeks,
5 days/week,
2-3 hr/day
18-68 months,
7 days/week,
16 hr/day
2-3 years
recovery In
ambient air.
UV-irradiatlon of propylene, S02, NO, and
N02 produced 03 and a mutagenic moiety
when collected particles were tested by
the pi ate- Incorporation test. Irradiation
did not alter and 03 tended to reduce
the mutagenic response.
Increased pulmonary infection. Decreased
fertility and infant survival. No signi-
ficant changes in pulmonary function. De-
creased spontaneous running activity during
the first few weeks of exposure.
Histopathologic changes resembling
tracheitis and bronchial pneumonia at
the higher concentration range of oxidants.
No cardiovascular effects
No significant differences In collagen:
protein ratios; prolyl hydroxylase in-
creased with high concentrations of all
mixes.
Salmonella
typhJBur1u«
House,
rat,
hamster,
guinea pig
Mouse
Dog
• Dog
Claxton and Barnes, 1981
Hueter et al . , 1966
Lewis et al., 1967
Nakajima et al . , 1972
Bloch et al., 1972, 1973
Gillespie, 1980
Orthoefer et al . , 1976
                                                                  Pulmonary function for groups receiving
                                                                  oxidants [irradiated exhaust (I) +  SOX]
                                                                  18 months:   no effects
                                                                  36 months:   no effects
Dog
Vaughan et al., 1969
Lewis et al., 1974

-------
                                                   TABLE 9-27.  EFFECTS OF COMPLEX POLLUTANT MIXTURES  (continued)
K>
M
U!
                Concentration,
                  (ppm)
Pollutant"
  Exposure
duration and
  protocol
Observed effect(s)
Species
Reference
                                                                61 months:  N2 washout increased (I);
                                                                R. increased (I, I+SO ).

                                                                2 years recovery:  P-.C02 increased (I+SQ );
                                                                Vp increased (I, I+SOX); Dico/TLC decreased

                                                                and V  increased in all groups; lung com-
                                                                partment volumes increased (I+SOX).

                                                                Morphology (32-36 months recovery):  air
                                                                space enlargement; nonciHated bronchiolar
                                                                hyperplasia; foci of ciliary loss with and
                                                                without squamous metaplasia in trachea
                                                                and bronchi.
                                                                            Dog
                                                                            Dog
                                                                            Dog
                                                                             Lewis et al., 1974


                                                                             Gillespie, 1980





                                                                             Hyde et al,,  1978
0.33 - 0.82
0.16 - 5.50
0.16 - 4.27
0.12 - 2.42
0.02 - 0.20


°x
NOg
NO
Formaldehyde
Acrolein


4-6 hr






Increased pulmonary flow resistance, in- Guinea pig Murphy et al., 1963;
creased tidal volume, decreased breathing Murphy, 1964
frequency due to formaldehyde and acrolein
at low 0 : aldehyde ratio. Decreased
tidal volume, increased frequency, in-
creased pulmonary resistance due to 0
and NO at high '0 : aldehyde ratio.
C. Other complex mixtures
0.08
0.7i
2.05
1.71
0.3
1.0
2.0

Os
S02
T-2 Butene
acetal dehyde
03
N02
S02

4 weeks
7 days/week
23 hr/day

2 weeks
7 days/week
23 hr/day

Alteration in distribution of ventilation Hamster Raub et al., 1983b
(AN2) and increased diffusing capacity.


Voluntary activity (wheel running) Mouse Stinson and Loosli, 1979
decreased 75% after 1-3 days, returning
to 853) of pre-exposure levels by the end
of 14 days.

-------
                                    TABLE 9-27.   EFFECTS  OF  COMPLEX  POLLUTANT MIXTURES   (continued)
Concentration,9
(ppm)
0.40 - 0.52
1.0 - 2.15
1.25
Pollutant6
03
Ox (§as vapor)
Ox (gas vapor)
Exposure
duration and
protocol Observed effect(s)
24 hr Decreased spontaneous wheel running
activity.
19 weeks, Decreased conception rate, litter rate,
continuous and newborn survival.
Species
Mouse
Mouse
Reference
Boche and
1960
Kotin and

Quilligan,
Thomas, 1957
.Ranked by nonspecific oxidant concentration (Os or 0 ).
 Abbreviations used:  0$ = ozone; 0  = oxidant; CO = carbon
 	 	  83 = ozone; 0  = oxidant; CO = carbon monoxide;  NO = nitrogen oxide,  K02 = nitrogen dioxide;
 NO  = nitrogen oxides; S02 = sulfur dioxide; HC = hydrocarbon;  CH4  =  methane;  H2S04 = sulfuric acid;  PAN - peroxyacetyl nitrate.
c
 See Glossary for the identification of pulmonary symbols.

-------
(Lewis et. al.,  1974),  48 to 61 months  (Bloch  et al. , 1972, 1973; Lewis et
al., 1974), and  68 months  (Orthoefer et  al., 1976) of exposure; the dogs were
examined again 24 months  (Gillespie,  1980) or  32 to 36 months  (Orthoefer et
al., 1976;  Hyde  et  al.,  1978)  after exposure  ceased.  Only those results
pertaining to oxidant exposure are described in this section, which limits the
discussion to groups exposed to irradiated automobile exhaust (I) and irradiated
exhaust supplemented with  sulfur oxides  (I+SO ).  See Table  9-27  for  exposure
                                             s\
concentrations.
     No specific cardiovascular effects  were reported during  the course of
exposures (Bloch et al., 1971, 1972, 1973) or 3 years after exposure (Gillespie,
1980).   Similarly, Orthoefer et al.-(1976) reported no significant biochemical
differences in the collagen to protein ratio in tissues of dogs exposed for 68
months or after  2.5  to 3 years of recovery in  ambient air.  However, prolyl
hydroxylase levels were reported to have increased in the lungs of dogs exposed
to  I and  I+SO ,  when compared to  control  air  and the nonirradiated exhaust
              J\
alone or in combination with SO .
                               /v
     No significant impairment of pulmonary function was found after 18 months
(Vaughan et al., 1969) or 36 months (Lewis et al., (1974) of exposure.  However,
by 61 months of  exposure, Lewis et al.  (1974) reported increases  in the nitrogen
washout of dogs  exposed to I, and higher total  expiratory resistance in dogs
exposed to both  I and I+SO , when compared to their respective controls receiv-
                          /\
ing  clean  air and  SO   alone.  Two years  after exposure  ceased,  pulmonary
                      /\
function was  remeasured by Gillespie (1980).  These  measurements  were made  in
a  different  laboratory than  the one  used during exposure,  but consistency
among measurements of the control group and another set of dogs of similar age
at  the  new  laboratory  indicated that this difference  did not have a major
impact on the findings.   Arterial  partial pressure  of  C0? increased in the
group exposed to I+SO   and total  deadspace  increased  in  the the  I  and  I+SO
                      )\                                                     />
groups.  The  diffusing capacity for  carbon monoxide  (D,   ) was  similar  in  all
exposure groups, but when normalized for total  lung capacity (TLC), the D|CO/TLC
ratio was  smaller in  exposed groups than in the air  control  group.   Mean
capillary blood  volumes also increased  in all  exposed groups.   No changes  in
lung  volumes  were reported  at the end  of exposure (Lewis  et  al., 1974).
However, when  lung  volumes from these animals  were  measured 2 years later,
increases were reported in the I+SO  group.  Unfortunately,  the sample size of
              1                    /\
dogs exposed  to  I was  too small  (n = 5)  to permit meaningful comparisons.   In
                                   9-215

-------
general, pulmonary  function  changes  were found to  be  similar in all groups
exposed to automobile  exhaust alone or supplemented with  SO  .   Exposure to
                                                            Pv
these mixtures with or without UV-irradiation produced lung alterations normal-
ly associated with injury to the airway and parenchyma.
     The  functional  abnormalities mentioned  above showed  relatively  good
correlation with  structural  changes  reported  by Hyde et al. (1978).  After  32
to 36 months of recovery in clean air, morphologic examination of the lungs by
light microscopy,  scanning electron  microscopy,  and transmission electron
microscopy revealed  a number  of exposure-related  effects.   The displaced
volume of the fixed  right  lung was  larger  in  the  I-exposed  group.  Both  the  I
and I+SQ  groups  showed  random enlargement of alveolar airspaces centered in
        f\
respiratory bronchioles  and  alveolar ducts.   Small hyperplastic lesions were
observed at the junction of the terminal bronchiole and the first-order respi-
ratory bronchiole.   Foci  of  ciliary loss associated with squamous metaplasia
were also observed in the intrapulmonary bronchi of the I+SO  group.   However,
                                                            A
because these was no significant difference in the magnitude of these lesions,
oxidant gases and SO  did not appear  to  act  in an additive  or  synergistic
                     X
manner.
     Additional  work  on  irradiated  and nonirradiated automobile exhaust has
been presented  by Claxton and Barnes  (1981).  The mutagenicity of diesel
exhaust particle extracts collected under smog-chamber conditions was evaluated
by the  Salmonella typhimurium pi ate-incorporation test (Ames et al.,  1975).
The authors demonstrated that the irradiation of propylene, S02, NO, and NQ"2
produced 0, and a mutagenic moiety.   In baseline studies on diesel  exhaust, in
which Qg was  neither added nor produced, the mutagenicity of each sample was
similar under dark or UV-light conditions.   When  0, was introduced  into the
smog chamber, the mutagenicity of the organic compounds was reduced.
     The behavioral effect of a nonirradiated reaction mixture was examined by
Stinson and  Loosli  (1979).   Voluntary wheel-running activity was recorded
                                                                           3
during a continuous  2-week exposure to synthetic  smog  containing  588 pg/m
(0.3 ppm) of 0,, 1 ppm of NOp, and 2 ppm of S0?, or to each component separate-
ly.  An immediate decrease in  spontaneous  activity occurred after  1  to 3 days
of exposure,  returning to 85 percent  of the  original  activity by  the  end of
exposure.   Activity  returned to  basal  levels 5 days after breathing filtered
air.  Ozone alone produced a response that was similar to that of the synthetic
smog mixture.   Since N02 and SOp alone had only moderate effects,  the  authors
concluded that 0~ had the major  influence on  depression of activity.
                                   9-216

-------
     More 'recently, Raub et al.  (1983b) reported pulmonary function changes in
hamsters exposed 23 hr/day for 4 weeks to a nonirradiated reaction mixture of
trans-2-butene, CL,  and SCL.   Decreases in the nitrogen washout slope and
                 o         £
increases in  the  diffusing capacity  indicated a  significant compensatory
change in distribution of ventilation in the lungs of exposed animals.  Animals
compromised by  the presence  of  elastase-induced emphysema  were unable to
respond to this pulmonary insult in the same manner as animals without impaired
lung function.
9.6  SUMMARY
9.6.1  Introduction
     The biological effects of 0., have been studied extensively in animals and
a wide array  of toxic effects have been ascribed to CL inhalation.  Although
much has been accomplished  to  improve the  existing  data base,  refine the  con-
centration-response relationships  and  interpret better the mechanisms  of CL
effects,  many of the  present data were  not  accumulated  with the  idea  that
quantitative  comparisons  to man  would be drawn.   In many  cases, only qualita-
tive comparisons can be made.  To maximize the extent that animal  toxicological
data can be  used to estimate the  human  health ris*k of exposure  to CL, the
qualitative as well as quantitative similarities between the toxicity of 0, to
animals and man must  be  considered more  carefully in the  future.   Significant
advances have been made in  understanding the, toxicity of 0, through appropriate
animal models.   This  summary highlights the significant  results  of selected
studies that  will  provide useful  data for  better predicting  and assessing, in
a scientifically sound manner, the possible human responses to 0-,.
     Summary  figures  and tables  are  presented  in the following sections.   The
practical purpose of this presentation of the data  is to help the reader focus
on what  types of effects or responses have been reported,  what concentrations
have been  tested (1.0 ppm  and lower), and as a  convenient list of references
with each  of  the biological parameters  measured.   Studies were selected  for
inclusion in  these figures  and tables on the basis  of specific criteria presen-
ted below:

1.   Studies  have  been cited when the reported  effects are clearly due to 0^
     exposure.   Effects  due to mixtures  of CL  with other  pollutants have  been
     summarized in  a  separate figure and  table.  Studies  involving exercise,
                                   9-217

-------
     diet deficiencies, or other possible modifiers of response to 03 have not
     been included.

2.   Cited studies  report  the effects of CL  exposure  over a broad range of
     animal species  and  strains  and for varying  lengths  of time.   Specific
     details  on  animal species, exposure duration,  and observed biological
     effects  can  be  obtained  from  the  tables  within  the body of  this  chapter,

3.   Each closed symbol on the figures represents one or more studies conducted
     at that particular concentration that caused effects. Specific references
     can be found in the accompanying tables.

4.   Each  open  circle represents  one or more studies  that used the given
     concentration, but reported no significant effects.  No-effect levels are
     also indicated by brackets in the accompanying tables.

5.   Only pulmonary function effects were divided by short-term (<14 days) and
     long-term exposures to follow the discussion in the text.

In order to keep this section brief, and concise, it was necessary to be somewhat
selective  in  determining what and  how this  information would be  presented.  A
number of important factors, such as the specific length of exposure, were not
included.  Also,  the  parameter selected  to  illustrate  a specific response was
usually  broad and very  general.   For example, the  category "decreases in
macrophage function"  includes such diverse  endpoints as measurements  of lyso-
somal and  phagocytic  activity, macrophage mobility, or chemotactic response.
These responses may  or may not be  related to  one  another.   Thus, care must  be
taken in how  these data are used and  interpreted. The  only appropriate  use  is
to gain an overview of the broad array of the effects of ozone and the concen-
trations which did and did not cause these effects.

9.6.2  RegionalDosimetry  in the Respiratory Tract
     The amount of  0» acting at a given site in the lung is related to the
airway luminal concentration at that level.   As a result,  0, does not immediately
interact with cellular components of the respiratory tract.  Instead, it first
comes into contact with the mucous or surfactant  layer  lining  the  airway.   It
should be noted that CL is quite reactive chemically.  Reactions with components
                                   9-218

-------
of this layer cause an increase in total absorption of 03 in the upper airways
and in a  reduction of the amount of CL reaching sensitive tissues.  The site
at which  uptake and subsequent  interaction occur and the  local  dose  (quantity
of CL absorbed per unit area per time), along with cellular sensitivity, will
determine the type and extent of the injury.   Also, the capacity for responding
to a specific dose may vary between animals and humans because of dissimilarities
in detoxification  systems, pharmacokinetics, metabolic rates, genetic makeup,
or other  factors.  Thus,  along with the above,  a  knowledge of the  complex
process of gas transport and absorption is crucial  to understanding the effects
of On and other oxidants in humans.
     The  animal  studies that  have been conducted on  ozone  absorption are
beginning to  indicate  the  quantity and site of 03 uptake in  the respiratory
tract.  Experiments on  the nasopharyngeal  removal  of  CL  in  animals suggest
that the  fraction  of  0, uptake  depends  inversely on flow  rate,  that  uptake is
greater  for  nose  than  for mouth breathing,  and that  trachea!  and chamber
concentrations are positively correlated.   Only one  experiment measured 0~
uptake in the lower respiratory tract,  finding 80  to 87 percent uptake  by the
lower  respiratory  tract of  dogs (Yokoyama and  Frank, 1972).   At present,
however,  there  are no  reported results for human nasopharyngeal  or lower
respiratory tract  absorption.  Caution must be used in estimating nasopharyngeal
uptake for normal  respiration  based  upon experiments employing  unidirectional
flows.
     To further  an understanding of 0,, absorption, mathematical models have
been  developed to  simulate the  processes involved  and  to  predict 03  uptake by
various regions and sites within the respiratory tract.  The model of Aharonson
et al. (1974) has  been used to analyze nasopharyngeal uptake data.   Applied to
Oo data,  the  model indicates that the average mass transfer coefficient in the
nasopharyngeal region increases with  increasing air  flow,  but the actual
percent uptake decreases.
     Three models  have been developed to  simulate lower respiratory uptake
(McJilton et  al. ,  1972; Miller et al., 1978b,  1985).  These  models  are very
similar in their treatment of 0» in the airways (taking into account convection,
diffusion, wall  losses, and  ventilatory patterns)  and  in  their  use of morpho-
logical  data  to  define the dimensions of the airways and liquid lining.  The
models differ in their treatment of the mechanism of absorption. Both  of the
models of Miller and co-workers take  into account chemical  reactions  of 0,
with  constituents  of  the  liquid lining, whereas the model of McJilton  et al.
                                   9-219

-------
does not.  The  models  of Miller et al. differ in their treatment of chemical
reactions, as well as in the fact that the newer model includes chemical reac-
tions of 03 in additional compartments, such as tissue and blood.
     Tissue dose  is  predicted by the models of Miller et al. to  be relatively
low  in  the trachea, to  increase to  a maximum between the  junction  of the
conducting airways and the gas-exchange region, and then to decrease distally.
This is  not  only true  for animal simulations  (guinea pig and  rabbit)  but  it
is also  characteristic of the human simulations  (Miller et  al.,  1978b;  1985).
     A comparison of the results of Miller and co-workers with morphological
data (that shows  the centriacinar region  to be most affected by  03)  indicates
qualitative agreement between predicted tissue doses and observed effects in
the  pulmonary  region.   However, comparisons in  the  tracheobronchial region
indicate that dose-effect correlations may be improved by  considering other
expressions  of  dose such as total  absorption by an  airway and  by further
partitioning of  the  mucous layer compartment in  mathematical models.   Further
research is  needed to define toxic mechanisms, as well as to refine  our know-
ledge of important chemical,  physical, and morphological  parameters.
     At present, there are few experimental results that are useful in judging
the  validity of  the  modeling efforts.  Such results  are needed,  not only  to
understand better the absorption of 03 and its role in toxicity, but also to
support  and  to  lend  confidence  to the  modeling  efforts.   With experimental
confirmation, models which further our understanding  of the role of  0~ in  the
respiratory tract will  become practical tools.
     The consistency and similarity of the human and  animal lower respiratory
tract dose curves obtained thus  far lend  strong  support to  the feasibility of
extrapolating to  man the results obtained on  animals exposed  to 03.   In the
past, extrapolations have usually been qualitative in nature.   With additional
research  in  areas which are basic to  the formulation of dosimetry  models,
quantitative dosimetric  differences  among species can be  determined.  If in
addition, more information is obtained on species sensitivity to a given dose,
significant advances can be made in quantitative extrapolations  and  in making
inferences about the likelihood of effects of 03 in man.   Since animal  studies
are  the  only available approach  for investigating the full  array of  potential
disease  states  induced  by  exposure to 03, quantitative use of animal data is
in the  interest  of better establishing CL levels to  which  man can safely  be
exposed.
                                   9-220

-------
9.6.3  Effects of Ozone onthe Respiratory Tract
9.6.3.1   Morphological  Effects.   The  morphological   changes  which  follow
exposure to  less than 1960 \jg/m   (1.0  ppm) 0, are very  similar  in all  species
of laboratory mammals  studied.   Of the many specific cell types found in the
respiratory  system, two  types,  ciliated cells and type 1 alveolar epithelial
cells, are  the  cells  most damaged morphologically  following  0-, inhalation.
Ciliated cells are  found in  the conducting airways,  e.g., trachea,  bronchi,
and nonrespiratory bronchioles.   Ciliated cells function in the normal clearance
of the airways  and the  removal  of  inhaled  foreign  material.    Following 0-
exposure of experimental animals, damaged ciliated cells have been reported in
all of these conducting airways (Schwartz et  a!.,  1976;  Castleman  et a!,,
1977).  In  rats, damage  to ciliated cells appears most  severe at the junction
of the conducting airways with the gas  exchange area  (Stephens  et al., 1974a;
Schwartz et al., 1976).  Damage to type 1 alveolar epithelial  cells  is limited
to those  cells  located  near  this  junction, i.e., the  centriacinar or proximal
alveolar  region of  the pulmonary acinus (Stephens et al., 1974b; Schwartz et
al.,  1976;  Castleman  et al., 1980; Barry et al., 1983; Crapo et al.,  1984).
Type  1 alveolar  cells form most of the blood-air barrier where gas  exchange
occurs.   Severely damaged  ciliated and type 1 alveolar epithelial cells are
shed  (sloughed) from  the tissue  surface and are replaced  by multiplication of
other cell  types  less damaged by 0,  (Evans et al., 1985).  This process has
been  most extensively studied in the  centriacinar  region where nonciliated
bronchiolar  cells  and type 2 alveolar  epithelial cells become  more  numerous
(Evans et al.,  1976a,b,c;  Lum et al.,  1978).   Some  of these nonciliated bron-
chiolar and  type 2 cells differentiate  into ciliated and type 1 cells, respec-
tively.  Cell multiplication  in bronchioles may be more than that required for
replacement  of  damaged ciliated cells, and nonciliated bronchiolar  cells may
become hyperplastic  (Castleman  et al., 1977;  Ibrahim et al., 1980;  Eustis et
al, ,  1981)  and  sometimes appear as nodules (Zitm'k et  al., 1978; Moore and
Schwartz, 1981; Fujinaka et al., 1985).  Inflammatory changes characterized by
a  variety of leukocytes with alveolar  macrophages  predominating,  intramural
edema, and  fibrin  are also seen  in the centriacinar  region (Stephens et al.,
1974a; Schwartz  et al. ,  1976; Castleman et al., 1977; Fujinaka  et  al., 1985).
      The  damage  to  ciliated  and  centriacinar  type 1 alveolar  epithelial cells
and the inflammatory changes  tend to occur soon after exposure  to concentrations
                                   9-221

-------
                         3
of 0, as  low as 392 pg/m  (0.2 ppm).  Damage to centriacinar type 1 alveolar
epithelium in rats has been well documented as early as 2 hours after exposure
                                3
to Og concentrations  of  980 \ig/m   (0.5 ppm)  (Stephens  et  al.s  1974a).   In  the
same publication the  authors  report damage to centriacinar  type  1 alveolar
                                                     o
epithelial cells after  2 hours exposure to  392 ug/m  (0.2 ppm) 03, but this
portion of their  report is not documented by published micrographs (Stephens
et al.,  1974a).  Loss of  cilia from cells in the  rat terminal  bronchiole
                                      3
occurs following exposure  to  980 |jg/m  (0.5 ppm) 03 for 2 hours (Stephens et
al.,  1974a).  Damage  to  ciliated cells has  been  seen  following exposure of
                                 3
both rats and monkeys to 392 ug/m  (0.2 ppm) 03, 8 hr/day for 7 days (Schwartz
et al., 1976; Castleman et al., 1977).  Centriacinar  inflammation has been
                                                       3
reported as early as 6 hours after exposure to 980 ng/m  (0.5 ppm) 0- (Stephens
                                                      3
et al., 1974b) and 4 hours after exposure to 1568 ug/m  (0.8 ppm) 03 (Castleman
et al., 1980).
     During long-term exposures, the damage to ciliated cells and to centriacinar
type 1 cells and centriacinar inflammation continue, though at a reduced rate.
Damage to cilia has been reported  in monkeys  following 90-day  exposure  to  980
    3
ug/m  (0.5 ppm) 0,,, 8 hr/day  (Eustis et al.,  1981)  and in rats exposed  to  980
    q            3
ug/ra  (0.5 ppm) 03, 24 hr/day for 180 days (Moore and Schwartz, 1981).   Damage
to centriacinar type  1 cells  was reported  following exposure of young rats to
490 ug/m  (0.25 ppm)  Q~, 12 hrs/day  for 42 days (Barry et al., 1983; Crapo et
                      •s                                                 ,,
al.,  1984).   Changes  in  type  1 cells were  not detectable  after 392 ug/nr (0.2
                                                                        3
ppm) 03,  8 hr/day for 90 days but  were seen  in rats exposed  to 980 ug/m (0.5
ppm) for  the  same  period (Boorman et al., 1980).   Centriacinar inflammatory
                                                            3
changes persist during 180-day exposures of rats to 980 pg/m  (0.5 ppm) 03, 24
hr/day (Moore and  Schwartz,  1981)  and one-year exposures of monkeys to 1254
ug/ii  (0.64 ppm) 03, 8 hr/day (Fujinaka et al., 1985),
     Remodeling of  distal  airways  and centriacinar  regions  occurs following
long-term exposures to  03.   Rats develop respiratory bronchioles between the
terminal bronchiole to alveolar duct junction seen  in control rats (Boorman et
al., 1980; Moore  and  Schwartz, 1981).   In monkeys,  distal airway  remodeling
results in  increased  volumes  of respiratory  bronchioles  which have thicker
walls and a  smaller internal  diameter (Fujinaka et al., 1985).  The walls of
centriacinar alveoli  are also thickened (Schwartz  et  al., 1976;  Boorman et
al.» 1980; Barry et al., 1983; Crapo et al., 1984; Last et al., 1984a).  Studies
of the  nature of  these  thickened  interalveolar septa  and bronchiolar walls
                                   9-222

-------
revealed increases in inflammatory cells, fibroblasts, and amorphous extracel-
lular matrix   (Last  et al., 1984a;  Fujinaka  et al., 1985).  Three  studies
provide morphological  evidence of  mild fibrosis  (i.e.,  local  increase of
collagen)  in  centriacinar interalveolar septa  following  exposure  to < 1960
    3
pg/m  (< I ppm)  of 03 (Last et  al.5 1979;  Boorman  et al.,  1980;  Moore and
Schwartz, 1981).   Changes in collagen location or amounts, or both, which occur
with the remodeling of the distal airways, were reported in two of those studies
(Boorman et al.,  1980; Moore and Schwartz, 1981).
     While morphometry of small  pulmonary arteries  is not  commonly studied in
0,,-exposed animals,  pulmonary  artery walls  thickened by muscular  hyperplasia
                                                      3
and edema were reported in rabbits exposed to 784 pg/rn  (0.4 ppm) 03, 6 hr/day,
5 days/week for  10 months (P'an'et  al.,  1972).   Thickened  intima and media in
                                                                  3'
pulmonary arterioles were reported in monkeys exposed to 1254 ug/m   (0.64 ppm)
03, 8 hr/day for 1 year (Fujinaka et al., 1985).
     Several of the effects of 0,, inhalation persisted after the 0~  inhalation
ended and  the  animals  breathed only filtered  air several  days or weeks.  Lungs
                               3
from rats  exposed  to 1568 (jg/m  (0.8 ppm)  0, for 72 hours appeared normal 6
days after the end of the exposure (Plopper et al.,  1978).  However, incomplete
resolution of  the  nonciliated  bronchiolar epithelial hyperplasia was reported
in monkeys 7 days after 50 hours exposure to 1568 |jg/m  (0.8 ppm) 0, (Castleman
                            .........        ,                         3
et al.5  1980)  and in mice 10  days after a 20-day exposure to 1568 jjg/m  (0.8
ppm) 03,  24 hr/day  (Ibrahim et  al., 1980).  Centriacinar  inflammation  and
distal airway  remodeling were  still  apparent  62 days after a 180-day exposure
to 980 |jg/m3 (0.5 ppm) Og, 24  hr/day (Moore and Schwartz,  1981).
     While not all species of  laboratory mammals  have been studied following  a
single 0,  exposure regimen  or  using the same  morphological techniques  because
investigators  have  asked different biological  questions,  there is a striking
similarity of  morphological  effects in the respiratory system of  all species
studied.   The  cell  types most damaged are the  same.  One  of these cells, the
type  1 alveolar epithelial  cell,  has  a wide distribution in the  pulmonary
acinus and yet is damaged only in one specific  location in all species  studied.
The  other, the ciliated  cell, appears  damaged  wherever  it is  located  in the
conducting airways.  Damage to these cells  is seen within  hours after exposure
to  concentrations  of 0-  much  lower  than 1 ppm  and continues during  exposures
of weeks  or months.   Hyperplasia of other  cell types is  reported to  start
early  in the exposure  period,  to continue throughout a long-term exposure, and
                                    9-223

-------
when  studied,  to persist following  postexposure  periods  of days or weeks.
Centriacinar inflammation is  also seen early and is reported throughout long
exposure periods.   Duration  of centriacinar inflammation during postexposure
periods has  been studied less often and appears dependent upon length of the
exposure period.
     Other effects  which have been reported in fewer  studies  or in a more
limited number of species include distal airway remodeling and thickened pulmo-
nary  arteriolar  walls.   Remodeling of distal airways  has only  been  reported
in rats and  monkeys after long-term exposures.  In rats, remodeling of distal
airways has  been reported to persist for several  weeks after the 03 exposure
has  ended.   Thickened pulmonary  arteriolar walls  have been reported only
twice,  once  after  long-term  exposure of rabbits  and  once  after long-term
exposure of monkeys.
      Studies on  the morphologic effects of 0~ exposures of experimental animals
are  summarized  in Figure 9-4 and Table 9-28 (see Section 9.6.1 for  criteria
used  to summarize the studies).
9.6.3.2   Pulmonary  Function.   One of  the  limitations  of animal studies  is
that  many  pulmonary function tests comparable to those conducted after acute
exposure of human subjects are difficult to  interpret.   Methods exist, however,
for  obtaining  similar measurements of many  variables pertinent  to understand-
ing  the effects of ozone on the respiratory tract, particularly after longer
exposure  periods.   A number  of  newer studies  reported here reflect recent
advances in studying the effects of 03 on pulmonary function in small animals.
      Changes in  lung function following ozone exposure have been studied  in
mice,  rats,  guinea  pigs, rabbits,  cats, dogs, sheep, and monkeys.  Short-term
                                                        3
exposure for 2 hr to concentrations of 431  to  980 ug/m  (0.22 to 0.5 ppm)
produces rapid,  shallow  breathing and increased  pulmonary  resistance  during
exposure  (Murphy et al., 1964; Yokoyama,  1969; Watanabe et al., 1973; Amdur
et al., 1978).    The onset of these effects is rapid and the  abnormal breathing
pattern usually  disappears  within 30 min after cessation of exposure.   Other
changes in lung  function measured following  short-term  ozone exposures lasting
3 hr  to 14 days  are usually greatest 1 day following exposure and disappear by
7 to 14 days following exposure.   These effects are associated  with  premature
closure of the small, peripheral airways and include increased  residual volume,
closing volume,  and closing capacity (Inoue  et al., 1979).
      Studies of airway reactivity following short-term ozone exposure  of  1 to
2 hr  duration  in experimental  animals show that 03  increases the reactivity of
                                   9-224

-------
                                    **.aSSP A
6.1-
0.2-
g 0.3-
Q.
Q.
| 0.4-
*

4 I'-' i 4 1 4 4 4 4 4 4 ' r 4 1 1 1 1 ( 1 4 » 4 > 4 1 1 1 4 1 4 i i 1 < t 1 l 4 l 1 1 l 1 » 1 1 t 1 1 • 1 > Figure 9-4. Summary of morphological effects in experimental animals exposed to ozone. See Table 9-28 for reference citations of studies ~ summarized here. $-225


-------
         TABLE 9-28.  SUMMARY TABLE:  MORPHOLOGICAL EFFECTS OF OZONE
                            IN EXPERIMENTAL ANIMALS
  Effect/response
03 concentration, ppm
         References
Damaged ciliated
  and type 1 cells
Proliferation of non-
  ciliated bronchiolar
  and type 2 cells
Centriacinar
  inflammation
 [0.2], 0.5, 0.8
 0.2, 0.5, 0.8
 0.2, 0.35
 0.25
 0.25
 0.26, 0.50, 1.0
 0.5
 0.5
 0.5
 0.5, 0.8
 0.5, 0.8
 0.54, 0.88
 0.8
 0.8
 0.85

 0.2, 0.35
 0.35, 0.50, 0.70,
  0.75, 1.0
 0.5
 0.5
 0.5
 0.5, 0.8
 0.54, 0.88
 0.64
 0.7
 0.8
 0.8
 0.8
 1.0

 [0.2], 0.5, 0.8
 0.2
 0.2, 0.5, 0.8
 0.25
 0.25
 0.35
 0.5
                         0
                         0.5, 0.8
                         0.5, 0.8
                         0.5, 0.8
                         0.54, 0.88
                         0.54, 0.88
                         0.64
                         0.8
                         1.0
Boorman et al.  (1980)
Schwartz et al.  (1976)
Castleman et al.  (1977)
Barry et al. (1983)
Crapo et al. (1984)
Boatman et al.  (1974)
Stephens et al.  (1974b)
Moore and Schwartz (1981)
Evans et al. (1985)
Eustis et al. (1981)
Mel lick et al.  (1975, 1977)
Stephens et al.  (1974a)
Castleman et al.  (1980)
Plopper et al.  (1978)
Stephens et al.  (1978)

Castleman et al.  (1977)
Evans et al. (1976b)

Evans et al. (1985)
Zitnik et al. (1978)
Moore and Schwartz (1981)
Eustis et al. (1981)
Freeman et al.  (1974)
Fujinaka et al.  (1985)
Evans et al. (1976a)
Castleman et al.  (1980)
Lum et al. (1978)
Ibrahim et al.  (1980)
Cavender et al.  (1977)

Boorman et al.  (1980)
Plopper et al.  (1979)
Schwartz et al.  (1976)
Barry et al. (1983)
Crapo et al. (1984)
Castleman et al.  (1977)
Stephens et al.  (1974b)
Moore and Schwartz (1981)
Mellick et al.  (1975, 1977)
Brummer et al.  (1977)
Last et al. (1979)
Stephens et al,  (1974a)
Freeman et al.  (1974)
Fujinaka et al.  (1985)
Castleman et al.  (1980)
Freeman et al.  (1973)
                                   9-226

-------
         TABLE 9-28.   SUMMARY TABLE:   MORPHOLOGICAL EFFECTS OF OZONE
                      IN EXPERIMENTAL ANIMALS (continued)
Effect/response
Distal airway
remodel ing
Thickened pulmonary
arteriolar walls
03 concentration, ppm
[0.2], 0.5, 0.8
0.2, 0.5, 0.8
0.5
0.64, 0.96
0.64
1.0
0.4
0.64
References
Boorman et al. (1980)
Schwartz et al. (1976)
Moore and Schwartz (1981)
Last et al. (1984a)
Fujinaka et al. (1985)
Freeman et al . (1973)
P'an et al. (1972)
Fujinaka et al. (1985)
the lungs to a number of stimuli.   Mild exercise, histamine aerosol inhalation,
and breathing air with reduced oxygen or elevated carbon dioxide concentrations
caused rapid, shallow  breathing  in conscious dogs immediately following 2-hr
exposures to  1100  to 1666 ng/m3 (0.56  to 0.85 ppm)  of  03  (Lee  et  al.,  1979,
1980).  Aerosolized  ovalbumin caused an increased  incidence  of  anaphylaxis  in
mice preexposed to 980 or 1568 pg/m  (0.5 or 0.8 ppm) of 03 continuously for 3
to 5  days  (Osebold et al. ,  1980).   In  addition,  increased airway sensitivity
to histamine  or  cholinomimetic drugs administered  by aerosol  or injection has
been  noted  in several  species after exposure  to  980  to  5880  [jg/m  (0.5  to 3.0
ppm)  of  03  (Easton and Murphy,  1967; Lee et al., 1977; Abraham  et al., 1980,
1984a,b; Gordon  and  Amdur,  1980;  Gordon et al.,  1981, 1984; Roum and Murlas,
1984).  The mechanism responsible for 0.,-induced bronchial reactivity is still
uncertain but may  involve more than one specific factor.  Ozone  has been shown
to cause increased sensitivity of vagal sensory endings in the  dog airway (Lee
et al., 1977, 1979,  1980),  Ozone exposure may also  enhance the  airway  respon-
siveness to  bronchoconstrictors  by altering sensitivity of the  airway  smooth
muscle.-directly  or through  re-leased cellular mediators (Gordon  et al. , 1981,
1984;  Abraham et al.,  1984a,b).   In  some species,  increased  airway hyperreac-
tivity may be explained by increased transepithelial  permeability  or decreased
thickness of  the airway mucosa  (Osebold et  al.,  1980; Abraham et al.,  1984b).
Ozone  exposure  may  also  decrease  airway hyperreactivity  by  causing  mucous
hypersectetion,  thereby  limiting the airway penetration of inhaled bronchocon-
strictors (Abraham et  al., 1984a).
                                   9-227

-------
     The time course  of airway hyperreactivity after exposure to 980 to 5880
pg/m (0.5 to 3.0 ppm) of 0~  suggests  a possible association with  inflammatory
cells and pulmonary  inflammation  (Holtzman et a!., 1983a,b; Sielczak et al.,
1983; Fabbri et al.,  1984; O'Byrne et al., 1984a,b; Murlas and Roum, 1985).
However, the time  course  of responsiveness is variable  in different species
and the  relationships between  airway  inflammation  and reactivity  at different
concentrations of  03  are not well  understood.  Additional studies that demon-
strate increased collateral  resistance  following  30 min local exposure of 03
or  histamine  in sublobar bronchi  of  dogs  (Gertner et al., 1983a,b,c,1984)
suggest that other mechanisms, along with amplification of reflex pathways,
may  contribute  to  changes  in  airway reactivity depending not only  on the
concentration of 03  in  the airways but  also  on the extent of penetration of
ozone into the lung periphery.
     The effects  of short-term  exposures  to 03 on pulmonary function and
airway reactivity  in  experimental  animals  are summarized in  Figure  9-5  and
Table 9-29  (see Section 9.6.1 for criteria used in developing this summary).
                                                                           3
     Exposures of  4  to  6 weeks to ozone concentrations  of 392 to 490 pg/m
(0.2 to  0,25 ppm)  increased  lung distensibility at high  lung  volumes in young
rats (Bartlett  et  al.,  1974; Raub et al.,  1983a).   Similar increases in lung
                                                                    3
distensibility were  found  in older rats exposed to 784 to 1568 pg/m  (0.4 to
0.8 ppm) for up to 180  days (Moore and  Schwartz,  1981;  Costa et al.,  1983;
                                                                        3
Martin et al., 1983).  Exposure to 03 concentrations of 980 to 1568 ug/m  (0.5
to  0.8 ppm) increased pulmonary resistance and caused impaired stability of
the small peripheral  airways in both rats and monkeys ( Wegner,  1982; Costa
et al., 1983; Yokoyama  et  al., 1984;  Kotlikoff et  al., 1984).  The effects  in
monkeys were  not   completely reversed by 3 months  following  exposure; lung
distensibility had also  decreased  in the postexposure period, suggesting the
development of lung fibrosis which has also been suggested morphologically and
biochemically.
     The effects of  long-term  exposures  to ozone  on pulmonary  function  and
airway reactivity  in  experimental  animals  are summarized in  Figure  9-6  and
Table 9-30  (see Section 9.6.1 for criteria used in developing this summary).
9.6.3.3  Biochemical  Effects  The lung is metabolically  active,  and- several
key steps in metabolism have been studied after 03 exposure.   Since the proce-
dures for such  studies  are invasive,  this research has been conducted only  in
animals.  Effects, to be  summarized below, have been observed on antioxidant
metabolism,  oxygen consumption,  proteins,  lipids,  and xenobiotic metabolism.
                                   9-228

-------
                    ^

jp
          <*&
                                          0°'
          X*
(rf
   \«



u.u —
0.1-
0.2-
E 0.3-
a
a
i 0.4-
o
1
S 0.5-
0)
u
8 0.6-
o>
<3 0.7-
0.8-
0.9-
1.0-
'

«
i


<
<
<
<


i
i . (
i

i
t " <
•
t 	 a


C
(
1
(

(
1
(



I
r

•

i

>
i 	 1


>
I



<





C
i <



l (



1
«J
)
1



i (
(
(
i 	 1

)




i
i
i
i 	
Figure 9-5. Summary of effects of short-term ozone exposures on
pulmonary function in experimental animals. See Table 9-29 for
reference citations of studies summarized here.

-------
          TABLE 9-29.  SUMMARY TABLE:   EFFECTS ON PULMONARY FUNCTION
           OF SHORT-TERM EXPOSURES TO OZONE IN EXPERIMENTAL ANIMALS
  Effect/response       03 concentration, ppm
                            References
Increased breathing
  frequency
0.22, 0.41, 0.8
0.34, 0.68, 1.0
0.5
Amdur et al.  (1978)
Murphy et al.  (1964)
Yokoyama (1969)
Decreased tidal volume     0.34, 0.68, 1.0      Murphy et al.  (1964)
Decreased lung
  compliance
Increased residual
  volume (RV),
  closing capacity
  (CC), and closing
  volume (CV)

Decreased diffusion
  capacity
[0.22], 0.41, 0.8
0.26, 0.5, 1.0
1.0

0.24 - 1.0
Amdur et al.  (1978)
Watanabe et al.  (1973)
Yokoyama (1974)

Inoue et al.  (1979)
0.26, 0.5, 1.0
Watanabe et al.  (1973)
Increased pulmonary
resistance


Increased airway
reactivity


:
[0.22]
0.26, 0.5, 1.0
0.5
1.0
[0.1]-0.8
[0.1]-0.8, 1.0
0.5, 1.0
0.7
1.0
Amdur et al. (1978) :
Watanabe et al . (1973)
Yokoyama (1969)
Yokoyama (1974)
Gordon and Amdur (1980)
Gordon et al. (1981, 1984)
Abraham et al. (1980, 1984a,b)
Lee et al. (1977)
Holtzman et al. (1983a,b)
     The lung contains several compounds (e.g., vitamin E, sulfhydryls, gluta-

thione)  and  enzymes (e.g.,  glutathione peroxidase, glutathione reductase,

glucose-6-phosphate dehydrogenase, and  superpxide dismutase) that function as

antioxidants, thereby  defending  the  lung against oxidant  toxicity  from the

oxygen in  air,  from oxidants produced  during  metabolic  processes,  and from

oxidizing  air pollutants  such as ozone.  Obviously, this  protection is only

partial  for  0-  since exposure to  ozone  causes  numerous effects on lung  struc-

ture, function,  and biochemistry.   Acute exposure to high ozone levels (2920

ug/m , 2 ppm)  typically decreases antioxidant metabolism, whereas  repeated
                                                     o
exposures  to lower  levels  (between 272 and 1568 ug/m  ,  0.2 and 0.8 ppm) in-

creases this metabolism (DeLucia et al., 1975b).  In rats maintained on normal
                                   9-230

-------
VD

M
CO
                                                              "" >x V*
:yp \ , • • i.
0*~ o6"
>
i
4
1




• . . <
<

1





»
1

                              Figure 9-6. Summary of effects of long-term ozone exposures on
                              pulmonary function in experimental animals. See Table 9-30 for
                              reference citations of studies summarized here.

-------
          TABLE 9-30.  SUMMARY TABLE:  EFFECTS ON PULMONARY FUNCTION
            OF LONG-TERM EXPOSURES TO OZONE IN EXPERIMENTAL ANIMALS
  Effect/response       0'3 concentration, ppm
                              References
Increased lung volume
Increased pulmonary
  resistance
Decreased lung
  compliance

Decreased inspiratory
  flow

Decreased forced
  expiratory volume
  (FEV,) and flow
  (FEFjr
[0.08], [0.12], 0.25
0.2
[0.2], 0.8
0.4

0.2, 0.8
0.5, 1.0
0.64
0.64

0.5, 0.8
0.64
0.2, 0.8
0.64
Raub et al.  (1983a)
Bartlett et al.  (1974)
Costa et al.  (1983)
Martin et al.  (1983)

Costa et al.  (1983)
Yokoyama et al. ,  1984
Wegner (1982)
Kotlikoff et al.5 1984

Eustis et al.  (1981)
Wegner (1982)
[0.08], 0.12, 0.25       Raub et al.  (1983)
Costa et al. (1983)
Wegner (1982)
diets, this response has been observed after a week of continuous or intermit-
tent exposure to 392 ug/m3 (0.2 ppm) 03 (Mustafa, 1975; Mustafa and Lee, 1976;
Plopper et al., 1979).  Similar responses are seen in monkeys and mice, but at
                                3
higher concentrations  (980 [jg/m ,  0.5 ppm)  (Fukase et al., 1978; Mustafa and
Lee, 1976).
     The effects of  03 on oxygen  consumption  have been  studied  since  oxygen

consumption  is  a fundamental  parameter of cellular metabolism,  reflecting
energy production by cells.   As with antioxidant metabolism, acute exposure to
                                3
high  ozone levels  Q>  3920 |jg/m ;  > 2 ppm)  decreases metabolism  (and  thus,
                                                                   3
oxygen consumption); repeated exposure to lower levels (> 1568 [jg/m , 0.8 ppm)
increases  oxygen  consumption (Mustafa et al.5  1973;  Schwartz  et al.,  1976;
Mustafa and  Lee, 1976)%  Effects   in  rats on  normal  diets  have been observed
                                                                 3
after  a  short-term  exposure to ozone  levels  as low as  392 Mg/m   (0.2  ppm)
(Schwartz et al., 1976; Mustafa et al., 1973; Mustafa and Lee,  1976).   Monkeys
                                                 3
are affected at a higher level of  ozone (980 yg/m , 0.5 ppm).
                                   9-232

-------
     Similar patterns of  response  for both antioxidant metabolism and oxygen
consumption are observed  after exposure to ozone.   A 7-day exposure to ozone
produces linear concentration-related  increases  in activities of glutathione
peroxidase,  glutathione  reductase,  glucose-6-phosphate  dehydrogenase, and
succinate oxidase  (Mustafa and Lee,  1976;  Chow et  al., 1974;  Schwartz  et al. ,
1976; Mustafa et al., 1973).   Rats on a vitamin E-deficient diet experience an
increase in  enzyme activities  at  196 ug/m  (0.1 ppm) ozone as compared to
        Q
392 ug/m  (0.2 ppm) in animals on normal diets (Chow et al., 1981; Mustafa and
Lee, 1976; Mustafa,  1975).   Research on these enzymes has  shown that there is
no  significant  difference in  effects from  continuous  versus intermittent
exposure; this, along with concentration-response data, suggests that the con-
centration of  ozone  is  more important  than  duration of  exposure in causing
these  effects  (Chow et al., 1974;  Schwartz  et  al.,  1976;  Mustafa and  Lee,
1976).
                                                                           /
     Duration of exposure still plays a role, however.  During exposures up to
1  or 4 weeks,  antioxidant  metabolism and 0^ consumption  generally do not
change on the first day of exposure; by about day 2,  increases are observed and
by  about day 4 a plateau is  reached (Mustafa and  Lee, 1976;  DeLucia et al.,
1975a).  Recovery  from these  effects occurs by 6 days post-exposure  (Chow
et  al., 1976b).  This  plateauing of effects in the  presence  of exposure does
not result  in  long-term tolerance.   If rats are re-exposed after recovery is
observed,  the  increase  in enzyme activities  is  equivalent  to  that observed in
animals exposed for the first  time  (Chow et  al., 1976b).
     The influence  of  age on  responsiveness is  also similar  for  antioxidant
metabolism and  oxygen  consumption (Elsayed et  al., 1982a;  Tyson  et al., 1982;
Lunan  et al., 1977).  Suckling neonates (5 to 20 days old) generally exhibited
a  decrease in enzyme activities; as  the animals grew older (up to about 180 days
old),  enzyme activities generally  increased  with age.  Species differences may
exist  in this response (Mustafa  and Lee,  1976;  Mustafa et al. ,  1982;  Chow
et al. ,  1975;  DeLucia et al. ,  1975a).   Studies  in  which monkeys  have been
compared  to  rats  did  not include a description of  appropriate  statistical
considerations applied (if any);  thus,  no  definitive conclusions  about  respon-
siveness of  monkeys  versus rats can be  made.
     The mechanism responsible for the increase  in antioxidant metabolism and
oxygen consumption  is not known.   The  response is  typically attributed, however,
to concurrent  morphological  changes, principally the loss  of type 1 cells and
an increase  in type  2 cells  that are richer  in the enzymes measured.
                                    9-233

-------
     Monooxygenases constitute  another class of  enzymes  investigated after
ozone exposure.  These  enzymes  function in the metabolism of both endogenous
(e.g., biogenic amines, hormones) and  exogenous (xenobiotic) substances.  The
substrates acted upon are  either activated or detoxified,  depending  on the
                                                                 3
substrate and  the  enzyme.    Acute  exposure to 1470 to  1960 (jg/m  (0.75 to
1 ppm) ozone decreased  cytochrome  P-450 levels and enzyme activities related
to both cytochrome  P-450  and P-448.   The  health  impact of these changes is
uncertain since only a few elements of a complex metabolic system were measured.
     The activity of lactate dehydrogenase is increased in lungs of vitamin E-
                                                            3
deficient rats receiving a short-term exposure to 196  ug/m  (0.1 ppm)  ozone
(Chow et al., 1981).  Higher levels caused a similar response in rats, but not
in monkeys, on normal diets (Chow et a!,, 1974, 1977).   This enzyme is frequent-
ly used as a marker of cellular damage because it is released upon cytotoxicity.
It is not known,  however,  whether  the increase  in  this enzyme  is a  direct
reflection of  cytotoxicity or  whether it is an  indicator  of  an increased
number of type 2 cells and macrophages in the lungs.
     An increase in  a few  of the measured- activities of lysosomal enzymes has
                                                      3
been shown in the lungs of rats exposed to > 1372 ug/m  (0.7 ppm) ozone (Dillard
et'al., 1972;  Castleman et al.} 1973a; Chow et al,s  1974).  This response is
most  likely  the result of an  increase in inflammatory cells  in the lungs
rather than  an induction  of enzymes,  since  lysosomal  enzymes in alveolar
macrophages decrease after jr\ vivo  or jji vitro exposure to ozone (Hurst et al.,
1970; Hurst and Coffin,  1971).
     As discussed  previously,   long-term exposure to  high 0, concentrations
causes mild  lung fibrosis  (i.e., local increase  of collagen in  centriacinar
interalveolar  septa).   This morphological  change has been correlated  with
biochemical  changes  in  the activity  of prolyl  hydroxylase (an  enzyme  that
catalyzes the  production of hydroxyproline) and in hydroxyproline content (a
component of collagen that is  present  in  excess  in  fibrosis)  (Last et al.,
1979; Bhatnagar et  al., 1983).   An increase  in collagen  synthesis  has been
                       3
observed, with 980 |jg/m  (0.5 ppm)  03  being the minimally effective concentra-
tion tested  (Hussain et al., 1976a,b;  Last et al., 1979),   During a prolonged
exposure, prolyl hydroxylase activity  increases by day 7 and returns to control
levels by 60 days  of exposure.  When a short-term exposure ceases,  prolyl
hydroxylase  activity returns to normal by  about  10 days  post-exposure, but
hydroxyproline levels remain elevated  28 days post-exposure.  Thus, the product
                                   9-234

-------
of the increased  synthesis,  collagen,  remains relatively stable.  One study
(Costa et al. ,  1983)  observed a small  decrease in collagen levels of rats at
                  o
392 and 1568  (jg/m   (0.2 and 0.8 ppm) 03 after  an  intermittent exposure for 62
days.
     The effects  of 0., on increasing  collagen content may be progressive;
i.e.,  after  a 6-week  intermittent exposure of rats  to 0.64 or 0.96 ppm  03
ceased, collagen  levels 6  week post-exposure were elevated over  the levels
immediately after exposure (Last  et al.,  1984b).  Also, there appears to be
little difference between  continuous  and  intermittent exposure in increasing
collagen levels in rat lungs  (Last et al,  1984b).  Thus, the intermittent clean
air periods were not sufficient to permit recovery.
     Although the ability  of  03 to  initiate peroxidation of unsaturated fatty
acids in vitro  is well established, few  in vivo  studies of lung  lipids have
been conducted.   Generally, ozone decreases unsaturated fatty acid content of
the lungs (Roehm et al., 1972) and decreases incorporation of fatty acids into
lecithin (a  saturated  fatty acid) (Kyei-Aboagye  et al., 1973).  These altera-
tions, however, apparently do not alter the surface-tension-lowering properties
of  lung  lipids  that are important to  breathing  (Gardner et al.,  1971; Huber
et al., 1971).
     One of  the earliest  demonstrated effects of ozone was that very high
concentrations  caused  mortality as  a  result  of  pulmonary  edema.   As  more
                                                            o
sensitive techniques were  developed,  lower levels (510 ug/m , 0.26 ppm) were
observed to increase the protein content of the lung (Hu et al., 1982).   Since
some of the excess protein could be attributed to serum proteins, the interpre-
tation was that edema had occurred.  This effect was more pronounced several
hours after  exposure  ceased.   At higher  concentrations, a  loss  of carrier-
mediated transport from the air side of the lung  to the blood side was observed
(Williams et  al., 1980).  These changes imply an  effect on the barrier function
of  the  lung, which  regulates fluxes  of  various  substances with potential
physiological activities across the alveolar walls.
     The biochemical  effects  observed  in experimental animals  exposed to 03
are summarized  in Figure  9-7 and Table 9-31  (see Section  9.6.1  for  criteria
used in developing this summary).
9.6.3.4   Host Defense Mechanisms.   Reports  over the  years  have presented
substantial  evidence that exposure to  ozone impairs the antibacterial activity
of  the  lung, resulting in  an  impairment of the lung's  ability  to  kill inhaled
                                   9-235

-------
NJ

U)

CTl
a
a
 «


o
                0
                U

                o
                U




                i

                O
                             XX*

0 0

0.1-
0,2-
0.3-
0.4-
0.5-
0.6-
0.7-
0.8-
0.9-
1.0


c
i

<
i
i


<



3
i C

i
i
i <

(
i 1

*


> C

1
(
5 <
1
i
> 1
1



5 <

<
l
\ \
<
i

! <
1


C
3
i
3
1 1
) l

i
	 1

1
3 0
I
1
1
l
1 0 I
I

* <



(
l t
i
i i
(
t i

l
i
l l
	 _<


l <
i l
t
1
l (
l i

»
1
l i
i

1
i
i

l
• »


t

                            Figure 9-7. Summary of biochemical changes in experimental animals

                            exposed to ozone. See Table 9-31 for reference citations of studies

                            summarized here.

-------
               TABLE 9-31.   SUMMARY TABLE:   BIOCHEMICAL CHANGES
                   IN EXPERIMENTAL ANIMALS  EXPOSED TO OZONE
  Effect/response
    concentration, ppm
   References
Increased 02
  consumption
Increased lysosomal
  enzyme activities
Increased lung
  hydroxyproline
  and prolyl
  hydroxylase
  activity
Altered mucus
  glycoprotein
  secretions

Increased alveolar
  protein and
  permeability
  changes

Increased LDH
  activity
Increased NADPH
  cytochrome c
  reductase
  activity

Increased GSH
 metabolism
[0.1], 0.2
[0.1], 0.2, 0.35, 0.5, 0.8
0.2, 0.5, 0.8
0.2, 0.5, 0.8
0.45
0.8
0.8

[0.2], [0.5], 0.8
0.7, 0.8
0.7, 0.8

[0.2], 0.5, 0.8
0.2, 0.8
0.45, 0.8
0.5, 0.64, 0.96
0.5
0.8

[0.2], [0.4], 0.5, 0.6, 0.8
0.5, 0.6, 0.8
0.6, 0.8

[0.1], 0.26, 0.51, 1.0
[0.25], 0.5, 1.0
0.6, 1.0
1.0

[0.1]
[0.5], 0.8
0.8

0.2, 0.35, 0.8
0.2, 0.5, 0.8
0.2, 0.5, 0.8
[0.1]
0.1, 0.2
0.2, 0.35, 0.5,
0.2, 0.5, 0.8
0.2, 0.5, 0.8
0.2, 0.5, 0.8
0.2, 0.5, 1.0
0.32
0.45
0.5
                                      0.8
Mustafa (1975)
Mustafa and Lee (1976)
Mustafa et al. (1973)
Schwartz et al. (1976)
Mustafa et al. (1982)
Chow et al. (1976b)
Elsayed et al. (1982a)

Chow et al. (1974)
Oil lard et al. (1972)
Castleman et al.  (1973a,b)

Hussain et al. (1976a,b)
Costa et al. (1983)
Bhatnagar et al.  (1983)
Last et al. (1979, 1984b)
Last and Greenberg (1980)
Hesterberg and Last (1981)

Last and Kaizu (1980)
Last and Cross (1978)
Last et al. (1977)

Hu et al. (1982)
Alpert et al.  (1971a)
Williams et al. (1980)
Reasor et al.  (1979)

Chow et al. (1981)
Chow et al. (1977)
Chow and Tappel (1973)

Mustafa and Lee (1976)
Schwartz et al. (1976)
DeLucia et al. (1972, 1975a,b)
Chow et al. (1981)
Plopper et al. (1979)
Mustafa and Lee (1976)
Chow et al. (1974)
DeLucia et al. (1972, 1975a,b)
Schwartz et al. (1976)
Fukase et al.  (1975)
Moore et "al.  (1980)
Mustafa et al. (1982)
Chow et al. (1975)
                                   9-237

-------
               TABLE 9-31.  SUMMARY TABLE:  BIOCHEMICAL CHANGES
             IN EXPERIMENTAL ANIMALS EXPOSED TO OZONE (continued)
  Effect/response
03 concentration, ppm
References
                      0.5, 1.0
                      0.7, 0.75, 0.8
                           Fukase et al.  (1978)
                           Chow and Tappel (1972, 1973)





Increased NPSH



Decreased
unsaturated
fatty acids
0.8

0.8
0.9
0.9
0.1, 0.2
0.2, 0.5, 0.8
0.45
0.8
0.5-


Elsayed et al. (1982a,b;
1983)
Chow et al. (1976b)
Tyson et al. (1982)
Lunan et al. (1977)
Plopper et'al. (1979)
DeLucia et al . (1975b)
Mustafa et al. (1982)
Chow et al. (19765)
Roehm et al . , 1972


microorganisms.  Suppression  of  this  biocidal  defense  of  the  lung can  lead to
microbial proliferation within the lung, resulting in mortality.  The mortality
response is  concentration-related  and is significant at concentrations as low
as 157 to  196 |jg/m3 (0.08 to  0.1  ppm)  (Coffin et  al., 1967;  Ehrlich et al.,
1977; Miller  et  al.,  1978a;  Aranyi et  al.,  1983).  The biological basis for
this response  appears  to be  that  ozone or  one of its reactive products can
impair or suppress the normal bactericidal  functions of the pulmonary defenses,
which results  in  prolonging  the life of the infectious agent, permitting its
multiplication and ultimately, in this animal model, resulting in death.   Such
infections can occur  because of 0, effects  on a complex  host defense .system
involving  alveolar  macrophage functioning,  lung  fluids,   and other immune
factors.
     The data  obtained in various experimental animal studies indicate that
short-terra ozone exposure can reduce the effectiveness of several vital defense
systems  including  (1) the ability  of the  lung to  inactivate bacteria and
viruses  (Coffin  et  al.,  1968; Coffin and Gardner,  1972b; Goldstein et al.,
1974a, 1977; Warshauer et §_!,._,. 1974j_Bergers et alj; 1983.   Schwartz and Christman,
1979; Ehrlich et al., 1979);  (2)  the mucociliary transport  system  (Phalen
et al., 1980; Frager et al.,  1979; Kenoyer et  al., 1981;  (3)  the immunological
system (Campbell  and  Hilsenroth, 1976; Fujimaki et al.,  1984; Thomas et al.,
                                   9-238

-------
1981b; Aranyi et al. ,  1983;  and (4) the pulmonary macrophage (Dowel! et al. ,
1970; Goldstein et al., 1971a,b, and 1977; Hadley et al.,  1977; McAllen et al.,
1981;-Witz-et al., 1983;  Hurst et al,,  1570; Hurst and Coffin, 1971; Amoruso
et al., 1981).  Studies  have also indicated that the activity level of the
test  subject and the presence  of  other airborne chemicals are  important vari-
ables that can  influence  the determination of the  lowest  effective concen-
tration of the  pollutant  (Gardner et al.,  1977; Aranyi et al., 1983;  Ehrlich,
1980, 1983; Grose  et al.,  1980, 1982;  Phalen et al.," 1980; Goldstein et al. ,
1974a; Illing et al. , 1980).
     Ciliated cells  are  damaged by 0™ inhalation,  as demonstrated by major
morphological changes  in  these cells, including necrosis and sloughing, or by
the  shortening  of  the  cilia in cells attached  to the bronchi.   Sufficient
ciliated cell damage should  result in decreased transport of viable and non-
viable particles from  the lung.  Rats exposed to 784,  1568, 1960,  or 2352
    3
ug/m   (0.4,  0.8,  1.0,  or 1.2  ppm)  for times as short as 4  hr have decreased
short-term clearance of  particles from  the lung (Phalen et al., 1980; Frager
et al. , 1979;  Kenoyer  et al.,  1981).  Short-term clearance is mostly due to
mucus  transport of particles,  and the decreased  short-term clearance is  an
anticipated functional result  predicted  from morphological  observations.  The
mucous glycoprotein  production of the trachea is  also altered by  0- exposure.
Mucous glycoprotein  biosynthesis, as measured ex vivo  in  cultured  trachea!
explants from exposed rats, was inhibited  by short-term continuous  exposure to
1568  ug/m3  (0.8 ppm) of 03  for 3 to 5  days (Last and Cross,  1978;  Last and
Kaizu, 1980; Last  et a!., 1977).  Glycoprotein synthesis and secretion recovered
to control values  after 5 to 10 days of exposure and increased to  greater than
control values  after 10 days of exposure.  With this increase  in production of
mucus, investigators have found that the  velocity  of the  trachea! mucus was
                                                             3
significantly reduced following a 2 hr exposure to 1960 M9/m   d-0  PPm) (Abraham
at al., 1980).
      A problem remains in assessing the  relevance  of these animal  data to
humans.  Green  (1984) reviewed the  literature and compared  the host antibacterial
defense systems of the rodent  and  man  and found that these two  species had
defenses that are  very similar and  thus provide a good basis  for  a qualitative
extrapolation.  Both defenses  consist of  an aerodynamic filtration system, a
fluid layer  lining the respiratory  membranes, a transport mechanism for "removing
foreign particles, microorganisms,  and  pulmonary  cells,  and immune secretions
                                   9-239

-------
of lymphocytes and plasma cells.   In both rodents and humans, these components
act in concert to maintain the lung free of bacteria.
     If the animal models are to be used to reflect the toxicological response
occurring in  humans,  then the endpoint for comparison of such studies should
be morbidity  rather  than mortality.   A better  index  of  0-  effect in humans
might be the  increased prevalence of infectious respiratory  illness in the
community.   Such a comparison may be proper since both mortality from respira-
tory infections  (animals)  and morbidity from respiratory infections (humans)
can result from  a loss in pulmonary defenses  (Gardner,  1984).   Whether the
microorganisms used in the various animal studies are comparable to the organ-
isms responsible  for  the  respiratory infections  in a  community  still  requires
further investigation.
     Ideally, studies  of  pulmonary host defenses should be performed in man,
using epidemiological  or volunteer  methods  of study.  Unfortunately,  such
studies have  not been reported yet.   Attention must therefore be paid to the
results of host-defense experiments conducted with animals.
     In the  area of  host  defense  of  the  lung against  infection,  present know-
ledge of the physiology, metabolism, and function have come primarily from the
study of various animal  systems,  but it is generally accepted that the basic
mechanisms of action of these defense cells and systems function  similarly in
both animals  and man.   There are also  human data to  support  this  statement,
especially in such areas  as immunosuppression, ciliostasis,  and alveolar
macrophages.   The effects seen  in animals represent alterations  in basic
biological systems.   One  can assume  that similar alterations  in basic  defense
mechanisms could occur in  humans since they  possess equivalent pulmonary
defense systems.   It is understood,  however, that different  exposure  levels
may be  required  to produce  similar responses in  humans.  The  concentration of
0, at which  effects  become evident can be influenced by a number of factors,
such as preexisting disease, virulence of the infectious agent, dietary factors,
concurrent exposure  to other pollutants, exercise, or the presence  of other
environmental stresses, or a combination of these.   Thus, one could hypothesize
that humans exposed to CU could experience effects on host defense mechanisms.
At the  present  time,  however, one cannot  predict the exact  concentration at
which effects may occur in man nor the severity of the effects.
     The effects of  0- on host defense  mechanisms  in  experimental  animals are
summarized in Figure 9-8 and Table 9-32 (see Section 9.6.1 for criteria used
in developing this summary).
                                   9-240

-------
                                        *'
ID
a
a

o
              o
              i
              01
              o
              N
             O
                                          /v
                                              <&?
                                              fP   ^    

( < ( i i i ** »*>' X I 1 1 j I 1 1 , AW s \ 1 1 1 < 1 t 1 1 >' • ( 1 1 < 1 < i \ 1 < ! IIT nl ii 1 t X'' ( ( 1 ( ) < 1 ( 1 \ ( < 1 $ ( 1 ., 1 ^5^ • • i < > • i < * * i 4 1 »xs 1 1 1 t J? Figure 9-8. Summary of effects of oione on host defense mechanisms in experimental animals. See Table 9-32 for reference citations of studies summarized here.


-------
        TABLE 9-32.  SUMMARY TABLE:  EFFECTS OF OZONE ON HOST DEFENSE
                      MECHANISMS IN EXPERIMENTAL ANIMALS
  Effect/response
03 concentration, ppm
                            References
Delayed mucociliary
  clearance; accelerated
  alveolar clearance,
  ciliary beating
  frequency
Inhibited bactericidal
  activity
Altered macrophage
  membrane
Decreased macrophage
  function
Altered no. of defense
  cells
Increased suscepti-
  bility to infection
  [0.1]
  0.4, 0.8, 1.0
  [0.5]
  [0.5], 1.0
  0.8
  1.2

  0.4
  0.4
  0.5
  0.62
                          0.
                          0.
  0.99

  0.1, 1.0
  0.5
  0.5
  0.5, 1.0

  0.25, 0.5

  0.5
  0.5, 0.67
  0.5, 0.67
  0.8
  1.0
  1.0
       0.35,
       0.35
0.2
0.2,
0.2,
0.2, 0.5,
0.25
0.5
0.5, 0.88
0.5
0.5, 0.88
0.5, 0.8
0.54, 0.88
0.8
1.0
1.0

0.08
0.08, 0.1
0.1
 0.5, 0.8

0.8
Grose et al.  (1980)
Kenoyer et al.  (1981)
Friberg et al.  (1972)
Abraham et al.  (1980)
Phalen et al. (1980)
Frager et al. (1979)

Coffin and Gardner (1972b)
Goldstein et al.  (1972b)
Friberg et al.  (1972)
Goldstein et al.  (1971b)
Bergers et al.  (1983)
Warshauer et al.  (1974)
Goldstein et al.  (1971a)

Gardner et al.  (1971)
Dowel! et al. (1970)
Hadley et al. (1977)
Goldstein et al.  (1977)

Hurst et al.  (1970)
Hurst and Coffin (1971)
Alpert et al. (197Jb)
Coffin et al. (1968)
Coffin and Gardner (1972b)
Schwartz and Christman (1979)
Shingu et al. (1980)
McAllen et al.  (1981)

Plopper et al.  (1979)
Dungworth et al.  (1975b)
Castleman et al.  (1977)
Boorman et al.  (1977, 1980)
Barry et al.  (1983)
Zltnik et al. (1978)
Stephens et al.  (1974a)
Last et al. (1979)
Brummer et al.  (1977)
Eustis et al. (1981)
Freeman et al.  (1974)
Castleman et al.  (1980)
Freeman et al.  (1973)
Cavender et al.  (1977)

Coffin et al. (1967)
Miller et al. (1978a)
Ehrlich et al.  (1977)
                                   9-242

-------
        TABLE 9-32.   SUMMARY TABLE:   EFFECTS OF OZONE ON HOST DEFENSE
                MECHANISMS IN EXPERIMENTAL ANIMALS (continued)
Effect/response
Increased suscepti-
bility (cont'd)






Altered immune
activity




03 concentration, ppm
0.1
0.1, 0.3
[0.2], 0.4, 0.7
0.3
0.5
[0.64]
0.7, 0.9
1.0
0.1
0.5, 0.8
0.5, 0.8
0.59

0.8
References
Aranyi et al . (1983)
111 ing et al. (1980)
Bergers et al. (1983)
Abraham et al . (1982)
Wolcott et al. (1982)
[Sherwood et al. (1984)]
Coffin and Blommer (1970)
Thomas et al. (1981b)
Aranyi et al . (1983)
Osebold et al. (1979, 1980)
Gershwin et al. (1981)
Campbell and Hilsenroth
(1976)
Fujimaki et al. (1984)
9.6.3.5  Tolerance.   Examination of responses to short-term, repeated exposures
to 0, clearly  indicates  that with some of  the  parameters measured, animals
have an increased capacity to resist the effects of subsequent exposure.   This
tolerance persists for varying  times,  depending on the degree of development
of the tolerance.  Previous  exposure to low concentrations  of 0, will protect
against the effects of subsequent exposure to lethal doses and the development
of lung edema  (Stokinger et al., 1956; Fairchild,  1967;  Coffin and Gardner-,
1972a; Chow, 1984).   The prolongation of mucociliary clearance reported for 03
can also be  eliminated  by pre-exposure to  a lower concentration  (Frager et
al., 1979).  This  effect is demonstrated for a  short period of time and is
lost as soon as  the mucus  secretion  rate  returns  to normal.  However, not all
of the toxic effects of 0~, such as reduced functioning activity of the pulmonary
defense system  (Gardner  et al.,  1972);  hyperplasia of the type 2 cells (Evans
et al., 1971, 1976a,b); increased susceptibility to respiratory disease (Gardner
and  Graham,  1977);  loss of  pulmonary  enzymatic activity (Chow, 1976, Chow
et al., 1976b); and inflammatory response (Gardner et al., 1972) can be totally
prevented by prior treatment with  low  levels of  0,,.   Dungworth et al. (1975b)
and Castleman  et al.  (1980) have attempted  to  explain tolerance  by careful
examination of the morphological changes that occur with  repeated 03 exposures.
These investigators suggest that during continuous exposure to 0., the injured

                                   9-243

-------
cells attempt  to  initiate early repair of the  specific lesion.   The repair
phase results in a reduction of the effect first observed but lasts only for a
short time  since,the  recovered  cells  are as  sensitive to re-exposure to 0-  as
the pre-exposed counterpart  (Plopper  et a!., 1978).  This  information is an
important observation because  it implies that the decrease in susceptibility
to 0, persists only as  long  as  the exposure  to  0, continues.  The biochemical
studies of Chow et al. (1976b) support this conclusion.
     At this time, there are a number  of hypotheses proposed to explain the
mechanism of  this phenomenon (Mustafa  and  Tierney, 1978;  Schwartz et al.,
1976; Mustafa  et al., 1977; Berliner et al.,  1978; Gertner et al., 1983b;
Bhatnagar et al., 1983).  Evidence by Nambu and Yokoyama (1983) indicates that
although the pulmonary antioxidant system (glutathione peroxidase,  glutathione
reductase, and glucose-6-phosphate dehydrogenase)  may play an active role in
defending the  lung against ozone, it  does not explain the mechanism of toler-
ance in that the development of tolerance does not coincide with the described
biochemical  enhancement of the antioxidant system in the lungs of rats.
     From this literature, it would appear that tolerance,  as seen in animals,
may not be  the result of any one single biological process, but instead may
result from a  number  of different events,  depending on the specific response
measured.   Tolerance  does  not  imply  complete or absolute protection,  because
continuing injury does  still occur, which could potentially  lead to nonrever-
sible pulmonary changes.
     Tolerance may not  be long-lasting.  During 0- exposure, the increase in
antioxidant metabolism  reaches  a plateau and recovery occurs a few days after
exposure ceases.   Upon re-exposure,  effects observed are similar to those that
occurred during the primary exposure (Chow et al., 1976b).

9.6.4  Extrapulmonary Effects of Ozone
     It is  still  believed that  CL, on contact with  respiratory system tissue,
immediately reacts and  thus  is  not absorbed or transported to extrapulmonary
sites to any significant degree.  However,  several studies  suggest that possibly
products  formed  by  the interaction of  0- and  respiratory  system fluids or
tissue can produce effects in lymphocytes,  erythrocytes, and serum,  as well  as
in the parathyroid gland,  the heart,  the liver, and the CNS.  Ozone exposure
also produces  effects on  animal  behavior  that may be  caused  by pulmonary
consequences of Q~, or by nonpulmonary (CNS) mechanisms.   The  mechanism by

                                   9-244

-------
which 0,  causes  extrapulmonary  changes  is  unknown.   Mathematical  models  of  0,
dosimetry predict  that very  little  03 penetrates  to  the  blood  of  the  alveolar
capillaries.  Whether these effects result from 0- or a reaction product of 0-,
which penetrates to the blood and is transported  is the subject of speculation.
9.6.4.1   Central Nervous  System and Behavioral  Effects.   Ozone significantly
affects  the behavior of  rats during exposure  to concentrations  as  low as
        3
235 |jg/m  (0.12  ppm)  for  6 hr.   With  increasing concentrations of 0,,  further
decreases in  unspecified  motor  activity and  in  operant  learned behaviors have
been  observed (Konigsberg and  Bachman, 1970;  Tepper  et a!.,  1982;  Murphy
et al . ,  1964;  and  Weiss et a!., 1981).   Tolerance to the observed decrease  in
motor activity  may occur  on  repeated  exposure.   At low  0, exposure concentra-
               3
tions (490  pg/m  ,  0.25 ppm),  an  increase in activity is observed after exposure
                                           q
ends.   Higher 0_ concentrations (980 ug/m ,  0.5  ppm) produce  a decrease in
rodent  activity that persists  for  several  hours  after  the end of exposure
(Tepper  et  al. ,  1982, 1983).
      The  mechanism by which  behavioral performance is  reduced is unknown.
Physically  active  responses appear  to- enhance the effects of Q3, although this
may  be  the  result of an  enhanced  minute volume that increases the effective
concentration  delivered  to the  lung.   Several  reports  indicate that  it is
unlikely  that, animals have  reduced physiological  capacity to  respond,  prompt-
ing Weiss et  al. (1981) to suggest  that 03 impairs the inclination to  respond.
Two  studies  indicate that  mice will respond to  remove themselves from  an
atmosphere  containing greater than 980  |jg/m  (0.5 ppm)  (Peterson  and Andrews,
1963, Tepper  et al.,  1983).   These studies suggest  that  the aversive  effects
of  03 may be due  to  lung irritation.   It  is  unknown whether lung  .irritation,
odor, or a  direct effect  on  the CNS causes change in rodent behavior at lower
03  concentrations.
9.6.4.2   Cardiovascular Effects.  Studies on the effects of 03 on the cardio-
vascular system are few,  and to date there are  no reports of attempts to con-
firm these  studies.  The exposure  of rats to 0,  alone or in combination with  ,
                   3
cadmium  (1176 |jg/m ,  0.6  ppm 03) resulted in measurable increases in systolic
pressure and  heart rate (Revis  et  al., 1981).   No additive or antagonistic
response was  observed with the  combined  exposure.   Pulmonary  capillary  blood
                                                                       3
flow and PaO  decreased 30 rnin following exposure of dogs to 588 MS/I"  (°-3
 ppm)  of 0-j  (Friedman  et  al.,  1983).   The  decrease  in  pulmonary  capillary  blood
 flow  persisted for as long as 24  hr  following  exposure.
                                    9-245

-------
9.6.4.3  Hematological  and  Serum  Chemistry Effects.   The data  base  for the
effects of 0^  on  the hematological system Is extensive and indicates that 0.,
or one  of its  reactive products can cross  the  blood-gas  barrier,  causing
changes in the  circulating  erythrocytes (RBC) as well as significant differ-
ences in various components of the serum.
     Effects of On  on the circulating RBCs can be readily identified by exa-
mining either  morphological and/or biochemical  endpoints.   These  cells are
structually and metabolically well understood  and  are available through rela-
tively non-invasive  methods, which makes them  ideal  candidates  for both human
and animal studies.   A  wide range of  structural  effects  have been reported  in
a variety of  species of animals, including  an increase  in the fragility of
                                               3
RBCs isolated from monkeys exposed to 1470 ug/m  (0.75 ppm) of 03 4 hr/day for
4 days (Clark et a!.,  1978).   A  single 4-hr exposure to 392 M9/m3 (0.2 ppm)
also caused increased fragility as well as sphering of RBCs of rabbits  (Brinkman
et a!., 1964).  An increase in the number of RBCs with Heinz bodies was detected
                                        3
following a  4-hr exposure  to 1666 pg/m  (0.85 ppm).   The presence of  such
inclusion bodies in  RBCs is an indication  of  oxidant stress (Menzel  et a!.,
1975a).
     These morphological changes  are frequently accompanied by a wide range of
                                                           2
biochemical effects.   RBCs  of  monkeys exposed to 1470 |jg/m  (0.75 ppm) of 0,
for 4  days  also had a decreased level of  glutathione (GSH)  and decreased
acetylcholinesterase  (AChE) activity,  an enzyme bound to  the RBC membranes.
The RBC GSH  activity remained  significantly lower 4 days postexposure  (Clark
et al,, 1978).
     Animals deficient  in  vitamin E  are more sensitive to 0-.   The RBCs from
these animals,  after being  exposed to  0,,  had a significant increase in the
activity of GSH peroxidase, pyruvate kinase, and lactic dehydrogenase,  but had
                                                 3
a decrease in  RBC GSH after exposure  to 1568 pg/m   (0.8  ppm) for 7 days (Chow
and Kaneko,  1979).   Animals with a vitamin  E-supplemented diet did  not have
any changes in glucose-6-phosphate dehydrogenase (G-6-PD), superoxide dismutase,
                                                    3
or catalase  activities.  At a lower  level  (980 |jg/m ,  0.5  ppm), there were  no
changes in GSH  level or in the activities of GSH peroxidase or GSH reductase
(Chow et al., 1975).  Menzel et al. (1972) also reported a significant  increase
in lysis of RBCs from vitamin E-deficient animals after 23 days of exposure to
        3
980 ug/m  (0.5 ppm).  These effects were not observed  in vitamin E-supplemented
rats.  Mice  on  a vitamin E-supplemented diet  and  those  on a deficient diet
                                   9-246

-------
                                                                   3
showed an increase in G-6-PD activity after an exposure of 627 ijg/m  (0.32 ppm)
of Q~ for 6  hr.   Decreases observed in ACHE activity occurred in both groups
(Moore et al.,  1980).
     Other blood changes  are  attributed to 0-.  Rabbits exposed for 1 hr to
        3
392 ug/m  (0.2 ppm)  of  0« showed a significant drop in total blood serotonin
                                                                    3
(Veninga, 1967).   Six- and 10-month exposures of rabbits to 784 pg/m  (0.4 ppm)
of 0~ produced an  increase in serum protein  esterase  and in serum trypsin
inhibitor.   This latter  effect may be a result  of  thickening of the small
pulmonary arteries.  The same exposure caused a significant decrease in albumin
levels and an  increase  in alpha and gamma  globulins (P'an and Jegier, 1971,
1976; P'an et al., 1972;  Jegier, 197-3).  Chow et al. (1974)  reported that the
serum lysozyme level  of rats increased significantly after 3 days of continuous
exposure to 0., but was not affected when the exposure was intermittent (8 hr/day,
                                                            3
7 days).  The 03 concentration in both studies was 1568 yg/m  (0.8 ppm) of Q-.
     Short-term exposure  to low concentrations  of  03  induced an  immediate
change  in the serum  creatine phosphokinase  level in mice.  In this  study, the
DO doses were  expressed as the product of concentration and time.   The C x T
value for this effect ranged from 0.4 to 4.0 (Veninga et al., 1981).
     A few of the hematological effects observed in animals (i.e.,  decrease in
GSH and AChE activity and the formation of Heinz bodies) following exposure to
03 have also been seen following j_n vitro exposure of RBCs from humans (Freeman
and Mudd, 1981;  Menzel  et al., 1975b; Verweij and Van  Steveninck,  1981).  A
common  effect observed  by a number of  investigators is  that 03  inhibits  the
membrane ATPase activity  of RBCs (Koontz and Heath, 1979; Kesner et al., 1979;
Kindya  and  Chan,  1976;  Freeman et al., 1979; Verweij  and Van Steveninck,
1980).  It has been  postulated that this inhibition of ATPase could be related
to the  spherocytosis :and increased fragility of  RBCs seen  in animal  and  human
cells.
     Although these  i_n vitro data are useful  in studying mechanisms of action,
it is difficult to extrapolate these data to any effects observed in man.  Not
only is the method of exposure not physiological, but the actual concentration
of 0, reaching the RBC cannot be determined with any accuracy.
9.6.4.4  Cytogenetic and  Teratogenic Effects.  Uncertainty still exists regard-
ing  possible reproductive, teratogenic,  and mutational  effects  of  exposure to
ozone.   Based  on various  rn vitro  data,  a number of chromosomal effects of
ozone have been described for  isolated cultured cell lines,  human lymphocytes,
                                   9-247

-------
and microorganisms  (Fetner,  1962;  Hamelin et  al., 1977a,b, Hamelin and Chung,
1975a»b, 1978; Scott  and Lesher, 1963; Erdman  and Hernandez, 1982; Guerrero
et a!., 1979;  Dubeau  and Chung, 1979, 1982),   The interpretation, relevance,
and predictive values  of such studies to human health are questionable since
(1) the concentrations  used  were many-fold  greater than what is found in the
ambient air  (see  Chapter 10);  (2)  extrapolation of  i_n vitro exposure concen-
trations to  human exposure  dose is not yet possible; and (3) direct exposure
of isolated  cells to  ozone  is highly  artifactual since  it bypasses all  the
defenses of  the  host that would  normally be  functioning in protecting the
individual  from the inhaled gas.  Furthermore, the direct exposure of isolated
cells jm vitro to  ozone may result  in chemical  reactions between ozone and
culture media that might not occur jjn vivo.
     Important questions still  exist  regarding  ijn vjvo cytogenetic effects of
ozone  in  rodents and  humans.   Zelac et al.  (1971a,b) reported chromosomal
abnormalities in peripheral  leukocytes of hamsters exposed to 0, (0.2 ppm).
Combined exposures  to  ozone  and  radiation (227-233 rads)  produced an additive
effect on the  number  of chromosome breaks in  peripheral  leukocytes.   These
specific findings were not confirmed  by Gooch  et al.  (1976) or by Tice et al.
(1978), but  sufficient differences  in  the various experimental protocols make
a direct comparison difficult.  The latter group did report significant increases
in the  number  of chromatid  deletions  and achromatic  lesions  resulting from
exposure to 0.43 ppm ozone.
     Because the  volume of  air  inspired  during pregnancy is significantly
enhanced,  the pregnant animal  may be  at greater risk to  low levels of ozone
exposure.   Early studies on  the possible teratogenic effects of  ozone have
suggested that exposures as low as 0.2 ppm can reduce infant survival  rate and
cause unlimited incisor growth (Brinkman et al,, 1964; Veninga,  1967).   Kavlock
et al. (1979, 1980) found that pregnant rats exposed to 1.0 and 1.49 ppm ozone
showed a significant increase in embryo resorption rate,  slower growth, slower
development  of righting reflexes,  and delayed grooming and rearing behavior,
but no increase in neonatal  mortality was observed.
9.6.4.5  Other Extrapulmonary Effects.  A series  of studies was conducted to
show that 03 increases drug-induced sleeping time in a  number of species of
animals (Gardner  et al.,  1974;  Graham, 1979;  Graham  et  al.,  1981,  1982a,b,
                          3
1983, 1985).  At 1960 pg/m  (1.0 ppm), effects were observed after 1,  2,  and 3
days of exposure.   As  the concentration of Q.,  was reduced, increasing  numbers
of daily 3-hr exposures were required to produce a significant effect.   At the
                                   9-248

-------
                                       3
lowest concentration studied  (196  (jg/m ,  0.1 ppm),  the increase was observed
at days 15 and 16 of exposure.  It appears that this effect is not specific to
the strain  of mouse or to the  three  species of animals tested,  but it is
sex-specific, with  females being  more susceptible.   Recovery was complete
within 24 hr  after  exposure.   Although a number of mechanistic  studies have
been conducted,  the  reason for this effect on pentobarbital-induced sleeping
time is  not known.   It has been hypothesized that some common aspect related
to  liver drug metabolism  is  quantitatively reduced (Graham et al., 1983).
     Several  investigators have attempted to  elucidate the  involvement  of  the
endocrine system in Og  toxicity.   Most of these studies  were designed to
investigate the  hypothesis that the -survival  rate of mice  and  rats exposed to
lethal concentrations  of 0™  could be increased by use  of various thyroid
blocking agents or by thyroidectomy.  To follow up these findings, demons and
Garcia (1980a,b) and demons and Wei  (1984)  investigated  the effects of  a
                           3
24-hr exposure to 1960 \ig/m  (1.0 ppm) of 0, on the hypothalamo-pituitary-thyroid
system of rats.  These three organs regulate the function of each other through
various  hormonal feedback mechanisms.   Ozone caused decreases in serum concen-
tration  of  thyroid  stimulating  hormone (TSH),  in circulating thyroid hormones
(T- and  T*) and  in protein-bound iodine.  No alterations were observed in many
other  hormone levels measured. - Thyroidectomy prevented the effect  of  0.,  on
TSH and  T.  and hypophysectomy prevented effects on T-, unless the animals were
supplemented  with T. in their drinking water.  The thyroid gland itself was
altered  (e.g.,  edema)- by 0-.   The authors  hypothesyzed  that 03  alters  serum
binding  of  these hormones.
     The extrapulmonary effects of ozone in experimental animals are summarized
in  Figure 9-9 and Table 9-33.  Criteria  used in developing the  summary were
presented in  Section 9.6.1.

9.6.5  Interaction of Ozone With Other Pollutants.
     Combined exposure studies in  laboratory animals have produced varied
results, depending  upon the  pollutant combination evaluated and the measured
variables.   Additive and/or  possibly  synergistic effects  of  03  exposure  in
combination with N0?  have  been described  for increased susceptibility to
bacterial infection  (Ehrlich et al., 1977, 1979; Ehrlich, 1980, 1983), morpho-
logical  lesions  (Freeman et  al., 1974), and  increased antioxidant metabolism
(Mustafa et al., 1984).  Additive or possibly  synergistic effects from exposure
                                   9-249

-------
f
M
U1
P
                                                                                       ,^e
                                                                   •&*    V

0.1 _
0.2-

E 0.3-
o.
a
c 0.4-
,2
**>
S
c O.S-
u
c
Q
8 0.6-
0
c
o
a 0-7-
0.8-

0.9_
1 n
I \ i
I
t

(



i



i



<




; :
• i
<
i
^


(
<
i «



i

<

i
<
<
i


i
i
i
• <
i <
<
i


i
, •





i


t
i
i

> 	 1
» *
•
» i
•



• ,








i



I
i

i




»








•
i


•> i 	 «« — ~
                                   Figure 9-9. Summary of extrapulmonary effects of ozone in

                                   experimental animals. See Table 9-33 for reference citations of studies

                                   summarized here.

-------
           TABLE 9-33.  SUMMARY TABLE:  EXTRAPULMONARY EFFECTS OF OZONE
                              IN EXPERIMENTAL ANIMALS
  Effect/response
03 concentration, ppm
         References
CNS effects
Hematological effects
Chromosomal, reproduc-
  tive, terato "logical
  effects
Liver effects
Endocrine system
  effects
                                  0
                                   0.5, 0.7
 0.05, 0.5
 0.1 - 1.0
 0.12 -
 0.2, 0.
 0.5
 0.5
 0.5
 0.6
 1.0
 1.0
 0.06, 0.12, 0.48
 0.2
 0.2, 1.0

 0.25, 0.32, 0.5
 0.4
 0.4
 0.5
 0.64
 0.75
 0.8
 0.8
 0.85
 0.86
 1.0
 1.0
 1.0

 0.1
 0.2
 0.24, 0.3
 0.43
 0.44
 1.0

 0.1, 0.25, 0.5, 1.0

 0.82
 1.0

 0.75
 0.75
 0.75
 0.75
 1.0
 1.0
Tepper et al
Murphy et al
Tepper et al
Konigsberg and Bachman (1970)
Weiss et al. (1981)
              (1982)
              (1964)
              (1983)
Reynolds and Chaffee (1970)
Xintaras et al.  (1966)
Peterson and Andrews (1963)
Fletcher and Tappel (1973)
Trams et al. (1972)

Calabrese et al. (1983a)
Brinkman et al.  (1964)
Veninga (1967, 1970)
Veninga et al. (1981)
Moore et al. (1980; 1981a,b)
Jegier (1973)
P'an and Jegier (1972, 1976)
Menzel et al. (1972)
Larkin et al. (1983)
Clark et al. (1978)
Chow and Kaneko (1979)
Chow et al. (1974)
Menzel et al. (1975a)
Schlipkoter and Bruch (1973)
Dorsey et al. (1983)
Mizoguchi et al. (1973)
Christiansen and Giese (1954)

Brinkman et al.  (1964)
Veninga (1967)
Zelac et al. (1971a)
Tice et al. (1978)
Kavlock et al. (1979)
Kavlock et al. (1980)

Graham (1979)
Graham et al. (1981, 1982a,b)
Veninga et al. (1981)
Gardner et al. (1974)

Atwal and Wilson (1974)
Atwal et al. (1975)
Atwal and Pemsingh  (1981, 1984)
Pemsingh and Atwal  (1983)
demons and Garcia  (1980a,b)
demons and Wei (1984)
                                     9-251

-------
to DO and  HpSO-  have also been reported for host defense mechanisms (Gardner
et a!.,  1977; Last and Cross, 1978; Grose et al., 1982), pulmonary sensitivity
(Osebold et  al.  1980),  and collagen synthesis  (Last  et al.,  1983), but not
for morphology (Cavender et al., 1977;  Moore and Schwartz, 1981).   Mixtures of
03 and (NH-)2 SO- had synergistic effects on collagen synthesis and morphometry,
including percentage of fibroblasts (Last et al., 1983, 1984a).
     Combining 03  with other particulate  pollutants  produces a variety of
responses, depending  on the endpoint measured.  Mixtures  of  03,  Fe2(SO,)3,
HpSCL, and (NH-^SO. produced the same-effect on clearance rate as exposure to
0~ alone.  However,  when  measuring changes in host defenses,  the combination
of GO with NOp and ZnSO- or 03 with S02 and (NH-^SO, produced enhanced effects
that can not be attributed to 03 only.
     However, since  these  issues  are complex, they must be addressed experi-
mentally using exposure regimens for combined pollutants that are more represen-
tative of ambient ratios of peak concentrations, frequency, duration, and time
intervals between events.
     The interactive  effects  of 03 with other  pollutants  are summarized in
Figure 9-10 and Table 9-34.

9.6.6  Effects of Other Photochemical Oxidants
     There have  been  far  too few controlled  toxicological  studies  with the
other oxidants to permit any sound scientific evaluation of their contribution
to the toxic action  of photochemical oxidant  mixtures.   When  the effects seen
after exposure to 03 and PAN are examined and compared,  it is obvious that the
test  animals  must  be  exposed  to concentrations  of  PAN much  greater  than  those
needed with 03 to produce a similar effect on lethality, behavior modification,
morphology, or significant alterations in host pulmonary defense system (Campbell
et al., 1967; Dungworth et al., 1969; Thomas et al., 1979, 1981a).   The concen-
trations of  PAN  required to  produce these  effects  are many  times greater than
what has been measured in the atmosphere (0.047 ppm).
     Similarly, most of the investigations reporting H202 toxicity have involved
concentrations much higher than those found in the ambient air, or the investi-
gations were conducted by using various i_n vitro techniques for exposure.  Very
limited information is available on the health significance of inhalation expo-
sure to gaseous Hp02.  Because  H202 is highly soluble, it is generally assumed
that  it does not penetrate into the alveolar regions of  the lung but is instead
                                   9-252

-------

NJ

UI

U)
                     o
                     «

                     s
                    o

                                       \pp
V.Vf —
0.1-
0.2-

0.3-
0.4-
0.5-
0.6-

0.7-
0.8-
O.S-
1.0,



c


c

(


(
(
1 1


)


) • • <
•
•
1 (

<
I
1 (
1
I 0



o
1 0

1

) O

1
                                 Figure 9-10. Summary of effects in experimental animals exposed to

                                 ozone combined with other pollutants. See Table 9-34 for reference

                                 citations of studies summarized here.

-------
               TABLE 9-34.  SUMMARY TABLE:  INTERACTION OF OZONE
                 WITH OTHER POLLUTANTS IN EXPERIMENTAL ANIMALS
Effect/response
Pollutant concentrations
  References
Increased
pulmonary
lesions
Increased
pulmonary
sensitivity

Increased anti-
oxidant metabolism
and QZ consumption

Altered mucus
secretion

Increased collagen
synthesis
Increased
susceptibility to
respiratory
infections
   [0.25 ppm 03
     +2.5 ppm N02]
   [0.5 ppm Q3
     + 1  mg/m3 H2S04]
   [0.5 ppm 03
     + 10 mg/m3 H2S04
   0.64, 0.96 ppm 03
     + 5 mg/m3 (NH4)2 S04
   0.9 ppm 03
     +0.9 ppm N02
   1.2 ppm 03
     + 5 mg/m3 (NH4)2S04

   0.5 ppm 03
     + 1 mg/m3 H2S04


   0.45 ppm 03
     +4.8 ppm N02
   0.5 ppm 03
     +1.1 mg/m3 H2S04

   [0.5], [0.8], 1.5 ppm 03
     + 5 mg/m3 (NH4)2S04
   0.5 ppm 03
     + 1 mg/m3 H2S04
   0.64, 0.96 ppm 03
     + 5 mg/m3 (NH4)2S04

   0.05 ppm 03
     + 3760 pg/m3 (NH4)2S04
   0.05 ppm 03
     + 100-400 Mg/m3 N02
     + 1.5 mg/m3 ZnS04
   0.1 ppm 03
     +0.9 mg/m3 H2S04
     (sequential exposure)
   0.1 ppm 03
     +4.8 mg/m3 H2S04
   0.1 ppm 03
     + 940 M9/m3 N02
   0.1 ppm 03
     +13.2 mg/m3 S02
     +1.0 mg/m3 (NH4)2S04
Freeman et al.  (1974)

Moore and Schwartz (1981)

Cavender et al.  (1978)

Last et al. (1984a)

Freeman et al.  (1974)

Last et al. (1983)


Osebold et al.  (1980)



Mustafa et al.  (1984)



Last and Cross (1978);
Last and Kaizu (1980)

Last et al. (1983)

Last et al. (1983)

Last et al. (1984a)
Ehrlich et al. (1977, 1979);
Ehrlich (1980)
Ehrlich et al. (1983)
                                                     Gardner et al. (1977)


                                                     Grose et al.  (1982)

                                                     Ehrlich (1980)

                                                     Aranyi et al. (1983)
                                   9-254

-------
              TABLE 9-34.   SUMMARY TABLE:  INTERACTION OF OZONE
                       WITH OTHER POLLUTANTS (continued)
Effect/response       Pollutant concentrations         References

Altered upper            [0.1 ppm 03                 Grose et al. (1980)
respiratory                +1.1 mg/m3 H2S04]
clearance                  (sequential exposure)
mechanisms               0.4 ppm 03                  Goldstein et al. (1974b)
                           +7.0 ppm N02
                         0.5 ppm 03                  Last and Cross (1978)
                           + 3 mg/m3 H2S04
                         [0.8 ppm 03                 Phalen et al. (1980)
                           +3.5 mg/m3
                             (Fe2(S04)3
                              + H2S04
                              + (NH4)2S04}]
deposited on  the  surface of the upper airways (Last et al., 1982).  Unfortu-
nately, there  have  not been studies designed to look for possible effects in
this region of the respiratory tract.
     A few i_n vitro studies have reported cytotoxic, genotoxic, and biochemical
effects of H202  when using isolated cells  or  organs (Stewart et al., 1981;
Bradley et al.,  1979;  Bradley and Erickson, 1981; Speit et al. , 1982; MacRae
and Stich, 1979).  Although these studies can provide useful data for studying
possible mechanisms  of action,  it is  not yet  possible to extrapolate these
responses to those that might occur in the mammalian system.
     Field and  epidemiological  studies have shown that  human  health  effects
from exposure  to  ambient mixtures of  oxidants  and other airborne  pollutants
can produce  human health effects (Chapter  11).   Few such studies have been
conducted  with laboratory  animals,  because testing  and measuring of such
mixtures is not only complicated, but  extremely costly.  In these studies, the
investigators  attempted  to simulate the photochemical reaction products pro-
duced  under  natural  conditions and to define  the cause-effect relationship.
     Exposure  to complex  mixtures  of  oxidants plus the various  components
found  in  UV-irradiated auto exhaust indicates  that  certain effects,  such  as
histopathological changes,  increase in susceptibility to infection,  a variety
of  altered pulmonary  functional  activities were  observed in this oxidant
atmosphere which  was not  reported in the nonirradiated exhaust (Murphy et  al.,
                                    9-255

-------
1963; Murphy,  1964; Nakajima et  a"L,  1972;  Hueter et al., 1966).   Certain
other biological responses were observed in both treatment groups, including a
decrease in  spontaneous  activity,  a  decrease  in  infant  survival  rate,  fertil-
ity, and  certain pulmonary  functional  abnormalities (Hueter et al.,  1966;
Boche and Quilligan, 1960; Lewis et al., 1967).
     Dogs exposed  to UV-irradiated  auto exhaust containing oxidants either
with or without SO  showed significant pulmonary functional abnormalities that
                  f{
had  relatively  good correlation with structural changes  (Hyde et  al.,  1978;
Gillespie, 1980; Lewis  et al.,  1974).  There were no significant differences
in the magnitude of the response  in  these  two treatment groups, indicating
that oxidant  gases  and  SO  did not  interact  in any synergistic or additive
                          A
manner.
                                   9-256

-------
9,7  REFERENCES


Abraham, W.  M.;  Januszkiewicz, A. J.;  Mingle,  M.; Welker,  M,;  Wanner,  A.;
     Sackner, M.  A.   (1980)  Sensitivity of bronchoprovocation  and tracheal
     mucous velocity  in  detecting airway responses to 03.  J. Appl.  Physiol.:
     Respir. Environ. Exercise Physiol. 48: 789-793.

Abraham, W. M.; Lauredo, I.; Sielczak, M.; Yerger, L.; King, M.  M.;  Ratzan, K.
     (1982) Enhancement of bacterial pneumonia  in sheep  by  ozone exposure.  Am.
     Rev. Respir. Dis. Suppl. 125(4 pt. 2): 148.

Abraham, W.; Chapman, G. A.; Marchette, B. (1984a) Differences  between  inhaled
     and intravenous  carbachol in detecting Q3-induced airway effects.  Environ.
     Res. 35: 430-438.

Abraham, W.  M.;  Delehunt,  J. C.; Yerger,  L.;  Marchette, B.; Oliver, W., Jr.
     (1984b)  Changes  in  airway permeability and responsiveness after exposure
     to ozone. Environ. Res. 34: 110-119.

Aharonson,  E. F. ; Menkes, H.; Gurtner, G,; Swift, D.  L.;  Proctor,  D.  F.  (1974)
     Effect of  respiratory  airflow rate on  removal  of soluble vapors by the
     nose. J. Appl. Physiol. 37: 654-657.

Altman,  P.  L.;  Dittmer,  D.  S.,  eds.  (1971)  Lung volumes and other ventilatory
     variables during pregnancy: man.  In: Respiration and circulation.  Bethesda,
     MD: Federation of American  Societies for Experimental  Biology;  pp.  61-63.

Alpert,  S.  M.;  Lewis, T. R.  (1971)  Ozone tolerance  studies utilizing uni-
     lateral  lung exposure. J. Appl. Physiol. 31: 243-246.

Alpert,  S.  M.;  Schwartz, B. B.;  Lee,  S.  D.; Lewis,  T.  R.  (1971a) Alveolar-
     protein  accumulation:  a sensitive indicator of  low level  oxidant toxi-
     city.  Arch. Intern. Med. 128: 69-73.

Alpert,  S.  M. ;  Gardner, D.  E. ;  Hurst, D. J. ;  Lewis, T.  R.; Coffin, D.  L.
     (1971b)  Effects  of exposure to ozone on defensive mechanisms of the lung.
     J. Appl. Physiol. 31: 247-252.

Amdur, M. 0.; Ugro, V.; Underbill, D.  W.  (1978) Respiratory response of guinea
     pigs  to  ozone alone and with sulfur dioxide. Am. Ind.  Hyg. Assoc.  J.  39:
     958-961.

American  Thoracic  Society.  (1962) Chronic bronchitis,  asthma,  and pulmonary
     emphysema.  Am.  Rev. Respir. Dis.  85: 762-768.

Ames, B. N.;  McCann,  J.; Yamasaki, E.  (1975) Methods  for detecting carcinogens
     and  mutagens  with the Salmonella/mammalian-microsome  mutagenicity test.
     Mutat.  Res. 31:  347-364.

Amoruso,  M.  A.; Witz, G.; Goldstein,  B.  D.  (1981) Decreased superoxide anion
     radical  production  by rat  alveolar  macrophages  following  inhalation  of
     ozone  or nitrogen dioxide.  Life  Sci. 28:  2215-2221.


                                   9-257

-------
Aranyi, C.;  Fenters,  J.;  Ehrlich, R.;  Gardner,  D.  (1976) Scanning  electron
     microscopy  of alveolar macrophages  after exposure to oxygen,  nitrogen
     dioxide, and  ozone. EHP Environ, Health Perspect.  16: 180.

Aranyi, C.; Vana,  S. C.; Thomas,  P. T.; Bradof, J. N.;  Fenters, J. D.;  Graham,
     J. A.;  Miller,  F. J.  (1983) Effects of subchronic exposure  to  a mixture
     of 03,  S02s  and  (NH4)2S04 on host  defenses of mice. J.  Toxicol. Environ.
     Health 12: 55-71.

Astarita, G. (1967) Mass transfer with  chemical reaction.  New York,  NY:  Elsevier
     Publishing Co.; p. 187.

Atwal, 0.  S.;  Pemsingh,  R. S.  (1981) Morphology of microvascular changes  and
     endothelial  regeneration  in experimental  ozone-induced parathyroiditis.
     III. Some pathologic considerations. Am.  J. Pathol.  102: 297-307.

Atwal, 0. S.; Pemsingh, R. S.  (1984)  Occurrence of mallory body-like inclusions
     in parathyroid chief cells  of ozone-treated dogs.  J.  Pathol.  142:  169-174.

Atwal, 0.  S.; Wilson, T.  (1974) Parathyroid  gland changes  following ozone
     inhalation: a morphologic study. Arch. Environ.  Health 28: 91-100.

Atwal, 0.  S.;  Samagh, B.  S.;  Bhatnagar,  M.  K. (1975)  A possible autoimmune
     parathyroiditis following ozone  inhalation. II.  A  histopathologic,  ultra-
     structural, and immunofluorescent  study.  Am. J.  Pathol. 80:  53-68.

Bailey, P.S. (1978) Ozonization  in organic chemistry.   V.  1, olefinic compounds.
     New York: Academic Press.

Ballew, M.; Calabrese, E. J.;  Moore,  G, S. (1983) The effect of dietary vitamin C
     on ozone-induced  oxidative  changes in guinea pig erythrocytes.  J.  Environ.
     Sci. Health A18:  597-610.

Barr, B.  C.  (1984) Terminal airway remodeling  in chronic ozone exposure. Diss.
     Abstr.  Int. 44: 3724.

Barry, B.  E.;  Miller, F.  J.;  Crapo,  J. D.  (1983)  Alveolar epithelial injury
     caused  by inhalation of  0.25  ppm of ozone.  In: Lee, S.  D.; Mustafa,
     M. G.;  Mehlman,  M.  A., eds. International  symposium on the biomedical
     effects of ozone  and related photochemical oxidants;  March 1982; Pinehurst,
     NC.   Princeton, NJ:  Princeton Scientific  Publishers, Inc; pp.  299-309.
     (Advances in  modern environmental  toxicology: v. 5).

Bartlett, D., Jr.; Faulkner, C.  S. II;  Cook, K. (1974)  Effect of  chronic ozone
     exposure  on  lung elasticity in  young rats. J. Appl.  Physio!. 37:  92-96.

Bergers, W.  W.  A.; Gerbrandy,  J. L.  F.; Stap,  J.  G.  M.  M.; Dura,  E.  A.  (1983)
     Influence  of air polluting  components viz ozone and the open air  factor
     on host-resistance towards  respiratory infection.  In: Lee, S. D.;  Mustafa,
     M. G.; Mehlman, M. A., eds.  International  symposium on the biomedical  ef-
     fects of ozone and related  photochemical  oxidants; March 1982;  Pinehurst,
     NC.   Princeton,  NJ:  Princeton Scientific  Publishers,  Inc.; pp.  459-467.
     (Advances in  modern environmental  toxicology: v. 5).
                                    9-258

-------
Berliner, J, A.;  Kuda,  A.; Mustafa, M.  G.;  Tierney, T. F.  (1978) Pulmonary
     morphologic studies  of  ozone tolerance in the  rat (a possible mechanism
     for tolerance). Scanning Electron Microsc. 2: 879-884.

Bhatnagar, R. S.; Hussain, M. Z.; Sorensen, K. R.; Mustafa,  M. G.; von  Dohlen,
     F. M.;  Lee,  S.  D.  (1983) Effect of ozone on  lung collagen  biosynthesis.
     In:  Lee,  S.  D. ;  Mustafa,   M.  G.;  Mehlman,  M.  A., eds.  International
     symposium  on the  biomedical effects of ozone and related photochemical
     oxidants;  March  1982;   Pinehurst,  NC.   Princeton,  NJ:  Princeton
     Scientific Publishers,  Inc.; pp.  311-321.  (Advances in modern  environ-
     mental toxicology: v. 5).

Bils,  R.  F.  (1970)  Ultrastructural alterations  of alveolar tissue of  mice.
     III. Ozone. Arch. Environ.  Health 20: 468-480.

Bloch, W. N. ,  Jr.;  Miller,  F. J.;  Lewis, T. R.  (1971)  Pulmonary hypertension
     in dogs  exposed to ozone.   Cincinnati, OH: U.S.  Environmental Protection
     Agency, Air Pollution Control  Office.

Bloch, W. N.,  Jr.;  Lewis, T. R.; Busch,  K.  A.;  Orthoefer,  J. G.; Stara, J. F.
     (1972) Cardiovascular status of female beagles  exposed  to air pollutants.
     Arch. Environ.  Health 24: 342-353.

Bloch,  W.  N.,  Jr.;  Lassiter,  S.; Stara, J. F.;  Lewis,  T.   R. (1973)  Blood
     rheology  of  dogs  chronically  exposed to  air pollutants. Toxicol. Appl.
     Pharmacol. 25:  576-581.

Boatman,  E.  S.; Sato,  S. ;  Frank, R.  (1974) Acute effects  of ozone on cat
     lungs.  II. Structural.  Am.  Rev, Respir. Dis.  110:  157-169.

Boatman,  E.  S.; Ward,  G.; Martin, C.  J.  (1983) Morphometric changes in rabbit
     lungs  before and after pneumonectomy and exposure to  ozone.  J.  Appl.
     Physio!.:  Respir. Environ.  Exercise Physio!,  54:  778-784.

Boche,  R.  D.;  Quilligan,  J.  J., Jr. (1960) Effects of synthetic smog on spon-
     taneous activity of  mice.  Science  (Washington,  DC) 131: 1733-1734.

Boorman,  G.  A.; Schwartz, L. W.; McQuillen,  N.  K. ; Brummer,  M.  E.  G.   (1977)
     Pulmonary  response  following  long-term intermittent  exposure  to  ozone:
     structural  and morphometric changes.  Am.  Rev.  Respir.  Dis.  115: 201.

Boorman,  G.  A.; Schwartz, L. W.; Dungworth, D.  L. (1980) Pulmonary effects of
     prolonged  ozone insult in  rats: morphometric evaluation of the central
     acinus.  Lab. Invest.  43: 108-115.

Boushey,  H.  A.;  Holtzman, M. J.; Sheller,  J.  R.;  Nadel, J.  A. (1980) Bronchial
      hyperreactivity.  Am.  Rev.  Respir.  Dis.  121:  389-413.

Bowes, C.;  Cumming, G.;  Horsfield,  K.;  Loughhead, J.; Preston,  S.  (1982) Gas
     mixing  in  a model  of the pulmonary acinus  with asymmetrical alveolar ducts.
     J.  Appl.  Physio!.:  Respir.  Environ.  Exercise Physiol.  52: 624-633.
                                    9-259

-------
Bradley, M.  0.;  Erickson, L. C. (1981) Comparison of the effects  of  hydrogen
     peroxide  and X-ray irradiation on toxicity,  mutation, and DNA  damage/
     repair  in mammalian cells (V-79). Biochim. Biophys.  Acta 654:  135-141.

Bradley, M.  0.;  Hsu,  I.  C.;  Harris,  C.  C.  (1979) Relationships between sister
     chromatid exchange and mutagenicity,  toxicity,  and DNA damage.  Nature
     (London) 282: 318-320.

Brain, J.  D.  (1980)  Macrophage damage  in  relation  to the pathogenesis of lung
     diseases. EHP Environ. Health Perspect. 35: 21-28.

Brain, J.  D.;  Sorokin, S. P.; Godleski, J, J.  (1977) Quantification,  origin,
     and fate of pulmonary macrophages.  In:  Brain,  J.  D.;  Proctor,  D.  F.;
     Reid, C.  M., eds. Respiratory  defense mechanisms.  New York,  NY: Marcel
     Dekker,  Inc.; pp. 849-891 (Lung biology  in health  and disease:  v.  5).

Brain, J.  D.;  Golde, D. W.; Green,  G.  M.; Massaro, D.  J., Valberg, P.  A.;
     Ward, P. A.; Werb, Z. (1978) Biological potential of pulmonary macrophages.
     Am. Rev. Respir.  Dis. 118: 435-443.

Brinkman,  R.;  Lamberts,  H. B.;  Veninga, T.  S.  (1964) Radiomimetic  toxicity of
     ozonised air. Lancet (7325): 133-136.

Brummer, M.  E. G.; Schwartz, L. W.; McQuillen, N.  K. (1977) A  quantitative
     study of  lung damage by scanning electron microscopy:  inflammatory cell
     response  to  high-ambient  levels of ozone.  Scanning Electron  Microsc. 2:
     513-518.

Byers, D.  H.;  Saltzman,  B. E.  (1959) Determination  of ozone in air by neutral
     and alkaline iodide procedures.  In:  Ozone chemistry  and  technology:
     proceedings  of  the first international ozone conference; November  1956;
     Chicago,  IL. Washington,  DC:   American Chemical  Society;  pp.  93-101.
     (Advances in chemistry series:  v. 21).

Calabrese, E. J.; Moore, G. S. (1980) Does the  rodent model  adequately predict
     the effects  of ozone induced changes to human erythrocytes? Med.  Hypothe-
     ses 6: 505-507.

Calabrese, E. J.; Moore, G.; Brown,  R. (1979)  Effects of environmental oxidant
     stressors on individuals with a G-6-PD deficiency with particular reference
     to an animal model. EHP Environ. Health Perspect. 29:  49-55.

Calabrese, E.  J.; Moore, G.  S.; Williams,  P.  (1982) Effect of methyl oleate
     ozonide,  a  possible ozone intermediate,  on normal  and G-6-PD deficient
     erythrocytes. Bull. Environ. Contam. Toxicol. 29: 498-504.

Calabrese, E. J.; Moore, G. S.; Grunwald, E. L.  (1983a) Ozone-induced decrease
     of erythrocyte survival in adult rabbits.  In: Lee, S.  D.; Mustafa,  M. G.;
     Mehlman, M.   A., eds. International symposium on the biomedical  effects of
     ozone and related  photochemical  oxidants; March 1982;  Pinehurst,  NC.
     Princeton,  NJ:  Princeton  Scientific  Publishers, Inc.;  pp.  103-117.  (Ad-
     vances in modern  environmental  toxicology:  v. 5).
                                   9-260

-------
Calabrese, E. J.;  Moore,  G.  S. ; Williams, P. S. (19835) An evaluation of the
     Dorset  sheep  as  a predictive animal  model  for the response  of G-6-PD
     deficient human  erythrocytes  to  a proposed systemic toxic ozone interme-
     diate,  methyl  oleate hydroperoxide.  Vet.   Hum.  Toxicol.  25: 241-246.

Calabrese, E. J. ;  Moore,  G.  S.;  Weeks, B.  L.; Stoddard,  A.  (1983c) The effect
     of ozone exposure upon  hepatic  and  serum  ascorbic  acid  levels  in male
     Sprague-Dawley rats. J.  Environ.  Sci. Health A18: 79-93.

Calabrese, E. J. ;  Williams,  P.  S. ; Moore, G. S. (1983d) An evaluation of the
     Dorset  sheep  as  a predictive  animal  model  for  the response of glucose-6-
     phosphate dehydrogenase-deficient human erythrocytes to a proposed systemic
     toxic ozone intermediate,  methyl oleate  ozonide. Ecotoxicol.  Environ.
     Safety 7: 416-422.

Campbell,  K.  I.; Hilsenroth,  R. H.  (1976) Impaired resistance  to toxin  in
     toxoid-immunized  mice exposed to ozone or  nitrogen  dioxide. Clin. Toxicol.
     9: 943-954.

Campbell,  K.  I.; Clarke, G.  L. ; Emik, L.  0.;  Plata,  R.  L.  (1967)  The atmos-
     pheric contaminant peroxyacetyl  nitrate. Arch. Environ. Health  15: 739-744.

Campbell,  K.  I.; Emik, L. 0.;  Clarke, G.  L.;  Plata,  R.  L.  (1970)  Inhalation
     toxicity of peroxyacetyl nitrate: depression  of voluntary activity in
     mice. Arch. Environ. Health 20:  22-27.

Carp, H.;  Janoff, A.  (1980)  Inactivation  of bronchial  mucous proteinase inhibi-
     tor  by  cigarette smoke and phagocyte-derived  oxidants.   Exp.  Lung Res.
     1: 225-237.

Case, G. D.; Dixon, J. S.; Schooley,  J. C. (1979) Interactions of  blood metal-
     loproteins  with  nitrogen  oxides  and  oxidant air pollutants. Environ. Res.
     20: 43-65.

Castleman, W. L.;  Dungworth, D.  L.; Tyler, W. S. (1973a)  Cytochemically detec-
     ted  alterations   of  lung  acid phosphatase  reactivity following ozone
     exposure. Lab. Invest.  29:  310-319.

Castleman, W.  L.;  Dungworth,  D.  L. ; Tyler,  W. S.  (1973b) Histochemically
     detected enzymatic  alterations  in rat  lung exposed to ozone. Exp.  Mol.
     Pathol. 19: 402-421.

Castleman, W.  L.;  Dungworth, D.  L.; Tyler, W.  S. (1975) Intrapulmonary airway
     morphology  in three species of monkeys:  A correlated scanning and trans-
     mission electron microscopic  study.  Am. J.  Anat.  142:  107-121.

Castleman, W. L.;  Tyler,  W.  S.;  Dungworth, D. L.  (1977)  Lesions  in respiratory
     bronchioles and  conducting airways  of  monkeys exposed to ambient levels
     of ozone. Exp. Mol.  Pathol. 26:  384-400.

Castleman, W. L.;  Dungworth,  D.  L.;  Schwartz, L. W.;  Tyler, W. S.  (1980)  Acute
     respiratory bronchiolitis:  an ultrastructural  and autoradiographic study
     of epithelial  cell  injury and renewal in rhesus  monkeys  exposed to  ozone.
     Am.  J.  Pathol. 98:  811-840.


                                    9-261

-------
Cavender, F.  L.;  Steinhagen,  W.  H.;  Ulrich,  C.  E.;  Busey,  W.  M.;  Cockrell,  B.
     Y.;  Haseman,  J.  K.;  Hogan,  M.  D.;  Drew,  R.  T.  (1977)  Effects in rats  and
     guinea pigs  of  short-term exposures to  sulfuric  acid mist,  ozone, and
     their combination. J. Toxicol.  Environ.  Health 3: 521-533.

Cavender, F.  L.;  Singh, B.;  Cockrell,  B.  Y.  (1978)  Effects in rats  and guinea
     pigs of  six-month exposures  to sulfuric  acid  mist,  ozone,  and  their
     combination. J.  Toxicol. Environ. Health 4: 845-852.

Chan, P.  C.; Kindya, R. J.; Kesner,  L. (1977)+Stud.ies  on the  mechanism of ozone
     inactivation of erythrocyte membrane (Na +  K )-activated ATPase.  J. Biol.
     Chem.  252: 8537-8541.

Chaney,  S.  G.  (1981) Effects of ozone on leukocyte DNA.  Research  Triangle
     Park,  NC:  U.S.  Environmental  Protection Agency,  Health  Effects Research
     Laboratory;  EPA  report no.  EPA-600-81-031.  Available  from:   NTIS,
     Springfield, VA; PB81-179277.

Chang, H.-K.; Farhi, L, E. (1973) On mathematical analysis of gas transport in the
     lung.  Respir. Physio!. 18: 37Q-38ci.

Chiappino,  G.;  Vigliani,  E.  C.   (1982)  Role  of  infective,  immunological,  and
     chronic  irritative factors  in the development of silicosis. Br.  J. Ind.
     Med. 39: 253-258.

Chow,  C.  K.  (1976) Biochemical  responses in lungs of ozone-tolerant  rats.
     Nature (London) 260:  721-722.

Chow, C. K.  (1983) Influence of dietary vitamin E on  susceptibility to ozone
     exposure.  In: Lee, S.  D., Mustafa, M. G.;  Mehlman, M. A., eds. Interna-
     tional  symposium  on  the biomedical effects of  ozone  and related  photo-
     chemical oxidants; March 1982; Pinehurst,   NC.  Princeton,  NJ:  Princeton
     Scientific Publishers, Inc.; pp. 75-93.  (Advances in modern  environmental
     toxicology: v. 5).

Chow, C.  K.  (1984) Glutathione peroxidase system in the lungs of  ozone-tolerant
     and non-tolerant  rats.  In:  Bors, W.; Saran,  M.;  Tait,   D.,  eds.  Oxygen
     radicals in chemistry and biology; proceedings of the third  international
     conference; July  1983;  Neuherberg, West Germany. Berlin,  West Germany:
     Walter de Gruyter & Co.; pp. 707-712.

Chow, C. K.;  Kaneko,  J.  J.  (1979)  Influence  of dietary  vitamin E on the  red
     cells of ozone-exposed rats. Environ. Res.   19: 49-55.

Chow, C. K.;  Tappel,  A.  L.  (1972)  An enzymatic protective mechanism against
     lipid peroxidation  damage  to  lungs of  ozone-exposed rats.  Lipids 7:
     518-524,

Chow, C. K.;  Tappel,  A.  L. (1973) Activities of pentose shunt and  glycolytic
     enzymes in lungs of ozone-exposed rats.  Arch. Environ. Health  26:  205-208.

Chow,  C.  K.; Dillard,  C.  J.; Tappel, A.  L.  (1974) Glutathione  peroxidase
     system and lysozyme in rats exposed to ozone or nitrogen dioxide.  Environ.
     Res. 7: 311-319.


                                   9-262

-------
Chow, C. K,; Mustafa, M. G.; Cross, C. E.; Tarkington,-B.  K.  (1975)  Effects of
     ozone exposure  on  the lungs and  the erythrocytes of rats  and  monkeys:
     relative  biochemical  changes.  Environ.  Physio!.   Biochem.  5:  142-146.

Chow, C.  K.;  Aufderheide,  W. M.; Kaneko, J.  J.;  Cross, C. E.; Larkin, E. C.
     (1976a) Effect of dietary vitamin E  on the red  blood  cells  of ozone-exposed
     rats. Fed. Proc. Fed. Am. Soc. Exp.  Biol. 35: 741.

Chow, C.  K.;  Hussain, M. Z.; Cross,  C.  E.;  Dungworth, D.  L.; Mustafa, M. G.
     (1976b) Effect of low  levels of ozone on  rat  lungs. I. Biochemical  respon-
     ses during recovery and reexposure.  Exp.  Mol. Pathol. 25:  182-188.

Chow, C. K.; Cross, C. E.;  Kaneko, J. J.  (1977) Lactate dehydrogenase activity
     and isoenzyme pattern  in lungs, erythrocytes, and plasma of ozone-exposed
     rats and monkeys. J. Toxicol. Environ. Health 3:  877-884.

Chow, C.  K.;  Plopper, C. G.; Dungworth,  D.  L. (1979)  Influence of dietary
     vitamin  E  on the lungs of  ozone-exposed rats:  a correlated biochemical
     and histological study. Environ.  Res. 20: 309-317.

Chow, C. K.; Plopper, C. G.; Chiu, M.; Dungworth,  D. L.  (1981)  Dietary vitamin E
     and  pulmonary  biochemical  and morphological  alterations of rats exposed
     to 0.1 ppm ozone. Environ.  Res. 24:  315-324.

Christensen,  E.;  Giese,  A.  C.  (1954)  Changes in absorption spectra of nucleic
     acids and  their derivatives following exposure to ozone and ultraviolet
     radiation. Arch. Biochem. Biophys.  51: 208-216.

Christman, C.  A.; Schwartz, L.   W.  (1982) Enhanced phagocytosis  by  alveolar
     maerophages  induced by short-term ozone  insult. Environ. Res.  28:  241-250.

Clark,  R.  A.;  Klebanoff, S. J.  (1975) Neutrophil-mediated tumor cell  cyto-
     toxieity:  role  of the peroxidase system.  J.  Exp. Med.  141:  1442-1447.

Clark,  K.  W.;  Posin, C. I. ;  Buckley,  R. D.  (1978)  Biochemical  response of
     squirrel monkeys to ozone.  J. Toxicol.  Environ. Health 4:  741-753.

Claxton,  L.  D.;  Barnes,  H. M.  (1981) The mutagenicity of  diesel-exhaust
     particle  extracts  collected  under  smog-chamber  conditions using  the
     Salmonella typhimurium test system.  Mutat.  Res. 88:  255-272.

demons,,  G.  K.;  Garcia, J.  F.  (1980a) Endocrine  aspects of  ozone exposure in
     rats,,... Arch,, „. Jqxico.L... Suppl..... (4);.. 301-304..,.

demons,  G. K.; Garcia,  J.  F. (1980b)  Changes in  thyroid function after short-
     term ozone  exposure in rats. J,  Environ. Pathol. Toxicol.  4:  359-369.

demons,  G. K.; Wei,  D.  (1984)  Effect  of short-term  ozone  exposure on exogenous
     thyroxine  levels in thyroidectomized and hypophysectomized rats. Toxicol.
     Appl. Pharmacol. 74:  86-90.
                                    9-263

-------
Coffin, D.  L.;  Blommer,  E. J.  (1970)  Alteration of the pathogenic  role of
     streptococci group C  in mice conferred by previous exposure  to  ozone.  In:
     Silver, I.  H.,  ed.  Aerobiology: proceedings  of the third international
     symposium; September  1969; Sussex, United Kingdom. New York, NY: Academic
     Press, Inc.; pp. 54-61,

Coffin, D.  L.;  Gardner,  D. E.  (1972a) The role  of tolerance  in protection  of
     the  lung  against secondary  insults,  In:  MEDICHEM:  first international
     symposium  of occupational physicians of  the chemical industry; April;
     Ludwigshafen,  West  Germany.  Ludwigshafen,  West Germany:   Badische
     Anilin-und Soda-Fabrik (BASF);  pp. 344-364.

Coffin, D.  L.;  Gardner,  D. E.  (1972b)  Interaction of biological agents and
     chemical air pollutants. Ann. Occup. Hyg. 15: 219-235.

Coffin, D.  L.;  Blommer,  E. J.; Gardner,  D.  E.;  Holzman,  R.  (1967)  Effect of
     air pollution on alteration  of  susceptibility to  pulmonary infection.  In:
     Proceedings  of  the  third annual conference on  atmospheric contamination
     in confined spaces; May; Dayton, OH. Wright-Patterson Air Force Base,  OH:
     Aerospace  Medical  Research  Laboratories;   pp.   71-80;  report no.
     AMRL-TR-67-200. Available  from: NTIS, Springfield, VA; AD-835008.

Coffin, D.  L.;  Gardner,  D. E.;  Holzman,  R.  S.;  Wolock, F.  J.  (1968)  Influence
     of ozone on pulmonary cells. Arch. Environ. Health 16: 633-636.

Cortesi,  R.;  Privett,  O.S. (1972) Toxicity  of  fatty ozonides and peroxides.
     Lipids 7: 715-721.

Cosio, M.  G.;  Hale,  K. A.; Niewoehner,  D.  E.  (1980)  Morphologic  and  morphome-
     tric  effects  of prolonged cigarette smoking  on the small airways.  Am.
     Rev.  Respir..Dis. 122: 265-271.

Costa, D.  L.;  Kutzman, R. S.;  Lehmann,  J.  R.;  Popenoe, E. A.;  Drew, R. T.
     (1983)  A  subchronic  multi-dose ozone study  in  rats.  In:  Lee,   S.  D.;
     Mustafa,  M.  G.;  Mehlman,  M. A,, eds.  International  symposium  on  the
     biomedical  effects  of ozone and  related  photochemical  oxidants; March
     1982;  Pinehurst,  NC.  Princeton,  NJ:  Princeton  Scientific  Publishers,
     Inc.;  pp.  369-393.  (Advances in modern environmental toxicology: v. 5).

Crapo, J.  D.;  Barry, B.  E.;  Chang,  L.-Y.;  Mercer,  R.  R.  (1984) Alterations in
     lung  structure caused by  inhalation  of oxidants. J.  Toxicol.  Environ.
     Health 13: 301-321.

Cross, C.  E.;  DeLucia, A.  J.;  Reddy,  A.  K.; Hussain, M.  Z.; Chow,   C.  K. ;
     Mustafa, M.  G.  (1976) Ozone interactions with  lung tissue:  biochemical
     approaches. Am. J. Med. 60:  929-935.

Curran, S.  F.; Amoruso,-M. A,;  Goldstein, B. D.; Berg, R. A.  (1984)  Degradation
     of soluble collagen by ozone or hydroxyl radicals. FEBS  Lett. 176:  155-160.

Dawber, T.  R.  (1980) Lipids and  atherosclerotic disease.  In:  The Framingham
     study.  The  epidemiology  of atherosclerotic  disease. Cambridge,  MA:
     Harvard University Press;  pp. 121-141.
                                    9-264

-------
DeLucia, A. J.; Hoque, P. M.; Mustafa, M. G.; Cross, C. E.  (1972) Ozone  inter-
     action with  rodent  lung:  effect on  sulfhydryls and sulfhydryl-containing
     enzyme activities. J.  Lab. Clin. Med. 80: 559-566.

DeLucia, A.  J. ;   Mustafa, M.  G.;  Cross,  C.  E. ;  Plopper,  C. G. ; Dungworth,
     D. L.  ; Tyler, W.  S.  (1975a)  Biochemical and morphological alterations in
     the lung  following  ozone exposure.  In:  Rai,  C.;  Spielman, L. A.,  eds.
     Air:   I.  pollution  control  and clean  energy; 1973; New  Orleans,  LA;
     Detroit,  MI; Philadelphia,  PA;  Vancouver,  British Columbia.  New York,
     NY: American Institute of Chemical  Engineers; pp. 93-100.  (AIChE sympo-
     sium  series: v.  71, no. 147).

DeLucia, A.  J. ;  Mustafa, M. G. ;  Hussain,  M. Z. ;  Cross,  C.  E.  (1975b) Ozone
     interaction  with  rodent  lung.   III.  Oxidation of  reduced  glutathione  and
     formation of mixed  disulfides between protein and nonprotein  sulfhydryls.
     J. Clin. Invest. 55: 794-802.

Dillard, C.  J. ;  Urribarri,  N.;  Reddy, K.;  Fletcher, B. ; Taylor, S. ;  de Lumen,
     B.; Langberg, S.;  Tappel, A. L.  (1972) Increased lysosomal enzymes  in
     lungs of ozone-exposed rats. Arch. Environ.  Health 25:  426-431.

Dixon,  J.  R. ;  Mountain,  J.  T.  (1965) Role of histamine and related substances
     in development of tolerance  to  edemagenic gases.  Toxicol.  Appl.  Pharmacol.
     7: 756-766.

Dohan,  F.  C. ;  Everts, G. S. ; Smith,  R. (1962) Variation  in air pollution  and
     the  incidence of respiratory  disease.   J.  Air Pollut.  Control  Assoc.
     12: 418-436.

Donovan, D. H. ; Menzel,  D.  B.  (1975) Dietary and  environmental  modification  of
     mouse lung lipids.  J.  Am. Oil Chem. Soc.  52:  122A.

Donovan, D.  H. ;  Menzel,  D.  B.  (1978) Mechanisms of lipid peroxidation:  iron
     catalyzed decomposition  of  fatty acid hydroperoxides  as  the  basis of
     hydrocarbon  evolution  i_n  vivo.  Experientia  34:  775-776.

Donovan, D.  H. ;  Williams,  S. J.;  Charles,  J. M. ; Menzel, D. B. (1977) Ozone
     tbxicity: effect of dietary vitamin E  and  polyunsaturated fatty acids.
     Toxicol.  Lett. 1: 135-139.

Dorsey, A. P.;  Morgan,  D.  L. ;  Menzel, D.  B.  (1983) Filterability  of  erythro-
     cytes from  ozone-exposed mice.  In: Abstracts of  the third international
     corigress  on  toxicology;  September;  San Diego,  CA.  Toxicol.  Lett.
     18(suppl. 1): 146.

Doty,  R. L.  (1975) Determination  of  odour  preferences  in  rodents:  a methodologi-
     cal  review.  In:  Moulton, D. G. ;  Turk,  A.;  Johnson, J. W., eds.  Methods
     in olfactory research. New York,  NY:  Academic Press; pp.  395-406.

Dowel!, A. R. ; Lohrbauer, L.  A.;  Hurst,  D.; Lee,  S. D. (1970)  Rabbit alveolar
     macrophage  damage  caused  by iji vivo ozone  inhalation.  Arch.  Environ.
     Health  21:  121-127.
                                    9-265

-------
Downey, J. E.; Irving, D. H.; Tappel, A.  L.  (1978)  Effects  of dietary antioxi-
     dants on  in  vivo lipid peroxidation in the rat as  measured  by pentane
     production. Lipids 13: 403-407.

Dubeau, H.;  Chung,  Y. S. (1979)  Ozone  response in wild  type and  radiation-
     sensitive  mutants of  Saccharomyces  cerevisiae. Mol.  Gen.  Genet. 176:
     393-398,

Dubeau, H.;  Chung,  Y. S. (1982)  Genetic  effects of ozone:  induction of point
     mutation  and  genetic recombination  in  Saccharomyces cerevisiae. Mutat.
     Res.  102: 249-259.

Dubick, M. A.; Critchfield, J. W.;  Last, J. A.;  Cross, L.  E.; Rucker, R. B.
     (1983) Ascorbic  acid turnover in the mouse following acute  ozone exposure.
     Toxicology (Ireland) 27: 301-313.

Dumelin,  E. E.; Dillard,  C. J.; Tappel, A.  L.  (1978a) Breath  ethane and pentane
     as measures  of vitamin E protection of Macaca radiata against 90 days of
     exposure  to ozone. Environ.  Res. 15:  38-43.

Dumelin,  E.  E.;  Dillard,  C.  J.;  Tappel, A. L.  (1978b) Effect of vitamin E and
     ozone on  pentane and ethane  expired by rats.  Arch.  Environ.  Health 33:
     129-135.

Dungworth, D.  L. (1976) Short-term effects of  ozone on  lungs  of  rats, mice,  and
     monkeys.  EHP Environ. Health Perspect.  16: 179.

Dungworth, D.  L.; Clarke, G.  L.;  Plata, R.  L.  (1969)  Pulmonary lesions produced
     in A-strain  mice by long-term exposure to peroxyacetyl nitrate. Am. Rev.
     Respir. Dis. 99:  565-574.

Dungworth, D.  L.;  Cross, C.  E.;  Gillespie,  J.  R.;  Plopper, C. G.  (1975a) The
     effects of  ozone on animals.  In:  Murphy, J.  S.; Orr,  J. R.,  eds. Ozone
     chemistry and  technology:  a  review  of the literature 1961-1974. Phila-
     delphia,  PA: The Franklin Institute  Press; pp.  29-54.

Dungworth, D.  L.;  Castleman, W.   L.; Chow,  C.  K.; Mellick,  P.  W.;  Mustafa,  M.
     G.;  Tarkington,  B.; Tyler,  W.  S.  (1975b) Effect  of ambient  levels of
     ozone  on  monkeys. Fed.  Proc.  Fed.  Am, Soc.  Exp.  Biol.  34: 1670-1674.

Easton, R.  E.; Murphy, S.  D.  (1967) Experimental  ozone preexposure and hista-
     raine. Arch. Environ. Health  15: 160-166.

Eglite, M.  E.  (1968)  A contribution to the hygienic assessment of atmospheric
     ozone. Hyg. Sanit. 33: 18-23.

Ehrlich,  R.  (1963)  Effect of air pollutants on respiratory infection. Arch.
     Environ.  Health   6r  638-642.

Ehrlich,  R. (1980)  Interaction between  environmental  pollutants  and respiratory
     infections. EHP  Environ. Health Perspect.  35:  89-100.
                                    9-266

-------
Ehrllch, R.  (1983)  Changes in susceptibility to respiratory infection caused
     by exposures to photochemical oxidant pollutants. In:  Lee, S. D.; Mustafa,
     M. G.;  Mehlman,  M.  A., eds.  International  symposium on the biomedical
     effects of ozone and  related photochemical oxidants; March 1982; Pinehurst,
     NC. Princeton,  NJ:  Princeton Scientific Publishers,  Inc.;  pp.  273-285.
     (Advances in modern environmental toxicology: v. 5).

Ehrlich, R.;  Findlay,  J.  C.;  Fenters, J.  D.;  Gardner,  D. .E.  (1977)  Health
     effects  of  short-term inhalation of nitrogen dioxide and ozone mixtures.
     Environ. Res. 14: 223-231.

Ehrlich, R.; Findlay, J. C.; Gardner, D. E. (1979) Effects  of  repeated exposures
     to peak concentrations of nitrogen dioxide  and ozone on resistance  to
     streptococcal pneumonia.  J. Toxicol. Environ. Health 5: 631-642.

Elsayed,  N.  M.;  Mustafa,  M.  G. ;  Postlethwait,  E. M. (1982a)  Age-dependent
     pulmonary response of  rats to ozone exposure. J. Toxicol. Environ.  Health
     9: 835-848.

Elsayed,  N.  M. ;  Hacker, A.;  Mustafa,  M.;  Kuehn, K.  ; Schrauzer,  G.  (19825)
     Effects  of  decreased  glutathione  peroxidase activity on the pentose
     phosphate cycle in mouse  lung.  Biochem.  Biophys.  Res. Commun. 104:
     564-569.

Elsayed,  N.  M. ;  Hacker, A. D.;  Kuehn,  K.;  Mustafa,  M.  G.;  Schrauzer, G.  N.
     (1983)  Dietary antioxidants  and the biochemical  response   to oxidant
     inhalation.   II.  Influence of dietary selenium on the biochemical effects
     of ozone exposure in mouse  lung. Toxicol. Appl. Pharmacol.  71:  398-406.

Emik,  L.  0.; Plata, R. L.  (1969)  Depression of running  activity in  mice  by
     exposure to polluted  air. Arch.  Environ. Health 18:  574-579.

Emik,  L.  0.; Plata, R. L.; Campbell,  K.  I.;  Clarke, G.  L.  (1971) Biological
     effects  of  urban air pollution:  Riverside summary.  Arch.  Environ.  Health
     23: 335-342.

Engel,  L.  A.; Macklem, P.  T.  (1977)  Gas mixing  and  distribution  in the  lung.
     In:  Widdicombe, J. G.,  ed.   Respiratory  physiology II.  Baltimore, MD:
     University  Park Press;  pp.  37-82  (International review  of  physiology:
     v. 14).

Erdman, H.  E.; Hernandez,  T.  (1982) Adult toxicity and dominant  lethals  induced
     by  ozone at specific  stages  in  spermatogenesis in Drosophila virilis.
     Environ. Mutagen.  4:  657-666.

Eustis,  S.  L.;  Schwartz,  L. W.;  Kpsch,  P.  C.;  Dungworth,  D. L. (1981) Chronic
     bronchiolitis  in  nonhuman primates  after prolonged ozone  exposure.  Am. J.
     Pathol.  105: 121-137.

Evans,  M.  J. (1984) Oxidant  gases.  EHP Environ.  Health Perspect. 55: 85-95.

Evans,  M.  J.; Mayr, W.; Bils, R.  F.;  Loosli, C.  G.   (1971) Effects of ozone on
     cell  renewal  in pulmonary alveoli  of  aging mice.  Arch. Environ. Health
     22:  450-453.


                                   9-267

-------
Evans, M. J.; Johnson,  L.  V.;  Stephens,  R.  J.;  Freeman,  G.  (1976a) Renewal  of
     the terminal bronchiolar  epithelium in the rat following exposure to N02
     or 03.  Lab. Invest. 35: 246-257.

Evans, M. J.; Johnson, L. V.; Stephens,  R. J.;  Freeman, G.  (1976b)  Cell  renewal
     in the  lungs  of rats exposed to  low levels of ozone.  Exp.  Mol.  Pathol.
     24: 70-83.

Evans,  M.  J.;   Stephens,  R.  J.; Freeman, G.  (1976c).  Renewal of  pulmonary
     epithelium following  oxidant  injury. In:   Bouhuys, A., ed.  Lung cells in
     disease: proceedings of a Brook Lodge conference; April; Augusta, MI. New
     York, NY:  Elsevier/North-Holland Biomedical  Press; pp. 165-178.

Evans,  M.  J.;   Dekker,  N.  P.;  Cabral-Anderson,  L.  J.;  Shami, S. G. (1985)
     Morphological basis of tolerance to  ozone.  Exp. Mol. Pathol.  42:  366-376.

Fabbri, L. M.;  Aizawa,  H.; Alpert,  S.  E.;  Walters, E. H.; O'Byrne,  P.  M.;
     Gold, B. D.; Nadel, J. A.; Holtzman^ M. J.  (1984) Airway hyperresponsive-
     ness and changes in cell  counts  in bronchoalveolar lavage after ozone
     exposure in dogs. Am. Rev. Respir.  Dis. 129: 288-291.

Fairchild, E. J., II. (1963) Neurohumoral factors in injury from inhaled irri-
     tants.  Arch. Environ. Health 6: 85-92.

Fairchild, E. J., II. (1967) Tolerance mechanisms:  determinants  of lung  responses
     to injurious agents. Arch. Environ.  Health 14: 111-126.

Fairchild, E. J., II; Graham, S. L.  (1963) Thyroid  influence  on  the toxicity  of
     respiratory  irritant  gases, ozone  and nitrogen dioxide. J.  Pharmacol.
     Exp. Ther.  139:  1-77-184.

Fairchild, E. J., II; Murphy, S. D.; Stokinger,  H.  E.  (1959)  Protection  by sulfur
     compounds  against  the air pollutants ozone and nitrogen  dioxide.  Science
     (Washington, DC) 130: 861-862.

Fairchild, E. J., II; Graham, S. L.; Hite, M.;  Killens, R.; Scheel, L. D.  (1964)
     Changes in thyroid I131 activity in ozone-tolerant and ozone-susceptible
     rats. Toxicol. Appl. Pharmacol. 6:  607-613.

Fairchild, G. A.  (1974) Ozone effect  on respiratory deposition of vesicular
     stomatitis virus aerosols. Am.  Rev.  Respir.  Dis.  109:  446-451.

Fairchild, G. A. (1977)  Effects of ozone and sulfur dioxide on  virus growth  in
     mice. Arch. Environ. Health 32: 28-33."'

Fetner, R. H. (1962)  Ozone-induced chromosome breakage in human  cell cultures.
     Nature  (London)  194: 793-794.

Fetner,  R. H. (1963) Mitotic inhibition induced in  grasshopper neuroblasts by
     exposure to  ozone.  Brooks Air  Force Base, TX: USAF School of Aerospace
     Medicine;  document no.  SAM-TOR 63-39.  Available from:  NTIS, Springfield,
     VA; AD-489014.
                                    9-268

-------
Filser, S.  G. ;  Bolt,  H.  M. ;  Muliawan,  H.;  Kappus,  H, (1983) Quantitative
     evaluation of ethane and n-pentane  as indicators  of  lipid peroxidation  HI
     vivo. Arch. Toxicol. 52:  135-147.

Fletcher,  B. L. ;  Tappel,  A.  L.  (1973)  Protective effects of dietary a-Toco-
     pherol  in  rats  exposed  to toxic  levels  of ozone and  nitrogen  dioxide.
     Environ.  Res. 6: 165-175.

Fletcher,  C. M.; Gilson, J. G.; Hugh-Jones, P.; Scadding, J.  G.  (1959)  Termino-
     logy, definitions,  and classification of chronic pulmonary emphysema and
     related conditions: a report of the conclusions  of a CIBA guest symposium.
     Thorax 14:  286-299.

Frager, N.  B. ;  Phalen,  R.  F.  ;  Kenoyer,  J.  L.  (1979)  Adaptations to ozone in
     reference  to mucociliary  clearance.  Arch.  Environ. Health 34: 51-57.

Freeman, B.  A.; Mudd, J. B. (1981) Reaction of  ozone  with sulfhydryls of human
     erythrocytes. Arch. Biochem. Biophys. 208: 212-220.

Freeman,  B. A.; Sharman, M.  C. ; Mudd,  J.  B.  (1979)  Reaction of ozone with
     phospholipid  vesicles and human  erythrocyte ghosts.   Arch.  Biochem.
     Biophys.  197: 264-272.

Freeman,  G.; Crane,  S.  C. ; Furiosi,  N.  J. ;  Stephens, R. J. ; Evans, M. J. ;
     Moore, W.  D. (1972)  Covert reduction  in  ventilatory surface  in  rats
     during prolonged  exposure to  subacute nitrogen dioxide. Am. Rev. Respir.
     Dis.  106:   563-579.

Freeman,  G. ; Stephens,  R.  J.  ;  Coffin,  D.  L.;  Stara,  J.  F.  (1973) Changes in
     dogs'  lungs  after long-term exposure to ozone:  light and electron micro-
     scopy. Arch. Environ. Health 26:  209-216.

Freeman,  G.; Juhos,  L.  T.; Furiosi,  N.  J.;  Mussenden, R.;  Stephens, R.  J. ;
     Evans, M.  J. (1974)  Pathology of  pulmonary disease  from  exposure to
     interdependent  ambient gases (nitrogen dioxide and  ozone).  Arch.  Environ.
     Health 29: 203-210.

Friberg,  L.;  Holma,  B. ;  Rylander,  R.  (1972)  Animal  lung  reactions after
     inhalation of   lead and  ozone.  Environ. Physio!. Biochem.  2:   170-178.

Friedman, M. ;  Gallo, J.  M.;  Nichols,  H.  P.;  Bromberg, P.  A.  (1983)  Changes  in
     inert  gas  rebreathing parameters after  ozone exposure in  dogs. Am.  Rev.
     Respir. Dis. 128: 851-856.

Fujimaki, H.;  Ozawa,..M..;,. Imal, ...T.; Shimizu,  F.  (1984)-Effect-of short-term
     exposure to  Og  on. antibody response  in  mice.  Environ.  Res. 35: 490-496.

Fujinaka,  L.  E. ;  Hyde,  D.  M.; Plopper, C.  G.; Tyler,  W.  S.;  Dungworth, D. L.;
     Lollini,   L.  0.  (1985)  Respiratory bronchiolitis following  long-term
     ozone  exposure  in Bonnet  monkeys:  a  morphometric study. Exp.  Lung Res.
     8: 167-190.

Fukase, 0.;  Isomura, K. ; Watanabe, H.  (1975) Effect of ozone on glutathione
     i_n vivo. Taiki  Osen Kenkyu 10: 58-62.


                                    9-269

-------
Fukase, 0.;  Watanabe, H.;  Isomura, K. (1978)  Effects  of exercise on mice
     exposed to ozone. Arch. Environ.  Health 33: 198-200.

Gardner,  D.  E,  (1981)  Effects of  environmental  chemicals on host defense
     responses of  the lung.  In:  Dean,  J.  H.;  Padarathsingh,  M.,  eds.  Biologi-
     cal  relevance  of immune suppression as  induced  by genetic, therapeutic
     and  environmental  factors:  second in a  series  of technology assessment
     workshops; November  1979;  Williamsburg,  VA. New  York,  NY:  Van Nostrand
     Reinhold Co.; pp. 275-287.

Gardner,  D.  E.  (1982a)  Use  of  experimental  airborne  infections for monitoring
     altered host defenses.  EHP Environ.  Health  Perspect.  43:  99-107.

Gardner,  D.  E.  (1982b)  Effects of gases and airborne particles on lung infec-
     tions.  In: McGrath,  J.  J.; Barnes,  C.  D.  eds.  Air  pollution  -  physio-
     logical effects. New York, NY: Academic  Press, Inc. ;  pp.' 47-79.

Gardner,  D.  E.  (1984) Oxidant-induced enhanced sensitivity to  infection  in
     animal models and their extrapolation to man.  In:   Miller,  F. J.; Menzel,
     D. B.,  ed.   Fundamentals  of extrapolation modeling of inhaled toxicants.
     Washington, DC:  Hemisphere Publ.  Co.; pp.  423-439.

Gardner,  D.  E.; Ehrlich,  R.  (1983) Introduction to host defense sessions.  In:
     Lee,  S.D.; Mustafa,  M.G.;  Mehlman, M.A., eds.  International symposium on
     biomedical effects  of  ozone  and  related photochemical  oxidants; March
     1982;  Pinehurst, NC.  Princeton,  NJ: Princeton  Scientific  Publishers,
     Inc.;  pp.  255-261.  (Advances  in  modern  environmental  toxicology: v.  5).

Gardner,  D.  E.; Graham,  J.  A.  (1977)  Increased pulmonary disease mediated
     through altered  bacterial defenses.  In:  Sanders,  C.  L.;  Schneider, R. P.;
     Dagle,  G.  E.; Ragen,  H.  A.,  eds. Pulmonary macrophage  and epithelial
     cells:  proceedings  of  the  sixteenth annual Hanford  biology  symposium;
     September 1976;  Richland, WA.  Washington,  DC: Energy Research  and Develop-
     ment Administration; pp.  1-21.  (ERDA  symposium series: 43).  Available
     from:  NTIS, Springfield, VA;  CONF-760927.

Gardner,  D.  E.; Pfitzer,  E.  A.; Christian; R. T.; Coffin,  D.  L.  (1971) Loss  of
     protective factor  for alveolar macrophages when  exposed to ozone. Arch.
     Intern. Med.   127: 1078-1084.

Gardner,  D. E.; Lewis, T.  R.;  Alpert, S. M.;  Hurst,  D.  J.; Coffin,  D.  L.
     (1972)  The role of  tolerance  in pulmonary defense  mechanisms.  Arch.
     Environ. Health  25:  432-438.

Gardner,  D.  E.;  Illing,  J.  W.; Miller, F. J.; Coffin, D.  L. (1974) The effect
     of  ozone on  pentobarbital  sleeping time  in mice. Res.  Commun.  Chem.
     Pathol. Pharmacol. 9: 689-700.

Gardner,  D.  E.;  Miller,  F.  J.;  Illing, J.  W.; Kirtz, J.  M. (1977) Increased
     infectivity  with exposure to  ozone  and  sulfuric  acid.  Toxicol,  Lett. 1:
     59-64.

Gelmont,  D.; Stein, R.  A.;  Mead,  J.  F.  (1981)  The  bacterial origin of rat
     breath pentane.  Biochem.  Biophys.  Res. Commun.  102:  932-936.


                                    9-270

-------
Gershwin, L. J.; Osebold, J. W.; Zee, Y. C. (1981) Immunoglobulin  E-containing
     cells  in  mouse lung following  allergen  inhalation and ozone exposure.
     Int. Arch. Allergy Appl.  Immunol. 65: 266-277.

Gertner,  A.;  Bromberger-Barnea,  B. ;  Dannenberg,  A.  M., Jr.; Traystman,  R.;
     Menkes, H.  (1983a)  Responses of the  lung periphery to  1.0 ppm ozone.  J.
     Appl. Physio!. Respir. Environ.  Exercise Physio!.  55: 770-776.

Gertner,  A.;  Bromberger-Barnea,  B.;  Traystman,  R. ;  Berzon, D. ;  Menkes,  H.
     (1983b) Responses of the  lung periphery  to  ozone and histamine.  J.  Appl.
     Physio!.: Respir. Environ. Exercise Physio!.  54: 640-646.

Gertner,  A.;  Bromberger-Barnea,  B.;  Traystman,  R.; Menkes, H.  (1983c) Effects
     of ozone  in peripheral lung  reactivity. J.  Appl. Physio!.: Respir.  Environ.
     Exercise  Physio!. 55:  777-784.

Gertner, A.; Bromberger-Barnea, B.;  Kelly, L.; Traystman,  R.; Menkes,  H.  (1984)
     Local  vagal  responses in the lung periphery. J. Appl.  Physio!.:  Respir.
     Environ.  Exercise Physio!. 57:  1079-1088.

Gillespie,  J.  R.  (1980)  Review of the  cardiovascular and pulmonary function
     studies on  beagles  exposed for 68 months  to auto exhaust and other air
     pollutants   -In:  Stara,  J.  F.  ; Dungworth,  D.  L. ; Orthoefer,  J. C. ;
     Tyler, W. S. ,  eds. Long-term  effects  of air pollutants  in  canine species.
     Cincinnati, OH:  U.S.  Environmental Protection Agency, Environmental  Criteria
     and Assessment Office; pp. 115-154; EPA report  no. EPA  600/8-80-014.  Avail-
     able from: NTIS, Springfield, VA; PB81-144875.

Giri, S.  N.; Benson,  J.;  Siegel,  D.  M.; Rice, S.  A.;  Schiedt, M.  (1975)  Effects
     of pretreatment  with  anti-inflammatory drugs  on  ozone-induced lung  damage  in
     rats.  Proc. Soc. Exp.  Biol.  Med. 150: 810-814.

Giri, S.  N.;  Hollinger,  M.  A.; Schiedt,  M. J.  (1980) The  effects  of ozone and
     paraquat  on PGF2  and  PGE2  levels in  plasma and  combined pleura!  effusion
     and  lung  lavage  of rats.  Environ. Res. 21:  467-476.

Goldstein,  B.  D.  (1973)  Hydrogen  peroxide in erythrocytes.  Detection in  rats
     and  mice  inhaling ozone.  Arch.  Environ. Health  26: 279-280.

Goldstein,  E.  (1984)  An  assessment of animal  models  for testing the effect of
     photochemical  oxidants on pulmonary susceptibility to bacterial  infection.
     J. Toxicol. Environ.  Health  13: 415-421.

Goldstein,  B.  D. ;  Balchum, 0. J.  (1974) Modification  of  the response of  rats
     to  lethal levels of  ozone  by  enzyme-inducing  agents.  Toxicol.  Appl.
     Pharmacol. 27: 330-335.

Goldstein,  B.  D. ;  McDonagh, E. M.  (1975)  Effect  of ozone on  cell membrane
     protein  fluorescence.  I.  Iji  vitro studies  utilizing the red cell  mem-
     brane.  Environ.  Res.  9: 179-186.

Goldstein,  B.  D.;  Pearson,  B.; Lodi,  C.; Buckley,  R.  D.;  Balchum,  0.  J.  (1968)
     The  effect of ozone  on mouse blood  i_n vivo. Arch.  Environ.  Health  16:
     648-650.


                                   9-271

-------
Goldstein, E.;  Tyler,  W.  S.;  Hoeprich,  P.  D.;  Eagle, G.  (1971a) Ozone and the
     antibacterial  defense  mechanisms of the murine lung. Arch.  Intern.  Med.
     127: 1099-1102.

Goldstein, E.; Tyler, W. S.; Hoeprich,  P. D.; Eagle,  C.  (1971b) Adverse  influ-
     ence of  ozone  on  pulmonary bactericidal  activity of murine lungs.  Nature
     (London) 229:  262-263.

Goldstein, B. D.;  Levine,  M.  R. ;  Cuzzi-Spada,  R.;  Cardenas, R.;  Buckley, R.
     D.; Balchum, 0.  J.  (1972a) p-aminobenzoic  acid as  a  protective  agent in
     ozone toxicity. Arch. Environ. Health 24: 243-247.

Goldstein, E.; Eagle, M. C.; Hoeprich,  P. D. (1972b) Influence  of ozone  on pulmo-
     nary defense mechanisms of  silicotic mice.  Arch.  Environ.  Health 24:  444-448.

Goldstein, B. D.; Lai,  L.  Y.;  Cuzzi-Spada,  R.  (1974a) Potentiation of comple-
     ment-dependent membrane damage  by ozone.  Arch.  Environ.  Health  28:
     40-42.

Goldstein, E.;  Warshauer,  D.; Lippert, W.;  Tarkington,  B. (1974b) Ozone and
     nitrogen dioxide exposure.  Arch. Environ. Health 28:  85-90.

Goldstein, B. D.; Solomon, S.; Pasternack, B. S.; Bickers, D. R.  (1975)  Decrease
     in rabbit lung microsomal cytochrome P-450  levels following  ozone exposure.
     Res. Commun. Chem. Pathol.  Pharmacol. 10: 759-762.

Goldstein, B. D.; Hamburger,  S.  J.;  Fal.k,  G.  W.; Amoruso,  M. A.  (1977) Effect
     of ozone and nitrogen dioxide on the agglutination  of rat  alveolar  macro-
     phages by concanavalin A. Life Sci. 21: 1637-1644.

Goldstein,  E.;  Gibson, J.  B.;  Tassan,  C.;  Lippert,  W.  (1978a)  Effect  of
     levamisole on  alveolar macrophage function. Am.  Rev.  Respir. Dis.  117:
     273.

Goldstein, E.;  Bartlema,  H. C.; van  der  Ploeg,  M.;  van Duijn,  P.; van der
     Stap, J.  G.  M.  M.;  Lippert,  W. (1978b)  Effect of ozone on lysosomal
     enzymes  of  alveolar  macrophages engaged  in phagocytosis and killing of
     inhaled Staphylococcus aureus. J.  Infect. Dis.  138: 299-311.

Gooch, P.  C.; Creasia, D. A.; Brewen,  J. G.  (1976) The  cytogenetic effect of
     ozone:  inhalation and  iji  vitro exposures.  Environ.  Res.  12: 188-195.

Gordon, T.;  Amdur,  M.  0.   (1980) Effect of ozone on respiratory  response of
     guinea pigs to histamine. J.  Toxicol. Environ.  Health 6: 185-195.

Gordon, T.;  Taylor, B. F.; Amdur, M. 0.  (1981) Ozone inhibition of tissue
     cholinesterase in guinea pigs. Arch.  Environ.  Health  36: 284-288,

Gordon, T.; Venugopalan, C. S.;  Amdur,  M. 0.;-Drazen,  J. M.  (1984) Ozone-induced
     airway  hyperreactivity in  the  guinea pig.  J.-  Appl.  Physio!.: Respir.
     Environ. Exercise Physio!.  57: 1034-1038.

Gottos A. M.» Jr. (1979) Status  report: plasma lipids, lipoproteins,  and coronary
     artery disease. In: Paoletti, R.;  Gotto, A. M., Jr.,  eds.  Atherosclerosis
     reviews: v. 4. New York,  NY:  Raven Press; pp.  17-28.

                                   9-272

-------
Graham, J. A. (1979) Alteration of hepatic xenobiotic metabolism by ozone  [dis-
     sertation].   Durham,  NC:  Duke  University.   Available from: University
     Microfilms, Ann Arbor, MI: publication no. 80-09646.

Graham, J. A.;  Menzel,  D.  B.; Miller,  F.  J. ;  Illing, J. W. ; Gardner,  D.  E.
     (1981) Influence of ozone on pentobarbital-induced  sleeping time  in mice,
     rats, and hamsters. Toxicol. Appl. Pharmacol. 61: 64-73.

Graham, J. A.;  Menzel,  D.  B.; Miller,  F.  J.;  Illing, J. W.; Gardner,  D.  E.
     (1982a) Effect  of  ozone on  drug-induced  sleeping time in mice  pretreated
     with  mixed-function oxidase  inducers  and inhibitors.  Toxicol.   Appl.
     Pharmacol. 62: 489-497.

Graham, J. A.;  Miller,  F.  J.;  Gardner,  D.  E.;  Ward,  R.;  Menzel,  D.  B.  (1982b)
     Influence  of  ozone and nitrogen dioxide  on  hepatic  microsomal  enzymes in
     mice. J. Toxicol.  Environ. Health 9: 849-856.

Graham, J. A.; Menzel,  D. B.; Miller, F. J.; Illing,  J.  W.; Ward, R.;  Gardner,
     D.  E.  (1983)  Influence  of ozone  on  xenobiotic metabolism. In:  Lee,
     S. D. ; Mustafa,  M. G. ; Mehlman, M. A. ,  eds.  International  symposium on
     the biomedical effects of ozone and related photochemical oxidants; March
     1982; Pinehurst,   NC.   Princeton,  NJ:   Princeton  Scientific  Publishers,
     Inc.; pp.  95-117.  (Advances in modern environmental  toxicology:  v.  5).

Graham,  J. A.; Menzel,  D.   B.;  Mole, M.  L.; Miller,  F.  J.; Gardner,  D. E.
     (1985)  Influence  of  ozone  on pentobarbital pharmacokinetics  in  mice.
     Toxicol.  Lett. 24:  163-170.

Green, G.  M.  (1984)  Similarities of host defense  mechanisms against pulmonary
     disease in  animals and man.  J. Toxicol.  Environ.  Health 13:  471-478.

Gregory, A. R.; Ripperton,  L. A.; Miller, B. (1967) Effect  of neonatal  thymec-
     tomy  on  the  development of ozone tolerance  in  mice. Am.  Ind.  Hyg. Assoc.
     J. 28: 278-282.

Grose, E.  C. ;  Gardner,  D.   E.; Miller, F. J. (1980) Response  of  ciliated epi-
     thelium to ozone and  sulfuric acid. Environ. Res. 22:  377-385.

Grose, E.  C. ;  Richards, J.  H.;  Illing,  J.  W. ; Miller,  F.  J.; Davies,  D.  W. ;
     Graham, J.  A. ;  Gardner, D.  E.  (1982)  Pulmonary host defense responses to
     inhalation of sulfuric acid and ozone. J.  Toxicol. Environ.  Health 10:
     351-362.

Gross, E.  A. ;  Swenberg, J. A.;  Fields,  S. ; Popp, J. A.  (1982)  Comparative
     morphometry  of the nasal, cavity in  rats  and mice.  J.  Anat.  135:  83-88.

Gross,  P.; Scheel, L.  D. ;  Stokinger,  H.  E. (1965) Ozone  toxicity  studies;
     destruction  of alveolar septa—a precursor  of emphysema?   In: Mitchell,
     R.  S. ,  ed.  The pathogenesis of the chronic  obstructive  bronchopulmonary
     disease:  seventh annual conference on research  in  emphysema;  June 1964;
     Aspen,  CO.  Basel,  Switzerland:  S.  Karger,  AG;  pp. 376-381.  (Progress
     in  research  in emphysema  and  chronic   bronchitis:  v.  2;  Medicina
     Thoracalis:  v.  22, nos. 1-3).
                                    9-273

-------
Gross, S.; Melhorn, D. K. (1972) Vitamin E  red cell  lipids  and  red  cell  stabil-
     ity in prematurity. Ann, N. Y. Acad. Sci. 203:  141-162.

Guerrero, R.  R.;  Rounds,  D.  E.;  Olson,  R.  S.;  Hackney, J.  D.  (1979) Mutagenic
     effects  of ozone on human cells exposed  j_n vivo and  iji vitro based  on
     sister chromatid exchange analysis. Environ.  Res. 18:  336-346.

Hadley, J.  G.;  Gardner,  D.  E. ;  Coffin  D.  L.;  Menzel, D. B. (1977) Enhanced
     binding of autologous cells to the macrophage plasma membrane  as  a  sensi-
     tive indicator  of  pollutant damage. In:  Sanders, C.  L.;   Schneider,  R.
     P.; Dagle  G.  E.; Ragan, H. A., eds. Pulmonary  macrophage  and  epithelial
     cells:  proceedings  of the  sixteenth  annual Hanford biology symposium;
     September  1976;  Richland,  WA.  Washington, DC:  Energy  Research and
     Development  Administration;  pp. 1-21.  (ERDA symposium series: v.  43).
     Available from: NTIS, Springfield, VA; CONF-760927.

Halliwell,  B.  (1982)  Production of superoxide, hydrogen peroxide and hydroxyl
     radicals by  phagocytic  cells: a cause  of chronic inflammatory disease?
     Cell Biol. Int. Rep. 6: 529-542.

Hamburger, S.  J.; Goldstein, B. D.  (1979) Effect of  ozone on the agglutination
     of  erythrocytes by  concanavalin  A.   I:  Studies  in  rats. Environ.
     Res. 19:  292-298.

Hamburger, S.  J.; Goldstein, B. D.; Buckley, R.  D.;  Hackney, J. D.; Amoruso,  M.
     A, (1979) Effect of ozone on  the agglutination  of erythrocytes by concana-
     valin  A.  II:  Study of hu'man  subjects  receiving supplemental  vitamin E.
     Environ.  Res. 19: 299-305.

Hamelin, C.;  Chung,  Y.  S. (1974)  Resistance a 1'ozone chez Escherichia  coli.
     I.  Relations avec certains  mecanismes de  reparation de  1'ADN [Ozone
     resistance  in Escherichia  coli.   I.  Relations   with  some  DNA  repair
     mechanisms].  Mol. Gen. Genet.  129: 177-184.

Hamelin, C.; Chung, Y. S. (1975a)  The effect of  low  concentrations  of  ozone  on
     Escherichia coli chromosome.  Mutat. Res.  28:  131-132.

Hamelin,  C.;  Chung,  Y.  S.  (1975b)  Characterization  of mucoid mutants  of
     Escherichia  coli K-12 isolated after  exposure  to ozone.   J. Bacteriol.
     122: 19-24.

Hamelin, C.;  Chung,  Y.  S. (1978)  Role  of the  pol, rec, and DNA gene products
     in the repair of lesions produced  in Escherichia coli  DNA  by ozone.  Stud.
     Biophys.  68:  229-335.

Hamelin, C.;  Sarhan,  F.;  Chung, Y.  S.  (1977a) Ozone-induced DNA degradation
     in  different  DNA polymerase  I mutants of Escherichia  coli K12. Biochem,
     Biophys.  Res. Commun. 77: 220-224.

Hamelin, C.;  Sarhan,  F.;  Chung,  Y. S.  (1977b)  DNA-degradation  caused by ozone
     in  mucoid  mutants of  Escherichia coli  K12.  FEMS Microbiol.   Lett.  2:
     149-151.
                                    9-274

-------
Hammond,  P.  B. ;  Bellies,  R.  P.  (1980)  Metals:  ch. 17.  In:  Doull,  J. ;
     Klaassen, C. D. ;  Amdur,  M. 0. ,  eds.  Toxicology, the  basic  science of
     poisons. New York, NY: Macmillan Press Company,  Inc.;  pp. 409-467.

Hathaway, J.  A.; Terrill,  R.  E.  (1962)  Metabolic effects  of  chronic ozone
     exposure on rats. Am.  Ind. Hyg. Assoc. J. 23: 392-395.

Hattori, K.; Kato, N.; Kinoshita, M.; Kinoshita,  S.;  Sunada, T. (1963) Protec-
     tive effect of ozone  in  mice  against whole-body X-irradiation.  Nature
     (London) 198: 1220.

Hayes,  W.  J., Jr.  (1975)  Toxicology of pesticides.   Baltimore, MD:  William
     and Wilkins Company; pp.  118-127.

Henschler, D.  (1960)  Schutzwirkung einer  Vorbehandlung  mit geringen Gaskon-
     zentrationen  gegen todliche  Reizgas-Lungenodeme [Protective effect  of
     previous treatment  with low gas concentrations  against lethal  pulmonary
     edema  from irritant  gas].  Naunyn Schmiedeberg's Arch.  Exper.  Pathol.
     Pharmakol.  238:  66-67.

Hesterberg,  T.  W.;  Last, J. A. (1981) Ozone-induced  acute  pulmonary  fibrosis
     in  rats:  prevention of increased rates  of collagen synthesis  by methyl-
     prednisolone. Am. Rev.  Respir.  Dis.  123:  47-52.

Hilding,  D.  A.; Hilding, A. C.  (1966)  Ultrastructure of trachea!  cilia and
     cell during regeneration. Ann.  Otol. Rhinol.  Laryngol.  75:  281-294.

Holtzman, M. J.; Fabbri,  L.  M.; Skoogh, B.-E.; O'Byrne,  P.  M.; Walters,  E.  H. ;
     Aizawa, H.; Nadel, J.  A.  (1983a) time course of  airway hyperresponsiveness
     induced  by ozone in dogs. J.  Appl.  Physio!.: Respir.  Environ,  Exercise
     Physio!. 55: 1232-1236.

Holtzman, M.  J.; Fabbri,  L.  M.;  O'Byrne,  P.  M.; Gold,  B.  D.; Aizawa,  H. ;
     Walters,  E.  H.;  Alpert, S. E.; Nadel, J. A.  (1983b) Importance of airway
     inflammation  for hyperresponsiveness  induced by  ozone. Am.  Rev.  Respir.
     Dis. 127:  686-690.

Holzman,  R.  S. ; Gardner, D. E. ; Coffin,  D.  L. (1968)  In vivo  inactivation of
     lysozyme by ozone. J.  Bacteriol. 96:  1562-1566.

Hu, P.  C. ; Miller,  F.  J.;  Daniels,  M. J.;  Hatch,  G.  E.;  Graham, J.  A.; Gardner,
     D.  E. ;  Belgrade, M.  K. (1982)  Protein accumulation in lung  lavage fluid
     following  ozone  exposure.  Environ. Res.  29:  377-388.

Huber,  G.  L. ;  Mason,  R.  J.;  LaForce, M.;  Spencer,  N.  J.; Gardner,  D.  E.;
     Coffin,  D.  L.  (1971)  Alterations  in  the  lung following the administration
     of ozone.  Arch.  Intern. Med.  128: 81-87.

Hueter,  F.  G. ;  Contner, G. L.; Busch, K.  A.; Hinners, R. G. (1966) Biological
     effects  of atmospheres  contaminated  by  auto exhaust. Arch.  Environ.
     Health  12:  553-560.

Hurst,  D.  J.; Coffin, D.  L.  (1971) Ozone  effect on  lysosomal  hydrolases of
     alveolar macrophages  i_n vitro.  Arch.  Intern. Med.  127: 1059-1063.


                                    9-275

-------
Hurst, D.  J.;  Gardner,  D. E.;  Coffin,  D.  L.  (1970)  Effect of ozone on acid
     hydro!ases  of the pulmonary  alveolar macrophage.  J, Reticuloendothel.
     Soc. 8: 288-300.

Hussain, M. Z.; Mustafa, M. G.; Chow, C, K.; Cross, C. E.  (1976a) Ozone-induced
     increase of lung proline  hydroxylase  activity and hydroxyproline content.
     Chest 69(suppl. 2): 273-275.

Hussain, M. Z.; Cross, C. E.;  Mustafa, M.  G.; Bhatnagar,  R. S. (1976b) Hydroxy-
     proline  contents  and prolyl  hydroxylase activities  in  lungs of rats
     exposed to low levels of  ozone. Life  Sci. 18: 897-904.

Hyde, D.; Orthoefer, J. G.; Dungworth, D.; Tyler, W.; Carter, R.; Lum, H.  (1978)
     Morphometric  and morphologic  evaluation  of pulmonary lesions  in  beagle
     dogs  chronically exposed to high ambient levels of  air pollutants.  Lab.
     Invest. 38: 455-469.

Ibrahim, A.  L.;  Zee, Y.  C.; Osebold, J.  W. (1976)  The effects  of ozone on the
     respiratory epithelium and alveolar macrophages of  mice.  I.  Interferon
     production. Proc.  Soc. Exp. Biol.  Med. 152: 483-488.

Ibrahim, A.  L.;  Zee, Y.  C.; Osebold, J.  W. (1980)  The effects  of ozone on the
     respiratory  epithelium of  mice.  II.   Ultrastructural alterations.   J.
     Environ. Pathol. Toxicol. 3: 251-258.

Ikematsu,  T.  (1978) Influence of ozone  on rabbit's tonsils—observation  of
     the surface  of the  tonsil with scanning  electron microscope.  Nippon  Jibi
     Inkoka Gakkai Kaiho 81: 1447-1458.

Illing,  J.  W.;  Miller,  F. J.;  Gardner,  D. E.  (1980)  Decreased resistance to
     infection in  exercised  mice exposed  to N02 and  03.  J. Toxicol. Environ.
     Health 6: 843-851.

Inoue,  H.;  Sato,  S.; Hirose,  T.;  Kikuchi, Y.;  Ubukata,  T.; Nagashima,  S.;
     Sasaki, T.; Takishima,  T. (1979)  A comparative  study between  functional
     and pathologic alterations in  lungs  of  rabbits  exposed to  an ambient
     level  of ozone: functional  aspects.  Nikkyo  Shikkai-Shi  17:  288-296.

Ishidate,  M.,  Jr.,; Yoshikawa,  K.  (1980)  Chromosome aberration  tests  with
     Chinese  hamster cells  jji vitro with  and without metabolic activation—a
     comparative study on mutagens and carcinogens. Arch. Toxicol. Suppl. 4:
     41-44.

Jackson, R. M.; Frank, L. (1984) Ozone-induced tolerance  to hyperoxia in  rats.
     Am. Rev. Respir. Dis. 129: 425-429.

Jegier,  Z.  (1973)  L1ozone en  tant  que  polluant  atmospherique  [Ozone as an air
     pollutant]. Can. J. Public Health 64: 161-166.

Johnson, D.  A.  (1980)  Ozone  inactivation  of  human at-proteinase inhibitor.
     Am. Rev. Respir. Dis. 121: 1031-1038.
                                   9-276

-------
Johnson, B. L.; Orthoefer, J. G.; Lewis, T. R.; Xintaras, C.  (1976)  The  effect
     of ozone on brain  function.  In:  Finkel, A.  J.;  Duel, W.  C., eds.
     Clinical implications of air pollution research: [proceedings of  the]  air
     pollution  medical  research conference;  December 1974;  San Francisco,
     CA. Acton, MA:  Publishing Sciences Group,  Inc.; pp. 233-245.

Johnson, K.  I.;  Fantone,  J.  C., III;  Kaplan, J.; Ward,  P.  A. (1981) In vivo
     damage of  rat  lungs  by oxygen  metabolites.  Clin.  Invest.   67:  983-993.

Juhos,  L.  T.; Evans,  M.  J.; Mussenden-Harvey,  R.;  Furiosi,  N. J.; Lapple,  C.
     E.; Freeman, G. (1978)  Limited  exposure  of rats to H2S04 with and without
     03. J. Environ. Sci. Health C13: 33-47.

Kaplan, J.;  Smaldone,  G.  C.; Menkes, H.  A.;  Swift, D. L.;  Traystman,  R. J.
     (1981)  Response  of collateral  channels  to  histamine:   lack  of vagal
     effect.  J.  Appl.  Physio!.: Respir. Environ.  Exercise Physio!.  51:  1314-
     1319.

Kavlock, R.;  Daston,  G.;  Grabowski,  C.  T.  (1979)  Studies  on the developmental
     toxicity of ozone.  I.  Prenatal  effects.  Toxicol.  Appl. Pharmaeol. 48:
     19-28.

Kavlock, R. J.; Meyer, E.; Grabowski, C. T. (1980)  Studies on the  developmental
     toxicity of ozone: postnatal effects. Toxicol,  Lett. 5:  3-9.

Kawachi,  T.;  Yahagi,  T. ;  Kada, T.;  Tazima,  Y.;  Ishidate,  M. ;  Sasaki,  M.;
     Sugiyama, T. (1980) Cooperative programme on short-term assays  for  carcino-
     genicity in  Japan.  In:  Montesano, R. ;  Bartsch,  H. ;  Tomatis, L.; Davis,
     W., eds.  Molecular  and cellular aspects of  carcinogen screening tests;
     June  1979; Hanover,  West  Germany. Lyon,  France:  International Agency
     for  Research  on  Cancer;  pp.  323-330.  (IARC scientific publications:
     no. 27).

Kenoyer, J.  L.;  Phalen, R. F.;  Davis, J. R.  (1981)  Particle  clearance from  the
     respiratory  tract as a test of toxicity: effect of ozone  on short and
     long  term  clearance.  Exp.  Lung  Respir. 2:  111-120.

Kensler, C. J.;  Battista,  S. D.  (1966)  Chemical and physical  factors affecting
     mammalian  ciliary activity. In:  Symposium on structure,  function and mea-
     surement of  respiratory cilia;  February  1965;  Durham, NC.  Am. Rev.  Respir.
     Dis.  93: 93-102.

Kesner, ^..;  l£indya, R. J.; Chan, P.  C.  (1979) Inhibition  of  erythrocyte membrane
     (Na + K  )-activated ATPase by ozone-treated  phospholipids.  J. Biol. Chem.
     254:  2705-2709.

Kimura, A.;  Goldstein, E.  (1981) Effect of ozone  on concentrations of lysozyme
     in phagocytizing alveolar  macrophages.   J.  Infect.  Dis. 143:  247-251.

Kindya,  R. J.; Chan,  P. C.  (1976)  Effect of  ozone on  erythrocyte  membrane
     adenosine  triphosphatase.  Biochim. Biophys.  Acta  429:  608-615,

Klebanoff, S. J.; Clark,  R.  A.  (1975) Hemolysis and iodination  of erythrocyte
     components by a myeloperoxidase-mediated system.  Blood 45:  699-707.


                                    9-277

-------
Kliment, V. (1972) Geometry of guinea pig respiratory  tract  and  application  of
     Landahl's model  of deposition or aerosol  particles.  J. Hyg.  Epidemic!.
     Microbiol. Immunol. 16: 107-114.

Kliment, V. (1973)  Similarity  and dimensional  analysis, evaluation of aerosol
     deposition in  the  lungs  of  laboratory  animals  and man. Folia  Morphol.
     21: 59-64.

Konigsberg, A. S.;  Bachman, C. H. (1970) Ozonized atmosphere and  gross  motor
     activity of rats.  Int. J. Biometeorol.  14: 261-266.

Koontz, A.  E.; (Jealjh, R.  L. (1979)  Ozone  alteration of transport of cations
     and the  Na  /K -ATPase in  human erythrocytes.  Arch.  Biochem.  Biophys.
     198:  493-500.

Kotin,  P.;  Thomas,  M. (1957) Effect of  air contaminants on reproduction and
     offspring survival  in mice.  AMA Arch.  Ind. Health 16: 411-413.

Kotlikoff,  M.  I.;  Jackson, A.  C.; Watson, J. W.  (1984)  Oscillatory  mechanics
     of the respiratory system  in ozone-exposed rats.  J.  Appl.  Physio!.;
     Respir. Environ. Exercise Physio!.  56:  182-186.

Krupnick,  A. J.;  Frank,  R.  (1984) Synergistic effects  on rat lungs of mixtures
     of oxidant  air  pollutants  (ozone  or nitrogen  dioxide) and respirable
     aerosols [letter to the editor]. Am. Rev.  Respir.  Dis.  129: 885.

Kulle,  T.   J,;  Cooper, G.  P. (1975)  Effects of formaldehyde and ozone on the
     trigeminal nasal  sensory system.  Archi Environ.  Health 30: 237-243.

Kyei-Aboagye, K.; Hazucha,  M.; Wyszogrodski, I.;  Rubinstein, D.; Avery,  M.  E.
     (1973) The effect  of ozone  exposure  in vivo on  the  appearance of  lung
     tissue lipids  in the endobronchial lavage of rabbits.  Biochem.  Biophys.
     Res.  Commun.  54: 907-913.

LaBelle, C. W.; Long, J.  E,; Christpfano,  E. E. (1955) Synergistic effects of
     aerosols: particulates as carriers  of toxic  vapors.  AMA Arch. Ind.  Health
     11: 297-304.

Laqueur, E.;  Magnus, R. (1921)  Experimental and theoretical bases  for  the
     therapy of phosgene sickness.  Z. Gesamte  Exp.  Med.  13: 284-291.

Larkin, E.  C.; Rao, G.  A.  (1984) Reduced  growth  in  rats exposed to  ozone as
     compared to  pair-fed  controls  breathing air. IRCS Med.  Sci. 12: 741-742.

Larkin, E.  C.; Kimzey, S.  L.; Siler, K.  (1978)  Response  of the  rat erythrocyte
     to ozone  exposure.  J.  Appl.  Physio!.:  Respir.  Environ.  Exercise Physiol.
     45: 893-898.

Larkin, E.  C.; Goheen,  S.  C.; Rao,  G.  A.  (1983) Morphology and fatty acid
     composition  of erythrocytes  from monkeys  exposed to ozone  for  one  year.
     Environ.  Res.  32: 445-454.
                                   9-278

-------
Last, J.  A.  (1983) Health effects of indoor air pollution: synergistic effects
     of nitrogen  dioxide  and a respirable aerosol. Environ.  Int.  (U.S.A.)  9:
     319-322.

Last, J.  A.;  Cross,  C.  E. (1978) A  new model  for health effects of air pol-
     lutants:  evidence for synergistic effects of mixtures of ozone  and  sulfuric
     acid aerosols on rat lungs. J.  Lab. Clin. Med. 91:  328-339.

Last, J.  A.;  Greenberg,  D.  B.  (1980)  Ozone-induced alterations in  collagen
     metabolism of rat lungs. II. Long-term exposure. Toxicol.  Appl.  Pharmacol.
     55:  108-114.

Last, J.  A.; Kaizu, T. (1980) Mucus glycoprotein  secretion by tracheal explants:
     effects of pollutants.  EHP Environ. Health Perspect. 35; 131-138.

Last, J.  A.; Jennings, M. D.  ; Schwartz,  L. W.; Cross, C.  E.  (1977) Glycoprotein
     secretion by tracheal  explants  cultured  from  rats  exposed to ozone. Am.
     Rev. Respir. Dis. 116.:  695-703.

Last, J.  A.;  Greenberg,  D.  B.; Gastleman, W.  L.  (1979)  Ozone-induced altera-
     tions  in  collagen metabolism of rat lungs.  Toxicol. Appl,  Pharmacol.  51:
     247-258.

Last, J.  A.; Dasgupta, P. K.; DeCesare,  K.; Tarkington,  B. K. (1982) Inhalation
     toxicology  of  ammonium  persulfate, an oxidant aerosol,  in  rats. Toxicol.
     Appl.   Pharmacol. 63: 257-263.

Last, J.  A.;  Gerriets,  J. E.;  Hyde,  D.  M.  (1983) Synergistic effects on rat
     lungs  of  mixtures  of oxidant air  pollutants  (ozone or  nitrogen dioxide)
     and respirable aerosols. Am. Rev.  Respir. Dis. 128:  539-544.

Last, J.  A.;  Hyde, D. M.; Chang, D. P.  Y.  (1984a)  A mechanism  of  synergistic
     lung damage  by ozone and a  respirable  aerosol. Exp.  Lung Res. 7: 223-235,

Last, J. A.;  Reiser, K.  M.;  Tyler,  W.  S.; Rucker, R.  B. (1984b) Long-term
     consequences  of exposure to ozone.  I.  Lung collagen content.  Toxicol.
     Appl.  Pharmacol. 72: 111-118.

Lee, L.-Y.; Bleecker, E.  R.;  Nadel, J.  A.  (1977)  Effect  of ozone on  bronchomotor
     response to  inhaled  histamine aerosol  in  dogs. J. Appl.  Physiol.:  Respir.
     Environ. Exercise Physiol.  43: 626-631.

Lee, L.-Y.; Dumont,  C.;  Djokic,  T.  D.; Menzel,  T. E.;  Nadel,  J.  A. (1979)
     Mechanism of rapid,  shallow breathing after ozone  exposure in  conscious
     dogs.  J. Appl.  Physiol.:  Respir.  Environ. Exercise  Physiol.  46: 1108-1114.

Lee, L.-Y.; Djokic, T. D.; Dumont, C.;  Graf,  P. D.; Nadel, J. A.  (1980)  Mechanism
     of ozone-induced tachypneic  response  to  hypoxia and hypercapnia in  conscious
     dogs.  J.  Appl.  Physiol.: Respir.  Environ. Exercise  Physiol. 48: 163-168,

Lewis, T. R.; Hueter, F.  G.;  Busch,  K.  A.  (1967)  Irradiated  automobile  exhaust:
     its effects  on  the reproduction of mice.  Arch. Environ.  Health  15:  26-35.
                                    9-279

-------
Lewis, T. R.; Moorman, W. J.; Yang, Y.; Stara, J, F. (1974) Long-term exposure
     to auto exhaust and other pollutant mixtures: effects on pulmonary  function
     in the beagle. Arch. Environ. Health 29: 102-106.

Ling, N,  R.;  Kay,  J.  E.  (1975)  Lymphocyte  stimulation.  Amsterdam,  The Nether-
     lands: North-Holland Publishing Co.

Lum, H.;  Schwartz,  L.  W.; Dungworth,  D.  L.;  Tyler,  W.  S.  (1978) A  comparative
     study  of cell  renewal  after  exposure  to ozone or oxygen:  response of
     terminal bronchiolar epithelium  in the  rat. Am.  Rev.  Respir.  Dis.  118:
     335-345.

Lunan,  K,  D.; Short,  P.; Negi,  D.; Stephens, R.  J.  (1977) Glucose-6-phosphate
     dehydrogenase  response  of  postnatal  lungs  to  N02  and 03.  In:  Sanders,
     C.  L.;  Schneider, R.  P.;  Dagle, G.  E.; Ragan,  H.  A., eds. Pulmonary
     macrophage  and epithelial  cells:  proceedings of the  sixteenth annual
     Hanford  biology  symposium;  September  1976;  Richland, WA.  Washington, DC:
     Energy  Research  and  Development  Administration;  pp. 236-247.  (ERDA
     symposium series: 43).  Available  from:  NTIS, Springfield,  VA;  CONF-760927.

MacLean,  S.  A.;  Longwell, A.  C.;  Blogoslawski,  W.  J.  (1973) Effects of ozone-
     treated  seawater  on  the spawned, fertilized, meiotic, and cleaving eggs
     of the commercial American oyster. Mutat. Res. 21: 283-285,

MacRae, W.  D.;  Stich,  H. F.  (1979) Induction of sister chromatid exchanges in
     Chinese  hamster ovary cells  by  thiol  and  hydrazene compounds.  Mutat.
     Res. 68: 351-365.

Martin, C. J.; Boatman, E. S.; Ward, G. (1983) Mechanical properties of  alveo-
     lar wall after pneumonectomy  and  ozone  exposure. J.  Appl.  Physio!.:  Respir.
     Environ. Exercise Physio!.  54: 785-788.

Matsumura, Y. (1970) The effects of ozone, nitrogen dioxide, and sulfur  dioxide
     on  the experimentally induced allergic respiratory disorder  in guinea
     pigs.  III.  The effect  on the occurrence of dyspneic attacks.  Am.  Rev.
     Respir.  Dis. 102: 444-447.

Matsumura,  Y.,  Mizuno, K.;  Miyamoto,  T.;  Suzuki,  T.;  Oshima,   Y.  (1972) The
     effects  of  ozone, nitrogen dioxide, and sulfur  dioxide on the experi-
     mentally  induced allergic  respiratory  disorder  in  guinea pigs.  IV.
     Effects  on  respiratory sensitivity to  inhaled acetylene!ine.   Am.  Rev.
     Respir.  Dis. 105: 262-267.

Matzen,  R.  N. (1957a) Development  of  tolerance  to ozone  in reference to pul-
     monary edema. Am, J. Physio!.  190: 84-88.

Matzen,  R.  N. (1957b)  Effect of vitamin C  and hydrocortisone on the pulmonary
     edema produced by ozone  in mice.  J. Appl. Physio!. 11: 105-109.

McAllen,  S.  J.;  Chiu,  S.  P.;  Phalen,  R.  F.;  Rasmussen, R. E.  (1981) Effect of
     lH vivo  ozone  exposure on j_n  vitro pulmonary alveolar  macrophage mobility.
     J. Toxicol. Environ. Health  7: 373-381.
                                    9-280

-------
McCord, J. M.; Fridovich, I. (1978) The biology and pathology  of  oxygen  radicals.
     Ann. Intern. Med. 89: 122-127.

McJilton, C.; Thielke, J.; Frank, R. (1972) Ozone  uptake  model  for the respira-
     tory system.  In: Abstracts  of technical  papers:  American  industrial
     hygiene conference;  May;  San  Francisco, CA. Am.  Ind.  Hyg.   Assoc.  J.
     33: paper no. 45,

Meiners, B.  A.;  Peters,  R.  E. ; Mudd, J. B. (1977) Effects of  ozone on indole
     compounds  and  rat  lung  monoamine oxidase.  Environ. Res. 14:  99-112.

Mellick, P.  W.;  Mustafa,  M. G.; Tyler,  W.  S.; Dungworth,  D.  L. (1975) Morpho-
     logic  and  biochemical  changes  in  primate lung due  to  low-level ozone
     exposure. In: Proceedings  of the 23rd  international  tuberculosis conference;
     September; Mexico. Bull.  Int. Union Tubercul. 51:  565-567.

Mellick, P.  W.;  Dungworth,  D.  L.; Schwartz,  L. W.; Tyler, W.  S.  (1977)  Short
     term morphologic effects of high  ambient levels of ozone on lungs of
     rhesus monkeys.  Lab. Invest. 36: 82-90.

Mendenhall,  R.  M.;  Stokinger,  H. E.  (1959)  Tolerance and  cross-tolerance
     development to atmospheric pollutants  ketene  and ozone. J. Appl. Physiol.
     14: 923-926.

Menzel,  D.  B.  (1970) Toxicity  of ozone,  oxygen,  and radiation.  Annu.  Rev.
     Pharmacol. 10:  379-394.

Menzel,  D.  B.  (1971) Oxidation of biologically active  reducing substances by
     ozone.  Arch. Environ. Health 23: 149-153.

Menzel,  D.  B. (1976)  The  role  of  free radicals in  the toxicity of air pollutants
     (nitrogen  oxides and  ozone).  In:  Pryor, W.  A.,  ed. Free radicals  in
     biology: v.. 2.  New York,  NY: Academic  Press,  Inc.; pp.  181-202.

Menzel,  D.  B.  (1983) Metabolic effects  of ozone exposure.  In: Lee,  S.  D.;
     Mustafa,  M.  G.;  Mehlman,  M. A.,  eds.  International symposium  on  the
     biomedical  effects  of ozone  and related photochemical oxidants; March
     1982;  Pinehurst, NC.  Princeton,  NJ:   Princeton  Scientific Publishers,
     Inc.;  pp.  47-55. (Advances  in  modern environmental toxicology: v.  5).

Menzel,  D.  B.;  Roehm, J. N.;  Lee, S. D. (1972) Vitamin E: the biological and
     environmental antioxidant. J. Agric.  Food Chem.  20:  481-486.

Menzel,  D.  B.; Slaughter, R.;  Donovan,  D.;  Bryant, A.  (1973) Fatty acid  ozonides
     as  the toxic agents  on ozone  inhalation.  Pharmacologist 15:  239.

Menzel,  D.  B.;  Slaughter, R.  J.; Bryant,  A. M.; Jauregui, H. 0. (1975a) Heinz
     bodies formed  in erythrocytes by  fatty  acid ozonides  and ozone. Arch.
     Environ. Health  30:  296-301.

Menzel,  D.  B.; Slaughter, R.  J.;  Bryant, A. M.; Jauregui, H. 0.  (1975b)  Preven-
     tion of ozonide-induced Heinz bodies in human erythrocytes by vitamin E.
     Arch.  Environ. Health  30:  234-236.
                                    9-281

-------
Menzel, D.  B.;  Anderson,  W.  G.;  Abou-Donia,  M.  B.  (1976) Ozone exposure modi-
     fies  prostaglandin biosynthesis  in  perfused  rat  lungs.  Res.  Commun.
     Chem. Pathol. Pharmacol. 15: 135-147.

Menzel, D. B.; Donovan, D. H.; Sabransky, M.  (1978) Dietary vitamin  E  and  fat:
     effect on murine lung and red blood cell glutathione peroxidase activity.
     Presented  at: 62nd annual  meeting, Federation of American Societies  for
     Experimental Biology;  April; Atlantic  City,  NJ.  Fed.  Proc.  Fed. Am.  Soc.
     Exp. Biol. 37: 707.

Menzel, D.  B.;  Smolko,  E. D.; Gardner, D. E.; Graham, J. A. (1983)  Frontiers
     in short-term testing  of pneumotoxicants.  In: Woodhead,  ed.  Short-term
     tests  for  environmentally-induced chronic  health effects. Washington,
     DC:  U.S.   Environmental  Protection  Agency,  Office  of Research  and
     Development; EPA report  no. EPA-600/8-83-002.

Miller, F. J. (1977) A mathematical model of  transport and removal of  ozone  in
     mammalian  lungs [Ph.D.  thesis].   Raleigh,  NC:  North Carolina State
     University.  Available from: University  Microfilms, Ann Arbor,  MI; publi-
     cation no. 77-21,560.

Miller, F.  J.  (1979) Biomathematical  modeling applications in the evaluation
     of ozone toxicity.   In:  Mudd, J.;  Lee, S. D.;  eds. Assessing toxic effects
     of environmental pollutants. Ann  Arbor,  MI: Ann Arbor Science Publishers,
     Inc.; pp. 263-286.

Miller,  S,;  Ehrlich, R.  (1958)  Susceptibility  to respiratory infections  of
     animals exposed  to ozone.  I. Susceptibility  to K_iebsi_e11a pneumoniae. 0.
     Infect. Dis. 103: 145-149.

Miller, F.  J.;  Illing,  J. W.; Gardner,  D.  E. (1978a) Effect of urban ozone
     levels  on  laboratory-induced  respiratory infections.  Toxicol.  Lett.
     2: 163-169.

Miller, F. J.; Menzel, D. B.; Coffin,  D. L. (1978b) Similarity between man and
     laboratory  animals in  regional deposition of  ozone.  Environ.   Res. 17:
     84-101.

Miller, F.  J.;  McNeal,  C. A.; Kirtz,  J.  M.;  Gardner, D. E.; Coffin,  D. L.;
     Menzel,  D.  B.   (1979)  Nasopharyngeal  removal  of  ozone  in rabbits and
     guinea pigs. Toxicology  14: 273-281.

Miller,  F.  J.; Overton,  J.  H.,  Jr.; Jaskot, R.  H.;  Menzel,  D.  B.; (1985)
     A model  of the  regional  uptake of gaseous  pollutants in  the lung. I.  The
     sensitivity of the uptake of ozone in the human lung to lower respiratory
     tract  secretions  and to exercise. Toxicol.  Appl.  Pharmacol.  79:  11-27.

Mizoguchi,  I.;  Osawa,  M.; Sato, Y.; Makino,  K,;  Yagyu,  H.  (1973)  Studies on
     erythrocyte  and photochemical   smog.  I.  Effects  of air pollutants on
     erythrocyte resistance.  Taiki Osen Kenkyu 8: 414.

Mole, M.  L.,  Jr.;  Stead,  A.  G.;  Gardner,  D.  E.  ;  Miller, F.  J.;  Graham,  J. A.
     (1985)  Effect  of ozone  on  serum  lipids  and lippoproteins  in the rat.
     Toxicol. Appl. Pharmacol. 80: 367-376.


                                   9-282

-------
Montgomery, M.  R.;  Niewoehner, D.  E.  (1979)  0x1dant-induced alterations  in
     pulmonary microsomal mixed-function oxidation:  acute effects of paraquat
     and ozone. J. Environ.  Sci, Health  C13: 205-219.

Moore, P.  F.;  Schwartz, L. W. (1981) Morphological effects of prolonged  exposure
     to ozone  and  sulfuric  acid aerosol  on the rat lung. Exp. Mol.  Pathol.
     35: 108-123.

Moore, G.  S.;  Calabrese,  E.  J.; Grinberg-Funes, R.  A.  (1980) The  C57L/J mouse
     strain as  a model  for extrapulmonary effects  of ozone exposure. Bull.
     Environ.  Contam. Toxicol. 25:  578-585.

Moore, G.  S.;  Calabrese,  E.  J.; Labato,  F. J.  (1981a) Erythrocyte survival in
     sheep  exposed  to ozone.  Bull. Environ.  Contam.  Toxicol.  27:  126-138.

Moore, G.  S.;  Calabrese,  E.  J.;  Schulz, E.  (1981b) Effect  of  in vivo ozone
     exposure  to  Dorset sheep,  an animal model  with low levels of erythrocyte
     glucose-6-phosphate dehydrogenase activity. Bull.  Environ. Contam.  Toxicol.
     26: 273-280.

Moore, G.  S.  ; Calabrese, E.  J.;  Molteni,  K.   H. (1984)  Plasmodium berghei
     infection  in  mice:  effect of  low-level  ozone  exposure. Bull.  Environ.
     Contam. Toxicol. 33:  99-105.

Moorman, W. J.;  Chmiel,  J.  J.; Stara, J. F.;-Lewis,  T. R. (1973) Comparative
     decomposition of  ozone  in the  nasopharynx of  beagles:  acute vs chronic
     exposure. Arch. Environ. Health 26:  153-155.

Morgan, M.  S.; Frank,  R.  (1977)  Uptake  of pollutant  gases by  the respiratory
     system.  In:  Brain,   J.  D.;  Proctor,  D.  F., eds.  Respiratory defense
     mechanisms.  Part  I.  New York, NY:  Marcel Dekker,  Inc.;  pp.  157-189.

Mountain,  J.  T.  (1963) Detecting hypersusceptibility to  toxic  substances:  an
     appraisal of simple blood  tests. Arch. Environ.  Health  6:  357-365.

Mudd, J. B.; Freeman, B. A.  (1977)  Reaction of  ozone  with biological  membranes.
     In:   Lee,  S.  D., ed. Biochemical  effects of environmental  pollutants.
     Ann Arbor,  MI:  Ann Arbor  Science Publishers, Inc.; pp.  97-133.

Mudd, J.  B. ;  Leavitt,  R.; Qngun, A.; McManus,  T. T.  (1969)  Reaction of ozone
     with  amino  acids and proteins.  Atmos. Environ.  3:  669-681.

Murlas, C. G.;  Roum, J. H. (1985) Sequence of pathologic  changes  in the  airway
     mucosa of guinea  pigs  during ozone-induced bronchial hyperractivity. Am.
     Rev.  Respir. Dis. 131:  314-320.

Murphy, S.  D.  (1964) A review of  effects on animals  of exposure to  auto exhaust
     and  some  of its components.  J.  Air Pollut. Control  Assoc.  14: 303-308.

Murphy,  S.  D.; Leng, J.  K.;  Ulrich, C.   E.; Davis,  H.  V. (1963) Effects on
     animals  of exposure to auto exhaust.  Arch. Environ. Health 7: 60-70.

Murphy,  S.  D.; Ulrich, C. E.;  Frankowitz, S.  H.; Xintaras,  C.  (1964) Altered
     function  in animals inhaling  low  concentrations of ozone and nitrogen
     dioxide.  Am.  Ind. Hyg.  Assoc.  J. 25:  246-253.

                                    9-283

-------
Mustafa, M. G. (1975) Influence of dietary vitamin E on  lung  cellular sensiti-
     vity to ozone in rats. Nutr. Rep.  Int. 11: 473-476.

Mustafa, M.  G.;  Cross,  C. E.  (1974)  Effects  of short-term ozone exposure on
     lung  mitochondria!  oxidative  and  energy metabolism. Arch.  Biochem.
     Biophys. 162: 585-594.

Mustafa, M.  G.;  Lee,  S.  D.  (1976)  Pulmonary biochemical  alterations resulting
     from ozone exposure. Ann. Occup. Hyg. 19:  17-26.

Mustafa, M.  G.;  Lee,  S.  D. (1979) Biological  effects  of environmental  pollu-
     tants:  methods   for  assessing biochemical  changes.  In:  Lee,  S.  D.;
     Mudd, J.  B.,  eds.  Assessing toxic  effects of environmental pollutants.
     Ann Arbor, MI: Ann Arbor  Science Publishers,  Inc.;  pp. 105-120.

Mustafa, M. G.; Tierney,  D. F. (1978) Biochemical  and  metabolic  changes in  the
     lung with oxygen,  ozone,  and  nitrogen dioxide toxicity.  Am. Rev. Respir.
     Dis. 118: 1061-1090.

Mustafa, M.  G.;  DeLucia,  A. J.; York,  G.  K.; Arth, C.;  Cross,  C.  E.  (1973)
     Ozone interaction  with rodent lung. II.  Effects on oxygen consumption of
     mitochondria. J. Lab. Clin. Med. 82: 357-365.

Mustafa, M. G.; Ospital,  J. J.; Hacker,  A. D.;  Lee,  S. D.  (1976) Effects of ozone
     and  nitrogen  dioxide exposure on  lung  metabolism.  EHP  Environ.  Health
     Perspect. 16: 184.

Mustafa, M.  G.;  Hacker, A. D.; Ospital, J.  J.; Hussain,  M.  Z.; Lee, S. D.
     (1977) Biochemical effects of environmental oxidants  pollutants  in animal
     lungs.  In:  Lee,  S. D., ed.  Biochemical  effects  of environmental  pollu-
     tants.  Ann  Arbor,  MI: Ann  Arbor Science Publishers,  Inc.;  pp.  59-96.

Mustafa, M. G.; Faeder, E. J.; Lee, S.  D. (1980) Biochemical  basis  of pulmonary
     response  to  ozone and nitrogen  dioxide  injury.  In:  Bhatnagar,  R. S.,
     ed. Molecular basis  of environmental toxicity. Ann Arbor,  MI: Ann Arbor
     Science Publishers,  Inc.; pp. 151-172.

Mustafa, M. G.; Elsayed,  N. M.; Quinn,  C. L.;  Postlethwait, E. M.;  Gardner, D.
     E.; Graham, J.  A.  (1982) Comparison of pulmonary biochemical  effects  of
     low level ozone  exposure on mice  and  rats.  J.  Toxicol.  Environ. Health
     9: 857-865.

Mustafa, M.  G.;  Elsayed,  N. M.; Graham, J. A.; Gardner,  D. E. (1983) Effects
     of ozone  exposure  on lung metabolism: influence  of animal  age,  species,
     and  exposure  conditions.  In:  Lee,  S.  D.; Mustafa, M. G.;  Mehlman, M.
     A., eds.  International  symposium on the  biomedical  effects of ozone and
     related photochemical oxidants; March 1982; Pinehurst, NC.  Princeton,  NJ:
     Princeton Scientific Publishers,  Inc.;  pp.  57-73.  (Advances  in modern
     environmental toxicology: v. 5).

Mustafa, M. G.; Elsayed,  N. M.; von Dohlen, F.  M.; Hassett, C. M.;  Postlethwait,
     E. M.; Quinn, C. L.; Graham, J. A.; Gardner,  D. E.  (1984) A comparison of
     biochemical effects  of nitrogen  dioxide,  ozone, and their combination in
     mouse lung.  I. Intermittent exposures. Toxicol. Appl.  Pharmacol.  72: 82-90.


                                   9-284

-------
Myers, B. A.;  Dubick,  M.  A.;  Reiser,  K.  M.;  Gerriets, J.  E. ;  Last, J.  A.;
     Rucker,  R.  B.  (1984)  Ozone  exposure,  food  restriction  and  protein
     deficiency: changes  in collagen  and elastin in  rodent  lung.  Toxicol.
     Lett. 23: 43-49.

Nadel, J.  A.   (1977)  Autonomic control of airway  smooth  muscle and  airway
     secretions. Am. Rev.  Respir. Dis. 115: 117-126.

Nakajima, T.;  Kusumoto, S.;  Tsubota,  Y.;  Yonekawa, E.; Yoshida,  R.; Motomiya,
     K.;   Ito,  K.;  Ide,  G.;  Otsu, H. (1972). Histopathological studies on the
     respiratory organs of  mice  exposed  to photochemical  oxidants and automo-
     bile exhaust.  Osaka-furitsu  Koshu  Eisei  Kenkyusho  Kenkyu  Hokoku  Rodo
     Eisei Hen 10:  35-42.

Nambu, Z.; Yokoyama, E. (1983) Antitoxidant system and ozone tolerance.  Environ.
     Res. 32:   111-117.

Nasr,  A.  N. M.  (1967)  Biochemical  aspects of ozone intoxication: a review.  J.
     Occup.  Med. 9: 589-597.

National  Air Pollution Control Administration. (1970)  Air quality criteria  for
     photochemical  oxidants.   Washington,  DC:  U.S.  Department  of  Health,
     Education and Welfare, Public Health  Service; NAPCA  publication  no. AP-63.
     Available from: NTIS, Springfield, VA; PB-190262.

National   Institutes  of Health. (1983) Supplement: comparative biology of the
     lung: workshop;  June 1982.  Am.  Rev.  Respir.  Dis.  Suppl.  128:  (2, pt.  2):
     S1-S91.

National   Institutes  of Health  (1985)  The definition of emphysema:  report of a
     National  Heart,  Lung,   and  Blood Institute,  Division  of  Lung  Diseases
     workshop.  Am. Rev.  Respir. Dis.  132: 182-185.

National  Research   Council.   (1977)   Ozone   and  other  photochemical
     oxidants.  Washington,  DC: National  Academy of Sciences,  Committee on
     Medical and Biologic Effects  of Environmental Pollutants.

Nayak, D. P.;  Kelley, G.  W.; Young, G. A.; Underdahl,  N.R.  (1964).   Progressive
     descending  influenza infection in mice  determined by immunofluorescence.
     Proc. Soc. Exp. Biol. Med. 116: 200-206.

Niewoehner, D.  E.;  Kleinerman, J.; Rice,  D.  B.  (1974) Pathologic changes in
     the  peripheral  airways of  young cigarette  smokers. N.  Engl.  J. Med.
     291: 755-758.

North  Atlantic Treaty Organization. (1974) Air quality criteria  for photochemical
     oxidants  and  related hydrocarbons:  a report  by the expert panel on air
     quality criteria. Brussels, Belgium:  NATO,  Committee on  the Challenges of
     Modern Society; NATO/CCMS n.  29.

O'Byrne,  P.  M.; Walters,  E. H. ;  Gold, B. D.; Aizawa,  H. A.;  Fabbri, L. M.;
     Alpert, S. E.; Nadel, J.  A.;  Holtzman, M. J.  (1984a) Neutrophil  depletion
      inhibits  airway hyperresponsiveness induced  by ozone  exposure.  Am.  Rev.
     Respir. Dis. 130: 214-219.


                                    9-285

-------
O'Byrne, P.  M.;  Walters, E. H.; Aizawa,  H.;  Fabbri, L.  M.;  Holtzman, M. J.;
     Nadel, J. A. (1984b) Indomethacin  inhibits the  airway hypperresponsiveness
     but not  the neutrophil  influx induced by ozone in dogs.  Am.  Rev. Respir.
     Dis. 130: 220-224.

O'Byrne, P. M.; Aizawa,  H.; Bethel, R.  A.; Chung,  K.  F.;  Nadel, J.  A.;  Holtzman,
     M.  J.  (1984c) Prostaglandin  F2a  increases responsiveness of  pulmonary
     airways  in dogs.  Prostaglandins 28:  537-543.

Oberst,  F.  W.;  Comstock, C. C.; Hackley,  E.  B. (1954)  Inhalation toxicity of
     ninety percent  hydrogen peroxide  vapor. AMA  Arch.  Ind.  Hyg. Occup.  Med.
     10: 319-327.

Ooraichi, S.;  Kita,  H.  (1974)  Effect of air pollutants  on ciliary activity of
     respiratory tract.  Bull.  Tokyo Med.  Dent.  Univ.  21:  327-343.

Orthoefer, J. G.; Bhatnagar, R. S.; Rahman, A.; Yang, Y.  Y.;  Lee, S.  D.;  Stara,
     J.  F.  (1976)  Collagen and prolyl  hydroxylase levels in  lungs  of beagles
     exposed  to air pollutants. Environ.  Res. 12:  299-305.

Osebold, J. W.; Owens, S. L.;  Zee, Y. C.;  Dotson,  W.  M.;  LaBarre, D.  D.  (1979)
     Immunological alterations in the lungs of mice following ozone exposure:
     changes  in immunoglobulin  levels  and antibody-containing cells.  Arch.
     Environ. Health 34: 258-265.

Osebold, J.  W.;  Gershwin,  L.  J.;  Zee,  Y.  C.  (1980) Studies  on the enhancement
     of  allergic  lung  sensitization by inhalation of ozone and sulfuric acid
     aerosol. J. Environ. Pathol.  Toxicol. 3: 221-234.

Pack,  A.;  Hooper, M.  B.;  Nixon, W.; Taylor,  J.  C. (1977) A computational
     model  of pulmonary gas  transport incorporating effective diffusion.
     Respir.  Physio!. 29: 101-124.

Paiva, M.  (1973) Gas transport in the  human  lung.  J. Appl. Physio!.  35:  401-
     410.

Palmer,  M.  S.j Swanson,  D, H.; Coffin,  D.  L.  (1971)  Effect of ozone on benzpyrene
     hydroxylase  activity in  the  Syrian  golden  hamster. Cancer  Res.  31:
     730-733.

Palmer,  M.  S.; Exley, R. W.; Coffin, D.  L.  (1972)  Influence of pollutant gases
     on  benzpyrene  hydroxylase activity.  Arch. Environ.  Health  25: 439-442.

P'an,  A.  Y.  S.;  Jegier,  I.  (1971) The serum trypsin inhibitor capacity during
     ozone  exposure. Arch.  Environ. Health 23:  215-219.

P'an,  A.  Y.  S.; Jegier, Z.  (1972) Trypsin protein  esterase  in  relation to
     ozone-induced vascular  damage. Arch.  Environ.  Health 24: 233-236.

P'an, A. Y. S.; Jegier,  Z.  (1976)  Serum protein changes during 'exposure to ozone.
     Am. Ind. Hyg. Assoc. J. 37: 329-334.

P'an,  A.  Y.  S.;  Beland,  J.; Jegier,  Z.  (1972) Ozone-induced  arterial lesions.
     Arch,  Environ. Health 24:  229-232.


                                   9-286

-------
Pemsingh, R. S.;  Atwal,  0.  S. (1983)  Occurrence of APUD-type cells  in  the
     ciliated cyst of the parathyroid  gland of ozone-exposed dogs.  Acta  Anat.
     116: 97-105.

Penha, P.  D. ;  Werthamer, S.  (1974)  Pulmonary lesions induced by  long-term
     exposure to  ozone.  II.  Ultrastructure observations of proliferative and
     regressive lesions.  Arch, Environ. Health 29: 282-289.

Pepelko,  W.  E.; Mattox, J.  K.; Yang, Y. Y.  (1980) Arterial blood gases, pulmonary
     function and pathology  in  rats  exposed to 0.75 or  1.0 ppm ozone.  J.
     Environ.  Pathol. Toxicol. 3: 247-259.

Peterson, D. C.; Andrews, H.  L. (1963) The  role of ozone in radiation avoidance
     in the mouse. Radiat.  Res. 19: 331-336.

Peterson, M. L.;  Rummo,  N.;  House,  D.;  Harder,  S.  (1978a)  The effect of  ozone
     on human immunity:  in vitro responsiveness of lymphocytes to phytohemmag-
     glutinin.  Arch.  Environ. Health 33: 59-63.

Peterson, M. L.;  Harder,  S.; Rummo, N.; House, D. (1978b). Effect of ozone  on
     leukocyte function  in  exposed human  subjects.  Environ.  Res.  15:  485-493.

Phalen,  R.  F.;  Kenoyer,  J.  L.; Crocker, T. T.; McClure, T. R. (1980) Effects
     of  sulfate aerosols  in combination with  ozone  on  elimination of tracer
     particles inhaled by rats. J. Toxicol. Environ. Health 6: 797-810.

Pirquet,  C.; Schittemhelm,  A., eds.  (1921) Uber Kampfgasvergiftungen [War gas
     poisoning]. I. Gesamte Exp.  Med. 13:  284-296.

Plopper,  C. G.;  Chow,  C. K.;  Dungworth,  D. L.; Brummer,  M.;  Nemeth,  T.  J.
     (1978) Effect of low  level of ozone  on  rat lungs.   II. Morphological
     responses during recovery and re-exposure. Exp. Mol.  Pathol. 29: 400-411.

Plopper,  C. G.;  Chow,  C. K.; Dungworth, D. L.; Tyler, W.  S.  (1979)  Pulmonary
     alterations  in rats exposed to 0.2 and 0.1 ppm  ozone: a  correlated morpho-
     logical and  biochemical  study. Arch.  Environ. Health  34: 390-395.

Prat,  R.; Nofre,  C.; Cier, A. (1968)  Effets  de Thypochlorite de  sodium de
     1'ozone et des radiations ionisantes  sur  les constituents pyrimidiques d1
     Escherichia  coli  [Effects of  sodium  hypochlorite,  ozone, and  ionizing
     radiations on pyrimidine constituents of  Escherichia co1i].  Ann.  Inst.
     Pasteur Paris 114: 595-607.

Previero, A.;  Signer,  A.;  Bezzi,  S.  (1964) Tryptophan  modification in poly-
     peptide chains.  Nature  (London) 204:  687-688.

Pryor, W. A.  (1976)  Free radical  reactions in biology:  initiation  of lipid
     autoxidation by ozone  and nitrogen dioxide. EHP Environ. Health  Perspect.
     16:  180-181.

Pryor, W. A.;  Stanley,  J.   P.; Blair,  E; Cullen, G.  B. (1976) Autoxidation  of
     polyunsaturated fatty  acids.  Part I.  Effect of ozone on the autoxidation
     of  neat methyl  linoleate and  methyl linolenate. Arch. Environ.  Health  31:
     201-210.


                                    9-287

-------
Pryor, W. A.; Dooley,  M.  M.;  Church,  D.  F.  (1983)  Mechanisms  for the reaction
     of ozone with  biological molecules, the  source  of the toxic effects of
     ozone.  In:  Lee, S. D.; Mustafa, M. G.; Mehlman, M. A., eds.  International
     symposium on the  biomedical  effects  of ozone  and photochemical  oxidants;
     March 1982; Pinehurst, NC. Princeton, NJ: Princeton Scientific  Publishers,
     Inc.; pp.  7-19.   (Advances  in modern environmental toxicology:  v.  5).

Raub, J. A.;  Miller,  F. J.; Graham, J. A. (1983a) Effects of low-level ozone
     exposure on  pulmonary function  in adult and neonatal  rats. In: Lee,
     S.  D.;  Mustafa, M.  G.;  Mehlman, M. A.,  eds.  International  symposium on
     the biomedical  effects of ozone and related photochemical oxidants;  March
     1982; Pinehurst,  NC  Princetgn,  NJ:  Princeton  Scientific  Publishers,
     Inc.; pp. 363-367.  (Advances in modern environmental toxicology: v. 5).

Raub, J.  A.;  Miller,  F.  J.;  Graham,  J.  A.; Gardner,  D.  E.;  O'Neil, J.   J.
     (1983b) Pulmonary function in normal and  elastase-treated hamsters exposed
     to a complex mixture of olefin-ozone-sulfur  dioxide  reaction products.
     Environ.  Res.  31:  302-310.

Raut, S.; Hurd,  J.;  Cureton, R.  J.;  Blandford,  G.; Heath, R. B.  (1975)  The
     pathogenesis of infections  of the mouse caused by virulent and avirulent
     variants of an influenza virus.  J. Med.  Microbiol. 8: 127-136.

Razumovstdi, S.  D.;  Zaikov, G. E. (1972) Effect of structure of  an unsaturated
     compound on rate  of its reaction with  ozone. J.  Org.  Chem. USSR (Engl.
     Transl.) 8: 468-472.

Reasor, M. J.; Adams, G.  K.,  III; Brooks, J. K.; Rubin, R. J. (1979)  Enrichment
     of albumin  and  IgG in the airway secretions  of dogs breathing  ozone. J.
     Environ.  Sci.  Health CIS: 335-346.

Reddy, C. C.; Thomas,  C. E.; Scholz,  R.  W.;  Labosh,  T. J.;  Massaro, E.  J.
     (1983)  Effects of ozone  exposure  under  inadequate vitamin E and/or selenium
     nutrition on enzymes associated with xenobiotic  metabolism.  In:  Lee, S.
     D,; Mustafa, M. G.; Mehlman, M. A., eds. International symposium on the
     biomedical   effects of ozone  and  related photochemical oxidants; March
     1982,  Pinehurst,  NC. Princeton,  NJ:  Princeton  Scientific  Publishers,
     Inc., pp.  395-410.  (Advances in modern environmental toxicology: v. 5).

Revis, N. W.; Major, T.;  Dalbey,  W.  E.  (1981)  Cardiovascular  effects of  ozone
     and  cadmium inhalation  in  the rat. In:  Northrop Services, Inc., ed.
     Proceedings of the research  planning workshop on  health effects  of oxidants;
     January 1980; Raleigh, NC. Research Triangle  Park, NC: U.S.  Environmental
     Protection Agency,  Health Effects Research  Laboratory;  EPA-600/9-81-001;
     pp. 171-179. Available from: NTIS, Springfield, VA; PB81-178832.

Reynolds, R.  W.;  Chaffee, R. R.  J. (1970) Studies on  the combined effects of
     ozone and  a hot environment on  reaction  time in subhuman  primates.  In:
     Project  clean  air: v. 2. Santa  Barbara,  CA:  University of California;
     research project S-6.

Roehm, J. N.; Hadley, J.  G.;  Menzel, D. B.  (1971a) Antioxidants  vs  lung disease.
     Arch. Intern. Med. 128:  88-93.
                                   9-288

-------
Roehm, J. N.;  Hadley,  d.  G.; Menzel,  D,  B.  (1971b)  Oxidation  of unsaturated
     fatty acids by ozone and nitrogen dioxide:  a  common  mechanism of action.
     Arch. Environ. Health 23:  142-148.

Roehm, J. N.;  Hadley,  J.  G.; Menzel, D. B. (1972) The influence of vitamin  E
     on the  lung fatty acids of rats exposed to ozone. Arch.  Environ. Health
     24:  237-242.

Ross, B.  K.; Hlastala, M.  P.; Frank, R. (1979) Lack  of ozone effects  on oxygen
     hemoglobin affinity.  Arch.  Environ. Health 34:  161-163.

Roth, R.   P.; Tansy, M.  F. (1972)  Effects  of gaseous  air pollutants on gastric
     secreto-motor activities in  the rat. J. Air  Pollut.  Control  Assoc.  22:
     706-709.

Roum, J.   H.; Murlas,  C. (1984) Ozone-induced changes  in muscaranic bronchial
     reactivity by different testing  methods.  J. Appl.  Physiol.: Respir.
     Environ. Exercise Physiol.  57: 1783-1789.

Roycroft, J. H.; Gunter, W.  B.; Menzel, D.  B.  (1977) Ozone toxictty:  hormone-
     like oxidation products from arachidonic acid by ozone-catalyzed autoxi-
     dation. Toxicol.   Lett.  1:  75-82.

Sachsenmaier,  W.;  Siebs,  W.;   Tan,  T.-A.  (1965) Wirkung von  Ozon  auf
     Mauseacitestumorzellen  und auf Huhnerfibroblasten  in der  Gewebekultur
     [Effect of ozone  on mouse ascites  on tumor cells  and  on chick fibroblasts
     in tissue culture].  Z. Krebsforsch. 67: 113-126.

Saltzman, B. E.;  Gilbert,  N,  (1959) lodometric microdetermination of organic
     oxidants and  ozone: resolution of mixtures by kinetic  colorimetry.   Anal.
     Chem. 31:  1914-1920.

Sato, S.; Kawakami,  M.; Maeda,  S.; Takishima,  T.  (1976a)  Scanning electron
     microscopy of the lungs of  vitamin  E-deficient rats  exposed to a low
     concentration of  ozone. Am.  Rev.  Respir. Dis. 113: 809-821.

Sato, S.; Kawakami, M.; Maeda,  S.; Takishima, T.  (1976b)  Electron microscopic
     studies on the effects of  ozone on the  lungs  of vitamin E-deficient  rats.
     Nippon  Kyobu  Shikkan Gakkai  Zasshi 14:  355-365.

Sato, S.; Shimura, S.  ; Kawakami,  M. ;  Hirose, T.;  Maeda,  S.;  Takishima,  T.;
     Kimura, S.;   Yashiro,  M.;  Okazaki, S.; Ito,  M.   (1978) Biochemical  and
     ultrastructural  studies  on the effects of long-term  exposure of ozone on
     vitamin E-depTeted rats.  Nippon Kyobu Shikkan Gakkai Zasshi 16:  260-268.

Sato, S.; Shimura, S.; Hirose, T.;  Maeda,  S.;  Kawakami,  M.;  Takishima,  T.;
     Kimura, S. (1980) Effects  of long-term ozone  exposure and dietary vitamin
     E in rats. Tohoku J. Exp.  Med. 130:  117-128.

Savino,  A.;  Peterson,  M. L.; House,  D.; Turner,  A. G.; Jeffries, H.  E.;  Baker,
     R.  (1978)  The effects of  ozone  on human cellular and humoral immunity:
     Characterization  of  T and B  lymphocytes by rosette  formation.  Environ.
      Res. 15:  65-69.
                                    9-289

-------
Savolainen, H.;  Zitting,  A.  (1980) dial  cell  effects of subacute formic acid
     vapour exposure [letter  to  the editor],  Acta  Pharmacol.  Toxicol.  47:
     239-240.

Scheel, L. D.; Dobrogorski, 0. J.;  Mountain, J. T.;  Svirbely,  J.  L.;  Stokinger,
     H. E. (1959) Physiologic, biochemical, immunologic  and  pathologic changes
     following ozone exposure. J.  Appl. Physio!.  14:  67-80.

Scherer, P. W.; Shendalman, L. H.;  Greene, N.  M.  (1972)  Simultaneous  diffusion
     and convection  in  single breath  lung washout.  Bull.  Math. Biophys. 34:
     393-412.

Scherer, P. W.;  Shendalman,  L.  H.; Greene,  N.  M.; Bouhuys, A.  (1975) Measure-
     ment  of  axial  diffusivities  in a model  of  the bronchial airways.  J.
     Appl. Physio!. 38: 719-723.

Schlipkoter, H.-W.; Bruch, J.  (1973) Funktionelle und morphologische  Veranderung
     bei Qzonexposition  [Functional and morphological alterations  caused by
     exposure to ozone], Zentralbl. Bakteriol.  Parasitenkd.  Infektionskr.  Hyg.
     Abt. 1: Orig. Reihe B 156: 486-499.

Schreider, J.  P.;  Raabe,  0. G. (1981) Anatomy of the nasal-pharyngeal airway
     of experimental animals. Anat. Rec. 200:  195-205.

Schwartz, L. W.; Christman, C. A.  (1979) Alveolar macrophage migration:  influ-
     ence  of  lung lining material  and acute  lung insult. Am.  Rev.  Respir.
     Dis. 120: 429-439.

Schwartz,  L.  W.;  Dungworth,  D.  L.; Mustafa,  M. G.; Tarkington, B. K.; Tyler,
     W.  S.  (1976) Pulmonary responses  of  rats to ambient levels of ozone:
     effects  of  7-day  intermittent or continuous exposure.  Lab.  Invest. 34:
     565-578.

Scott, M.  L.  (1970)  Studies on vitamin E and related factors in nutrition and
     metabolism.  In: DeLuch,  H.  F.; Suttie,  J. W., eds.' The fat-soluble vita-
     rains. Madison, WI: University of Wisconsin Press; pp. 355-374.

Scott, D.  B.  M.;  Lesher,  E.  C.  (1963)  Effect of ozone on survival and permea-
     bility of Esdieridrla co]_[. J. Bacteriol.  85: 567-576.

Selgrade, M.  J.  K.;  Mole,  M.  L.;  Miller,  F.  J.; Hatch,  G.  E.;  Gardner,  D. E.;
     Hu, P. C.  (1981)  Effect  of  N02  inhalation and  vitamin C deficiency on
     protein  and  lipid  accumulation in the  lung.  Environ. Res. 26:  422-437.

Shakman, R. A.  (1974)  Nutritional  influences on the toxicity of environmental
     pollutants: a review. Arch. Environ. Health  28:  105-113.

Sherwin, R. P.; Richters, V.; Okimoto, D. (1983)  Type 2  pneumocyte  hyperplasia
     in  the lungs of mice exposed to an ambient level (0.3 ppm) of ozone. In:
     Lee,  S.  D.;  Mustafa,  M.  G.;  Mehlman,  M.  A.,  eds. International symposium
     on  the biomedical  effects of  ozone  and  related photochemical  oxidants;
     March 1982;  Pinehurst, NC. Princeton, NJ:  Princeton Scientific  Publishers,
     Inc.; pp.  289-297.  (Advances  in  modern environmental toxicology: v.  5).
                                    9-290

-------
Sherwood, R,  L.;  Kimura, A.;  Donovan,  R.;  Goldstein,  E.  (1984)  Effect of 0.64
     ppm ozone on rats with chronic pulmonary bacterial infection. J. Toxicol.
     Environ. Health 13: 893-904.

Shimura, S.; Maeda, S.; Takismima, T. (1984) Giant lamellar bodies in alveolar
     type II cells of rats exposed to a low concentration of ozone.  Respiration
     46: 303-309.

Shingu, H.;  Sugiyama,  M.;  Watanabe, M.; Nakajima, T. (1980) Effects of  ozone
     and photochemical  oxidants  on interferon production by  rabbit  alveolar
     macrophages. Bull. Environ.  Contam. Toxicol. 24: 433-438.

Shinriki, N.; Ishizaki, K.; Miura, K.; Ueda, T.; Harada, F. (1983) Degradation
     of nucleic acids with ozone. III. Mode of ozone-degradation  of  mouse  pro-
     line transfer  ribonucleic acid (tRNA) and  isoleucine tRNA.  Chem. Pharm.
     Bull. (Tokyo) 31: 3601-3608.

Sielczak, M.  W.;  Denas, S. -M.;  Abraham,  W.  M.  (1983) Airway  cell changes in
     tracheal lavage of sheep  after ozone  exposure. J. Toxicol. Environ. Health
     11: 545-553.

Sparrow, A.  H.;  Schairer,  L.   A. (1974) Mutagenic response of Tradescantia to
     treatment with X-rays, EMS, DBE, ozone, SOg, N20 and several  insecticides.
     Am. Environ. Mutagen  Soc. 445: 265-276.

Speit,  G.;  Vogel, W. (1982)  The effect of sulfhydryl compounds  on  sister-
     chromatid  exchanges.  II.  The question of cell  specificity and  the role
     of H202. Mutat. Res.  93:  175-183.

Speit,  G.;  Vogel, W.; Wolf, M.  (1982)  Characterization of sister chromatid
     exchange  induction by hydrogen peroxide. Environ.  Mutagen.  4:  135-142.

Stara,  J.  F.; Dungworth, D.  L.; Orthoefer,  J.  G.;  Tyler,  W.  S.,  eds.  (1980)
     Long-term  effects  of air pollutants: in canine  species. Cincinnati,  OH:
     U.S. Environmental Protection Agency, Environmental Criteria and  Assessment
     Office;  EPA  report no. EPA-600/8-80-014. Available  from: NTIS,  Springfield,
     VA; PB81-144875.

Stephens, R.  J.; Freeman, G.; Stara,  J.  F.;  Coffin, D. L. (1973) Cytologic
     changes  in  dog  lungs  induced by chronic exposure to ozone. Am.  J.  Pathol.
     73: 711-726.

Stephens, R.  J.;  Sloan, M.  F;; Evans,  M.  d.; Freeman, 6. (1974a)  Early response
     of lung to  low  levels of ozone. Am.  J. Pathol.  74:  31-58.

Stephens, R.  J.;  Sloan, M.  F.; Evans,  M.  J.; Freeman, G. (1974b)  Alveolar type
     1  cell  response to exposure  to  0.5 ppm 03 for  short periods. Exp. Mol.
     Pathol.  20:  11-23.

Stephens,  R. J.; Sloan, M.  F.;  Groth,  D.  G.  (1976)  Effects of  long-term,  low
     level  exposure of N02 or 08  on  rat lungs.  EHP  Environ.  Health  Perspect.
     16:  178-179.
                                    9-291

-------
Stephens, R.  J.;  Sloan,  M.  F.;  Groth,  D.  G.;  Negi, D.  S.;  Lunan, K.  D. (1978)
     Cytologic response  of postnatal rat  lungs  to 03  or N02  exposure.  Am.  J.
     Pathol. 93: 183-200.

Stephens, R.  J.;  Buntman^ D.  J.;  Negi,  D.  S.; Parkhurst,  R.  M.;  Thomas, D.  W.
     (1983) Tissue  levels' of vitamin E in the lung and the cellular  response
     to injury resulting from oxidant gas  exposure.  Chest 83: 37S-39S.

Stewart, R. M.; Weir, E. K.; Montgomery, M. R.;  Niewoehner, D. E.  (1981)  Hydro-
     gen peroxide contracts airway smooth  muscle:  a possible  endogenous mecha-
     nism. Respir. Physio!. 45: 333-342.

Stich, H. F.; Wei, L.; Lam, P.  (1978) The  need for a mammalian test for mutagens:
     action of some reducing agents. Cancer Lett.  5: 199-204.

Stinson, S. F.; Loosli, C. G. (1979) The effect  of synthetic  smog  on  voluntary
     activity of  CD-I mice.  In: Animals  as  monitors of environmental pollu-
     tants:  symposium on  pathobiology  of environmental pollutants,  animal
     models and wildlife as monitors; 1977; Storrs,  CT. Washington, DC: National
     Academy of Sciences; pp. 233-240.

Stokinger, H. E.;  Coffin,  D.  L.  (1968)   Biological effects of air pollutants.
     In: Stern, A. C., ed. Air pollution v. 1. pp.  446-546.

Stokinger, H. E.; Wagner, W. D.; Wright, P-. G. (1956)  Studies on ozone toxicity.
     I.  Potentiating  effects  of exercise and tolerance development.  AMA Arch.
     Ind. Health 14:  158-162.

Stokinger,  H.  E.; Wagner, W. D.;  Dobrogorski,  0. J.  (1957)  Ozone toxicity
     studies. III.  Chronic injury to lungs  of  animals following exposure at
     low level. AMA Arch.  Ind. Health 16:  514-522.

Strawbridge,  H.  T. G.  (1960)  Chronic pulmonary  emphysema  (an  experimental
     study).  II.  Spontaneous  pulmonary emphysema  in rabbits.  Am. J.  Pathol.
     37: 309-331.

Suttorp, N.;  Simon,  L.  M.; (1982)  Lung  cell  oxidant  injury: enhancement of
     polymorphonuclear leukocyte-mediated  cytotoxicity in  lung cells  exposed to
     sustained ui vitro hyperoxia. J. Clin. Invest.  70: 342-350.

Suzuki, T.  (1976)  Studies  on  the acute  toxicity of 5-hydroxytryptamine in the
     rats  pre-exposed to ozone.  Bull.  Tokyo Med. Dent. Univ. 23:  217-226.

Swann,  H.  E., Jr.;  Balchum,  0. J.  (1966) Biological  effects of  urban air
     pollution. IV. Effects of acute smog  episodes on  respiration  of  guinea pigs.
     Arch. Environ. Health 12: 698-704.

Sweet,  F.;  Kao,  M.-S.;  tee, S.  D.;  Hagar, W. L.;  Sweet, W.  E.  (1980) Ozone
     selectively  inhibits  growth of human cancer cells. Science (Washington,
     DC) 209: 931-933.

Teige, B.; McManus, T. T.; Mudd, J.  B. (1974) Reaction of ozone  with  phosphati-
     dylchoMne liposomes and the  lytic effect of products on red  blood cells.
     Chen. Phys. Lipids 12: 153-171.


                                   9-292

-------
Tapper, J.  L. ; Weiss,  B.;  Cox,  C.  (1982) Microanalysis of ozone depression of
     motor activity. Toxicol. Appl. Pharmacol. 64:  317-326.

Tepper, J.  L.;  Weiss,  B.;  Wood, R. W.  (1983)  Behavioral indices of  ozone
     exposure. In:  Lee,  S.  D.;  Mustafa,  M.  G.;  Mehlman, M. A.,  eds.  Inter-
     national symposium  on  the  biomedical effects of ozone and related photo-
     chemical oxidants;  March 1982; Pinehurst,  NC.  Princeton, NJ: Princeton
     Scientific Publishers, Inc.; pp.  515-526. (Advances  in  modern environmen-
     tal toxicology: v. 5).

Thienes, C. H.; Skillen, R. G.;  Hoyt,  A.;  Bogen,  E.  (1965)  Effects of  ozone on
     experimental tuberculosis  and on natural pulmonary  infections in mice.
     Am. Ind. Hyg. Assoc. J.  26: 255-260.

Thomas, G.;  Renters,  J.  D.;  Ehrlich,  R.  (1979)  Effect of exposure to  PAN and
     ozone on susceptibility  to  chronic  bacterial  infection.  Research  Triangle
     Park,  NC: U.S.  Environmental   Protection Agency,  Health Effects Research
     Laboratory;  EPA  report  no.  EPA-600/1-79-001.  Available  from:  NTIS,
     Springfield, VA;  PB-292267.

Thomas, G.  B. ;  Renters,  J.  D.;  Ehrlich,  R.;  Gardner, D. E. (1981a) Effects of
     exposure to  peroxyacetyl nitrate on  susceptibility  to  acute and  chronic
     bacterial infection. J.  Toxicol.  Environ. Health  8:  559-574.

Thomas, G.  B. ;  Renters,  J.  D.;  Ehrlich,  R.;  Gardner, D. E. (1981b) Effects of
     exposure to  ozone on susceptibility to  experimental  tuberculosis.  Toxicol.
     Lett.  9: 11-17.

Thompson,  D.  J. ;  Lebowitz, M. ; Cassell  E.  J.  (1970)  Health  and  the  urban
     environment,  air  pollution, weather  and  the common  cold.  Am.  J.  Public
     Health 60: 731-739.

Tice,  R.  R. ; Bender,  M.  A.;  Ivett, J.  L.;  Drew, R.  T.  (1978) Cytogenetic
     effects  of inhaled  ozone.  Mutat.  Res. 58:  293-304.

Trams,  E. G.; Lauter,  C. J.;  Brown, E. A.  B.;  Young, 0.  (1972) Cerebral corti-
     cal metabolism after  chronic  exposure to  ozone. Arch.  Environ.  Health 24:
     153-159.

Tremer, H.  M.; Falk. H.  L.; Kotjn,  P.  (1959) Effect of air pollutants on cili-
     ated  mucus-secreting  epithelium.  JNCI J.  Natl. Cancer  Inst.  23: 979-994.

Tyson,,.£.._ A,.;. Lunan.,  K,.. D,;. Stephens,__ R.. J,.., (.1982)..Age-re 1 atedL _di fferences  in
     GSH-shuttle  enzymes in  N02-   or  03-exposed  rat  lungs.  Arch. Environ.
     Health 37: 167-176.

U.S. Department of  Hea/lth,  Education and Welfare. (1967) The health consequences
     of smoking.  Washington, DC: Public Health  Service; publication no.  1696.

U.S. Department of  Health,  Education and Welfare. (1969) The health consequences
     of smoking:  1969 supplement  to  the 1967 Public  Health Service review.
     Washington,  DC:  Public Health Service.
                                    9-293

-------
U.S. Environmental  Protection Agency.  (1978) Air  quality criteria for ozone
     and photochemical  oxidants.  Research Triangle Park, NC:  U.S.  Environmen-
     tal Protection Agency,  Environmental  Criteria and Assessment Office;
     EPA-600/8-78-004.  Available  'from:  NTIS,  Springfield, VA;  PB8Q-124753.

U.S. Environmental  Protection Agency.  (1982) Air quality criteria  for  particu-
     late matter  and sulfur oxides.  Research Triangle  Park,  NC:  U.S.  Environ-
     mental  Protection  Agency,  Environmental Criteria and Assessment  Office;
     EPA report  no. EPA-6QQ/8-82-029a,b,c.  Available from:  NTIS, Springfield,
     VA; PB84-156777.

Vaughan, T.  R.,  Jr.; Jennelle, L. F.; Lewis, T. R.  (1969) Long-term exposure
     to low  levels  of air pollutants: effects  on  pulmonary function  in  the
     beagle. Arch.  Environ.  Health 19: 45-50.

Vaughan, W.  J.;  Adamson,  G.  L.;  Lindgren, F.  T.;  Schooley,  J.  C. (1984) Serum
     lipid  and  lipoprotein  concentrations  following exposure to  ozone.  J.
     Environ. Pathol. Toxicol. Oncol.  5:  165-173.

Veninga, T.  S. (1967) Toxicity of ozone  in  comparison with ionizing radiation.
     Strahlentherapie 134: 469-477.

Veninga, T.  S.  (1970) Ozone-induced alterations in murine blood  and  liver.
     Presented  at:   Second  international  clean air congress;  December;
     Washington,  DC.  Washington, DC:  International Union of  Air  .Pollution
     Prevention Associations; paper no.  MB-1SE.

Veninga, T.  S.;  Lemstra,  W.  (1975)  Extrapulmonary  effects of ozone whether in
     the presence  of nitrogen  dioxide  or  not.  Int. Arch.  Arbeitsmed. 34:
     209-220.

Veninga, T.  S.;  Wagenaar, J.; Lemstra, W. (1981)  Distinct enzymatic responses
     in mice exposed to a range  of  low  doses of  ozone.  EHP Environ.  Health
     Perspect. 39:  153-157.

Verweij, H.; Van Steveninck, J.  (1980)  Effects of semicarbazide on oxidative
     processes in human red blood cell membranes.  Biochim.  Biophys. Acta 602:
     591-599.

Verweij, H., Van  Steveninck,  J. (1981) Protective  effects of semicarbazide  and
     p-aminobenzoic acid  against ozone toxicity. Biochem. Pharmacol. 30:  1033-
     1037.

Warshauer,   D.; Goldstein, E.; Hoeprich,   P.  D.;  Lippert, W.  (1974) Effect of
     vitamin E and  ozone on the pulmonary  antibacterial defense mechanisms.
     J. Clin. Med.  83: 228-240.

Watanabe, S.; Frank,  R.-; Yokoyama, E. (1973) Acute  effects of ozone on lungs
     of cats. I.   Functional.  Am.  Rev.  Respir. Dis.  108:  1141-1151.

Wayne,  L. G.;  Chambers,  L.  A.  (1968)  Biological  effects of urban air pollu-
     tion.   V.  A  study  of effects of  Los Angeles  atmosphere  on laboratory
     rodents. Arch.  Environ.  Health 16:  871-885.
                                   9-294

-------
Wegner, C. D. (1982) Characterization of dynamic respiratory mechanics by mea-
     suring  pulmonary and  respiratory  system  impedances  in adult  bonnet
     monkeys (Macaca  radiata):  including  the effects of long-term exposure to
     low-level   ozone  [dissertation]. Davis,  CA:  University of  California.
     Available  from:  University  Microfilms,  Ann  Arbor,  MI;  publication
     no. 82-27900.

Weibel, E.  R.  (1963)  Morphometry of the  human lung. New York,  NY: Academic
     Press, Inc.

Weiss,  B.;  Ferin,  J.; Merigan, W.; Stern, S.;  Cox,  C.  (1981) Modification of
     rat  operant  behavior  by  ozone exposure. Toxicol.  Appl.   Pharmacol.
     58: 244-251.

Wilkins,  E.  S.;  Fettissoff, P. (1981)  The  effect  of air pollutants  on  lung
     surfactant surface tension. J. Envi-ron.  Sci. Health A16: 477-491.

Williams,  P. S. ;  Calabrese, E.. J.; Moore, G.  S. (1983a) An evaluation of  the
     Dorset  sheep as  a predictive  animal  model for the response  of  G-6-PD
     deficient  human  erythrocytes  to a proposed systemic  toxic ozone inter-
     mediate. J. Environ. Sci. Health A18: 1-17.

Williams,  P. S.;  Calabrese, E. J.; Moore, G.  S. (1983b) The effect of methyl
     linoleate  hydroperoxide  (MLHP),  a possible toxic  intermediate of ozone,
     on human  normal  and glucose-6-phosphate dehydrogenase (G-6-PD)  deficient
     erythrocytes. J.  Environ. Sci. Health A18: 37-49.

Williams,  P.;  CaVabrese,  E. J.; Moore, G.  S. (1983c)  The effect  of  methyl
     oleate  hydroperoxide,  a  possible  toxic  ozone  intermediate, on human
     normal  and  glucose-6-phosphate  dehydrogenase-deficient  erythrocytes.
     Ecotoxicol. Environ. Saf. 7: 242-248.

Williams,  S. J.;  Charles, J.  M.; Menzel,  D. B. (1980)  Ozone induced altera-
     tions  in  phenol  red absorption from the rat  lung.  Toxicol. Lett. 6:
     213-219.

Wilmer, J.  W.  G.  M.;  Natarajan, A.  T.  (1981)  Induction of  sister-chromatid
     exchanges and chromosome  aberrations by Y~irradiated  nucleic  acid consti-
     tuents  in CHO cells. Mutat. Res. 88: 99-107.

Wilson, D.  W. ;  Plopper,  C. G.;  Dungworth,  D. L.  (1984) The response of the
     macaque tracheobronchial  epithelium  to  acute ozone injury.  A  quantitative
     ultrastructural  and autoradiographic study.  Am. J. Pathol.  116:  193-206.

Winterbourn, C. C. (1979) Protection by ascorbate against  acetylphenylhydrazine-
     induced Heinz body  formation in glucose-6-phosphate dehydrogenase deficient
     erythrocytes. Br. J. Haematol. 41: 245-252.

Witz,  G. ;  Amoruso, M. A.; Goldstein, B. D.  (1983)  Effect of ozone on alveolar
     macrophage  function:  membrane dynamic  properties.  In:  Lee, S. D.;
     Mustafa,  M.  G.;  Mehlman,  M.  A.,  eds.  International  symposium  on  the
     biomedical  effects  of ozone  and  related photochemical oxidants; March
     1982;  Pinehurst, NC.  Princeton,  NJ: Princeton Scientific Publishers,
     Inc.;  pp.  263-272,  (Advances in modern  environmental toxicology: v.  5).


                                    9-295

-------
Wolcott, J.  A.;  Zee,  Y. C.; Osebold,  J.  W.  (1982) Exposure to ozone reduces
     Influenza disease severity and alters  distribution of influenza viral
     antigens in murine lungs. Appl. Environ; Mlcrobiol. 44: 723-731.

Wong, D. S.;  Hochstein, P. (1981) The potentiation of ozone toxicity by thy-
     roxine. Res. Commun. Chem. Pathol. Pharmacol. 31: 483-492.

Wood, R,  W. (1979)  Behavioral evaluation of  sensory irritation evoked by
     ammonia. Toxicol. Appl. Pharmacol. 50:  157-162.

World Health Organization.  (1961) Chronic cor pulmonale: report of  an expert'
     committee.  Geneva,  Switzerland:  World  Health Organization. (Technical
     report series: no. 213).

Wright,  J.   L.; Lawson,  L. M.;  Pare,  P. D.; Wiggs,  B.  J,; Kennedy, S.; Hogg,  J.  C.
     (1983) Morphology of peripheral airways in current  smokers and  ex-smokers.
     Am. Rev. Respir. Dis.  127: 474-477.

Xintaras, C.;  Johnson,  B.   L.; Ulrich,  C.  E.;  Terrill,  R.  E.;  Sobecki,  M.  F.
     (1966)  Application of the  evoked response technique  in  air  pollution
     toxicology,  Toxicol. • Appl. Pharmacol. 8: 77-87.

Yilma,  T.;  Zee,  Y.  C.;  Osebold,  J.  W.  (1'979) Immunofluorescence determination
     of the pathogenesis  of infection with  influenza virus  in mice  following
     exposure to aerosolized virus.  J. Infect. Dis.  139: 458-464.

Yokoyama, E.  (1969)  A comparison of the  effects  of  S02, N02  and  03 on the
     pulmonary ventilation: guinea  pig  exposure  experiments.  Sangyo  Igaku
     11: 563-568.

Yokoyama, E.  (1972)  The effects of  03 on lung functions.  In:  Proceedings  of
     the science lecture  meeting  on air pollution caused  by  photochemical
     reaction; March;  Tokyo,  Japan.   Tokyo, Japan:  Japan Society  of  Air
     Pollution; pp. 123-143.

Yokoyama, E. (1973) The effects of  low-concentration  ozone  on  the lung pressure
     of  rabbits  and  rats.  Presented at: 46th  annual meeting of the Japan
     Society of  Industrial  Hygiene;  April; paper 223.

Yokoyama, E.  (1974)  On the maximal  expiratory  flow volume curve of rabbits
     exposed to  ozone.  Nipon Kyobu  Shikkan Gakkai  Zasshi 12: 556-561.

Yokoyama,  E.;  Frank,  R. (19/2)  Respiratory  uptake of ozone  in  dogs.  Arch
     Environ. Health 25: 132-138.

Yokoyama, E.;  Ichikawa, I. (1974)  Study  on  the biological  effects  of atmos-
     pheric pollutants  (FY 1972-1975). In:  Research  report  for funds of the
     Environmental  Agency  in  1974.  Tokyo, Japan:  Institute  of  Public Health,
     Department  of Industrial  Health;  pp. 16-1 - 16-6.

Yokoyama,  E.;  Ichikawa,  I.;  Nambu, Z.;  Kawai,  K.;  Kyono,  Y. J.  (1984)
     Respiratory effects  of intermittent exposure to ozone  of  rats. Environ.
     Res. 33: 271-283.
                                    9-296

-------
Younes, M.;  Siegers, C.-P.  (1980) Lipid  peroxidation as a consequence  of
     glutathione depletion  in  rat and mouse liver.  Res,  Commun.  Chem.  Pathol.
     Pharm. 27: 119-128.

Yu, C.  P.  (1975)  On equations of  gas  transport  in  the lung. Respir. Physio!.
     23: 257-266.

Zelac, R.  E.; Cromroy, H. L.; Bolch, W. E., Jr.;  Dunavant,  B.  G.;  Bevis,  H.  A.
     (1971a)  Inhaled ozone  as  a  mutagen.  I. Chromosome aberrations induced in
     Chinese  hamster lymphocytes.  Environ.  Res.  4:  262-282.

Zelac, R.  E.; Cromroy, H. L.; Bolch, W. E., Jr.;  Dunavant,  B.  G.;  Bevis,  H.  A.
     (1971b)  Inhaled ozone  as a  mutagen. II.  Effect on the  frequency of chromo-
     some aberrations  observed in irradiated  Chinese  hamsters.  Environ.  Res.
     4: 325-342.

Zitnik, L.  A.;  Schwartz, L. W. ;  McQuillen, N. K.;  Zee,  Y.  C.; Osebold, J.  W.
     (1978) Pulmonary changes  induced  by low-level  ozone: morphological obser-
     vations. J. Environ. Pathol.  Toxicol.  1:  365-376.

Zitting, A.;  Savolainen, H. (1980) Biochemical effects of subacuteformic acid
     vapor exposure. Res. Commun,  Chem. Pathol.  Pharmacol.  27: 157-162.
                                    9-297

-------

-------
            APPENDIX A:   GLOSSARY OF PULMONARY TERMS AND SYMBOLS*


Acetylcholine  (ACh):   A naturally  occurring substance  in  the body having
     important parasympathetic effects;  often used as a bronchoconstrictor.

Aerosol:  Solid particles  or liquid droplets that are dispersed or suspended
     in a gas, ranging in size from 10   to 10  micrometers (urn).

Air spaces:   All alveolar ducts,  alveolar sacs, and alveoli.  To be contrasted
     with AIRWAYS.

Airway  conductance  (Gaw):   Reciprocal  of airway resistance.   Gaw  = (I/Raw).

Airway  resistance  (Raw):   The (frictional) resistance to airflow afforded by
     the  airways between the airway opening  at  the mouth and the  alveoli.

Airways:  All passageways of the respiratory tract from mouth or nares down to
     and including respiratory bronchioles.  To be contrasted with AIR SPACES.

Allergen:  A material that, as a result of coming into contact with appropriate
     tissues of an animal body, induces a state of allergy or hypersensitivity;
     generally associated with idiosyncratic hypersensitivities.

Alveolar-arterial  oxygen pressure  difference [P(A-a)02]:  The difference  in
     partial pressure of 02 in the alveolar gas spaces and that  in the systemic
     arterial blood, measured in torr.

Alveolar-capillary  membrane:   A fine  membrane (0.2  to  0.4 pm) separating
     alveolus from capillary; composed of epithelial cells lining the alveolus,
     a  thin  layer  of connective tissue, and a layer of capillary endothelial
     cells.

Alveolar  carbon  dioxide pressure  (P/fiQ?)'  'Partial. pressure of carbon  dioxide
     in the air contained in the lung alveoli.

Alveolar  oxygen  partial  pressure (P/vOo^  Partial  pressure  of oxygen  in the
     air contained  in the alveoli or tne lungs.

Alveolar  septum  (pi.  septa):  A thin tissue  partition between two adjacent
     pulmonary  alveoli,  consisting of a close-meshed capillary network and
     interstitium  covered  on  both  surfaces  by  alveolar epithelial cells.
 ^References:   Bartels, H.; Dejours,  P.;  Kellogg,  R. H.; Mead, J.  (1973) Glossary
               on  respiration  and gas  exchange.  J. Appl.  Physio!.  34:  549-558.

               American College  of  Chest  Physicians - American Thoracic Society
               (1975)  Pulmonary  terms and symbols:  a  report of the ACCP-ATS
               Joint  Committee on pulmonary nomenclature.   Chest 67:  583-593.
                                     A-l

-------
Alveolitis:  (interstitial pneumonia):  Inflammation of the lung distal to the
     terminal non-respiratory  bronchiole.   Unless  otherwise indicated, it is
     assumed that the condition is diffuse.  Arbitrarily, the term is not used
     to refer to  exudate in air spaces resulting from bacterial infection of
     the lung.

Alveolus:   Hexagonal  or spherical  air cells of  the  lungs.   The majority of
     alveoli arise  from the alveolar ducts which are lined with the alveoli.
     An alveolus  is an  ultimate respiratory unit where the  gas  exchange takes
     place.

Anatomical dead space (V~    .):   Volume of the conducting airways down to the
     level where, during^ airbreathing,  gas exchange with blood can occur, a
     region probably situated at the entrance of the alveolar ducts.

Arterial oxygen saturation  (Sa02):   Percent saturation of dissolved oxygen  in
     arterial blood.

Arterial partial  pressure of  carbon dioxide (PaCOp):   Partial pressure  of
     dissolved carbon dioxide in arterial blood.

Arterial partial  pressure of oxygen  (PaOp):   Partial  pressure of dissolved
     oxygen in arterial blood.

Asthma:  A disease characterized by an increased responsiveness of the airways
     to various stimuli  and manifested by slowing of forced expiration which
     changes in severity either .spontaneously  or as  a  result of therapy.  The
     term asthma  may be modified by  words  or phrases indicating its etiology,
     factors provoking attacks, or its duration.

Atelectasis:  State of collapse  of  air  spaces  with  elimination of the gas
     phase.

ATPS condition (ATPS):  Ambient temperature and pressure, saturated with water
     vapor.  These are the conditions existing in a water spirometer.

Atropine:   A poisonous  white crystalline alkaloid, Cy/H^NO-,  from belladonna  ,
     and related  plants,  used  to  relieve spasms of smodtn muscles.   It is an
     anticholinergic agent.

Breathing  pattern:   A general term  designating the  characteristics of the
     ventilatory  activity,  e.g.,  tidal volume,  frequency of  breathing,  and
     shape of the volume time curve.

Breuer-Hering reflexes (Hering-Breuer reflexes):  Ventilatory reflexes originat-
     ing in the lungs.  The reflex arcs are formed by the pulmonary mechanore-
     ceptors, the vagal afferent fibers,  the respiratory centers,  the medullo-
     spinal pathway,  the motor neurons,  and the  respiratory muscles.  The af-
     ferent link informs the respiratory centers of the volume state or of the
     rate of change of  volume of the lungs.  Three types of Breuer-Hering re-
     flexes have been described:   1) an inflation reflex in which lung inflation
     tends  to  inhibit inspiration  and stimulate expiration; 2) a deflation
     reflex in which lung deflation  tends  to inhibit expiration and stimulate
     inspiration;  and 3)  a  "paradoxical  reflex," described  but  largely disre-
     garded by  Breuer and Hering, in which sudden inflation may stimulate
     inspiratory muscles.
                                    A-2

-------
Bronchiole:   One of  the  finer subdivisions of the airways, less than 1 mm in
     diameter, and having no cartilage In Its wall.

Bronchiolitis:  Inflammation of the bronchioles which may be acute or chronic.
     If the etiology is known, it should be stated.   If permanent occlusion of
     the lumens is  present,  the term bronchiolitis obliterans  may  be used.

Bronchitis:   A non-neoplastic disorder of structure or function of the bronchi
     resulting from infectious or noninfectious irritation.  The term bronchitis
     should be modified by appropriate words or phrases to indicate its etiol-
     °fly> its chronicity,  the presence of associated airways dysfunction, or
     type of  anatomic  change.   The term  chronic bronchitis, when  unqualified,
     refers to a  condition associated with prolonged exposure to nonspecific
     bronchial irritants and  accompanied by  mucous hypersecretion and  certain
     structural  alterations  in the  bronchi.   Anatomic changes may  include
     hypertrophy of  the mucous-secreting apparatus and epithelial metaplasia,
     as  well  as  more  classic evidences  of inflammation.  In  epidemiologic
     studies, the presence, of cough  -or sputum production on most days for at
     least three months of the year has sometimes been accepted as a criterion
     for the diagnosis.

Bronchoconstrictor:   An agent that causes a reduction in the caliber (diame-
     ter) of airways.

Bronchodilator:   An agent that causes an increase in the caliber  (diameter) of
     airways.

Bronchus:  One of the  subdivisions of the trachea serving to convey air to and
     from the lungs.   The  trachea divides into right  and left main bronchi
     which in turn form lobar, segmental, and subsegmental bronchi.

BTPS conditions (BTPS):  Body temperature, barometric  pressure,  and  saturated
     with water vapor.  These are  the conditions  existing in  the  gas  phase of
     the lungs.   For man the  normal temperature is taken as 37°C, the pressure
     as the barometric pressure, and the partial pressure of water vapor as 47
     torr.

Carbachol:  A parasympathetic stimulant  (carbamoyjcholine chloride, CgH-jj-ClNpOp)
     that produces constriction of the bronchial smooth muscles.

Carbon dioxide production  (VC02):  Rate of carbon dioxide production by organ-
     isms, tissues, or cells.  Common units:  ml C02 (STPD)/kg*min.

Carbon monoxide (CO):  An  odorless,  colorless,  toxic  gas formed by  incomplete
     combustion,  with, a strong affinity for hemoglobin and  cytochrome;  it
     reduces  oxygen absorption  capacity, transport, and  utilization.

Carboxyhemoglobin  (COHb):   Hemoglobin in which the-iron is associated with
     carbon  monoxide.  The affinity  of  hemoglobin for  CO is  about  300 times
     greater  than for  0^-
                                    A-3

-------
Chronic obstructive  lung disease (COLD):  This  term  refers  to diseases of
     uncertain etiology  characterized by persistent slowing of airflow during
     forced expiration.  It  is recommended that  a more specific term, such as
     chronic obstructive bronchitis or chronic obstructive emphysema, be used
     whenever possible.  Synonymous with chronic obstructive  pulmonary disease
     (COPD).

Closing capacity  (CC):   Closing  volume  plus residual  volume,  often expressed
     as a ratio of TLC, i.e.  (CC/TLCX).

Closing volume (CV):   The  volume exhaled after the expired gas concentration
     is inflected  from an alveolar  plateau during a  controlled  breathing
     maneuver.   Since  the  value  obtained is dependent on  the  specific  test
     technique, the  method  used  must be designated in the text,  and when
     necessary, specified  by a qualifying symbol.  Closing volume  is often
     expressed as a ratio of the VC,  i.e. (CV/VC%).

Collateral resistance  (R   ,-,):   Resistance to flow through indirect pathways.
     See COLLATERAL VENTItWlON and RESISTANCE.

Collateral  ventilation:  Ventilation  of air  spaces  via indirect pathways,
     e.g., through pores in  alveolar septa, or anastomosing respiratory bron-
     chioles.

Compliance (C,,C  .):   A  measure  of distensibility.   Pulmonary compliance is
     given by  the  slope  of a static volume-pressure curve at a point, or the
     linear approximation  of a nearly  straight portion of such a curve, ex-
     pressed in liters/cm  HpO  or ml/cm HpO.  Since the static volume-pressure
     characteristics of  lungs  are nonlinear (static compliance decreases as
     lung volume  increases)  and  vary according to the previous volume history
     (static compliance  at a given volume increases  immediately  after  full
     inflation and decreases following  deflation),  careful specification of
     the conditions of measurement are  necessary.  Absolute values also depend
     on organ size.  See also DYNAMIC COMPLIANCE.

Conductance (G):   The reciprocal  of  RESISTANCE.  See  AIRWAY  CONDUCTANCE.

Diffusing capacity of  the  lung (D. ,  D.Op, DiCO?* D.CO):  Amount of gas (0^,
     CO, COg)  commonly expressed as  mr gas cSTTO) diffusing  between alveorar
     gas ana pulmonary capillary blood  per torr mean gas pressure difference
     per min,  i.e.,  ml QgAmin-torr).   Synonymous with transfer  factor and
     diffusion factor.

Dynamic compliance (Cjyn):   The  ratio  of the tidal volume to  the change in
     intrapleural pressure between the points of zero flow at the extremes of
     tidal volume  in  liters/cm H,0 or ml/cm H^O.  Since at the points of zero
     airflow at the  extremes of "tidal  volume, volume acceleration is usually
     other than zero,  and  since, particularly in abnormal states,  flow may
     still be  taking  place within lungs between regions which are exchanging
     volume, dynamic  compliance  may differ from  static compliance, the latter
     pertaining to condition of  zero  volume acceleration and  zero  gas  flow
     throughout the lungs.    In normal  lungs at ordinary volumes and respiratory
     frequencies, static and dynamic compliance are the same.
                                    A-4

-------
Elastance (E):  The  reciprocal  of  COMPLIANCE;  expressed  in  cm  H90/liter  or  cm
     H20/ml.                                                     *

Electrocardiogram (ECG,  EKG):   The graphic record of the electrical currents
     that are associated with the heart's contraction and relaxation.

Emphysema:   A  condition of  the lung  characterized  by abnormal,  permanent
     enlargement of  airspaces  distal  to the terminal bronchiole, accompanied
     by the  destruction of their walls, and without obvious fibrosis.

Expiratory reserve volume  (ERV):   The maximal volume of air exhaled from the
     end-expiratory level.

FEV./FVC:  A ratio of timed (t = 0.5, 1,  2,  3 s) forced expiratory volume
     (FEV.)  to  forced vital  capacity (FVC).  The ratio is often expressed  in
     percent 100 x FEVt/FVC.  It is an index of airway obstruction.

Flow volume curve:   Graph  of instantaneous  forced expiratory flow  recorded  at
     the mouth,  against corresponding lung volume.  When  recorded over the
     full vital capacity,  the  curve  includes  maximum expiratory flow rates  at
     all lung  volumes  in  the  VC range and is  called  a  maximum expiratory
     flow-volume curve  (MEFV).   A  partial  expiratory flow-volume curve  (PEFV)
     is one which describes maximum expiratory flow rate over a portion of  the
     vital  capacity only.

Forced  expiratory  flow (FEFx):  Related to some  portion of the FVC curve.
     Modifiers refer to the amount of the FVC already exhaled when the measure-
     ment is made.   For example:

                 = instantaneous forced expiratory flow after 75%
                   of the  FVC has been exhaled.

          FEF9nn 19nn = mean forced expiratory flow between 200 ml
             ^uu-i^uu   gnd 1200 ml Qf the pvc (formerly called the
                        maximum expiratory flow rate (MEFR).

          FEF?r -jf-o/ = mean forced expiratory flow during the middle
             "~/3A   half of the FVC [formerly called the maximum
                      mid-expiratory flow rate (MMFR)].

          FEF    = the maximal   forced expiratory  flow achieved  during
             max   an FVC.

Forced expiratory volume (FEV):  Denotes the volume of gas which is exhaled in
     a  given  time  interval during the execution  of  a  forced vital  capacity.
     Conventionally, the times  used are 0.5, 0.75, or 1 sec, symbolized  FEVg ,-,
     FEVn 7r,  FEV-, n.   These values are often  expressed  as a  percent  of the'
     forCed5vitaTLcaJpacity, e.g. (FEV-j^ Q/VC) X 100.

Forced  inspiratory vital capacity  (FIVC):   The maximal volume  of  air inspired
     with a  maximally forced effort from  a position of  maximal expiration.

Forced vital capacity (FVC):  Vital capacity performed with a maximally  forced
     expiratory effort.


                                    A-5

-------
Functional residual  capacity  (FRC):  The  sum of  RV and ERV  (the volume of air
     remaining in the  lungs at the end-expiratory position).  The method of
     measurement should be indicated as with RV,

Gas exchange:  Movement of oxygen from the alveoli into the pulmonary capillary
     blood as carbon dioxide  enters the  alveoli  from the blood.  In broader
     terms, the exchange of gases between alveoli and lung capillaries.

Gas exchange ratio (R):  See RESPIRATORY QUOTIENT.

Gas trapping:  Trapping  of gas behind small airways that were opened during
     inspiration but closed during forceful expiration.   It is a volume differ-
     ence between FVC and VC.

Hematoerit (Hct):  The  percentage of the volume  of red blood cells in whole
     blood.

Hemoglobin (Hb):  A  hemoprotein  naturally occurring in most vertebrate blood,
     consisting of four  polypeptide chains (the  globulin)  to each  of  which
     there is attached  a heroe group.  The heme is made of four pyrrole rings
     and a divalent  iron  (Fe   -protoporphyrin) which combines reversibly with
     molecular oxygen.

Histamine:  A depressor  amine  derived  from the amino acid histidine and found
     in all body tissues, with the highest concentration in the lung;  a powerful
     stimulant of gastric secretion, a constrictor of bronchial  smooth  muscle,
     and a vasodilator that causes a fall in blood pressure.

Hypoxemia:   A  state in  which the oxygen  pressure and/or  concentration  in
     arterial and/or venous blood is lower than its normal  value at sea level.
     Normal oxygen pressures  at  sea level are 85-100 torr  in arterial blood
     and 37-44 torr  in mixed  venous  blood.  In adult humans the normal oxygen
     concentration is 17-23 ml 02/100 ml arterial blood; in mixed venous  blood
     at rest it is 13-18 ml 02/100 ml blood.

Hypoxia:  Any state  in which  the oxygen  in the lung, blood, and/or tissues  is
     abnormally low  compared with that of normal  resting man breathing air  at
     sea level.   If  the  PQ2 is  low  in the environment,  whether because of
     decreased barometric  pressure  or  decreased  fractional concentration of
     Og, the condition is termed environmental  hypoxia.   Hypoxia when referring
     to the  blood is termed hypoxemia.  Tissues  are said to be hypoxic when
     their P02 is low, even if there is no arterial hypoxemia, as in "stagnant
     hypoxia" which  occurs  when  the local circulation is low compared  to the
     local metabolism.

Inspiratory capacity (1C):  The sum of IRV and TV.

Inspiratory  reserve  volume  (IRV):   The maximal volume of air  inhaled from the
     end-inspiratory level.

Inspiratory  vital capacity  (IVC):   The maximum volume of air  inhaled from the
     point of maximum expiration.
                                    A-6

-------
Kilogram-meter/min (kg-m/min):  The work performed each min to move a mass of 1
     kg through  a vertical  distance of 1 m against  the  force of gravity.
     Synonymous with kilopond-meter/min.

Lung volume  (V,):   Actual  volume of the  lung,  including  the volume of the
     conducting airways.

Maximal aerobic  capacity  (max VO^):   The rate  of  oxygen  uptake by the body
     during repetitive  maximal  respiratory  effort.  Synonymous with maximal
     oxygen consumption.

Maximum breathing capacity (MBC):  Maximal volume of air which can be breathed
     per minute  by  a subject breathing as quickly and as deeply as possible.
     This tiring lung function test is usually limited to 12-20 sec, but given
     in liters (BTPS)/min.  Synonymous with maximum voluntary ventilation (MVV).

Maximum expiratory  flow (V      ):   Forced  expiratory flow, related to the
     total lung capacity or tne actual volume of the lung at which the measure-
     ment is  made.   Modifiers refer to the  amount of lung volume remaining
     when the measurement is made.  For example:

          V    -.cw - instantaneous forced expiratory flow when the
           max /a*   lung 1g flt ?5% Qf Us TLC>

          V    ~ n = instantaneous forced expiratory flow when the
           max j.u        volume is 3-0 -|-jters
Maximum expiratory flow rate (MEFR):  Synonymous with ^^ 200-1200"

Maximum mid-expiratory  flow rate (MMFR or MMEF):  Synonymous with ^05-75%-

Maximum ventilation (max V>):  The volume of air breathed in one minute during
     repetitive maximal respiratory effort.  Synonymous with maximum ventilatory
     minute volume.

Maximum voluntary  ventilation  (MVV):   The  volume of  air breathed by  a  subject
     during voluntary  maximum  hyperventilation lasting a specific period of
     time.  Synonymous with maximum breathing capacity  (MBC).

Methemoglobin  (MetHb):   Hemoglobin in  which iron is  in the ferric state.
     Because the iron, is oxidized, methemoglobin is  incapable of oxygen trans-
     port.  Methemoglobins  are formed by various drugs  and occur under pathol-
     ogical conditions.   Many methods  for hemoglobin  measurements  utilize
     methemoglobin (chlorhemiglobin, cyanhemiglobin).

Minute  ventilation (Vp):   Volume  of  air breathed in one minute.   It is a
     product of  tidal  volume (VT)  and breathing frequency  (fR).   See VENTILA-
     TION.                      '         .                    °

Minute volume:  Synonymous  with minute ventilation.

Mucociliary transport:   The process by which mucus is  transported,  by  ciliary
     action, from  the lungs.
                                    A-7

-------
Mucus:  The clear, viscid  secretion  of mucous membranes, consisting of mucin,
     epithelial cells,  leukocytes,  and  various  inorganic salts suspended in
     water.

Nasopharyngeal:   Relating  to the nose or the nasal  cavity and the pharynx
     (throat).

Nitrogen oxides:  Compounds of N and 0 in ambient air; i.e., nitric oxide (NO)
     and others with  a higher oxidation state of N, of which N02 is the most
     important toxicologically.

Nitrogen washout  (AN-,  dN«):   The curve obtained by  plotting  the  fractional
     concentration of N? "in  expired alveolar gas vs. time,  for  a subject
     switched from breattiing ambient air to an inspired mixture of pure Op.   A
     progressive  decrease  of N« concentration ensues which  may be analyzed
     into  two  or  more exponential components.  Normally, after 4 min of pure
     0? breathing the fractional N7 concentration in expired alveolar gas is
     dSwn to less than 2%.

Normoxia:  A state in which the  ambient oxygen pressure is approximately 150 ±
     10  torr  (i.e.,   the  partial pressure  of oxygen  in  air  at sea level).

Oxidant:  A chemical  compound that has the ability to remove, accept, or share
     electrons from another chemical species, thereby oxidizing it.

Oxygen  consumption (VO?J  Q0?):   Rate of oxygen uptake of organisms, tissues,
     or cells.  Common unitf:   ml  02 (STPD)/(kg-min)  or  ml  02  (STPD)/(kg-hr).
     For whole organisms the oxygen Consumption is commonly expressed per unit
     surface  area or, some power of  the  body weight.   For tissue samples or
     isolated cells QQ2 = pi Op/hr per mg dry weight.

Oxygen  saturation (S0?):   The  amount of oxygen  combined with hemoglobin,
     expressed  as a  percentage of the oxygen capacity of that  hemoglobin.   In
     arterial blood,  SaO~.

Oxygen  uptake  ($02):   Amount of oxygen taken up by  the body  from  the environ-
     ment, by  the blood from the alveolar gas, or by an organ or tissue from
     the blood.   When this amount of oxygen  is expressed per unit of time one
     deals with an  "oxygen uptake  rate."  "Oxygen  consumption" refers more
     specifically to  the  oxygen uptake rate  by all  tissues  of  the body  and  is
     equal to  the oxygen uptake rate of the organism only when the $„ stores
     are constant.

Particulates:   Fine solid particles  such as  dust, smoke, fumes, or smog, found
     in the air or in emissions.

Pathogen:  Any virus,  microorganism,  or etiologic  agent  causing disease.

Peak  expiratory flow (PEF):   The highest forced expiratory  flow measured with
     a  peak flow meter.

Peroxyacetyl  nitrate  (PAN):   Pollutant created by action  of UV component of
     sunlight on hydrocarbons and NO  in the air;  an ingredient of photochem-
     ical  smog.


                                     A-8

-------
Physiological  dead  space  (Vn):   Calculated  volume which accounts  for  the
     difference between the pressures  of  C02 in expired and alveolar gas (or
     arterial  blood).  Physiological dead  space reflects the combination of
     anatomical  dead space and  alveolar dead space,  the volume of the latter
     increasing  with  the  importance   of  the  nonuniformity   of  the
     ventilation/perfusion ratio in the lung.

Plethysmograph:   A  rigid  chamber placed around  a  living structure for  the
     purpose of measuring  changes in the volume of the structure.   In respira-
     tory measurements, the entire body is ordinarily enclosed  ("body plethys-
     mograph") and  the plethysmograph  is used  to measure changes in volume of
     gas  in  the system produced 1) by solution  and  volatilization (e.g.,
     uptake of  foreign gases  into  the  blood),  2)  by  changes in pressure or
     temperature (e.g., gas compression in the lungs,  expansion of gas  upon
     passing into the warm, moist lungs),  or 3) by breathing through a tube to
     the  outside.   Three  types  of plethysmograph are used:  a) pressure, b)
     volume, and c)  pressure-volume.   In type  a, the  body chambers  have  fixed
     volumes and  volume  changes  are  measured  in  terms  of  pressure change
     secondary to gas  compression  (inside  the  chamber,  outside  the  body).  In
     type b, the body  chambers  serve essentially as conduits between  the body
     surface and devices (spirometers or integrating flowmeters) which measure
     gas  displacements.   Type c combines  a  and  b  by  appropriate summing of
     chamber pressure and volume displacements.

Pneumotachograph:   A device for measuring instantaneous gas flow  rates in
     breathing  by recording the pressure drop across a fixed flow resistance
     of  known pressure-flow characteristics, commonly connected to  the airway
     by  means  of  a  mouthpiece,   face mask,  or cannula.   The flow resistance
     usually consists  either of parallel  capillary tubes  (Fleisch  type)  or of
     fine-meshed screen (Silverman-Lilly type).

Pulmonary alveolar  proteinosis:  A chronic or recurrent disease characterized
     by  the filling of alveoli  with an insoluble exudate,  usually poor in
     cells, rich in  lipids and proteins, and accompanied by minimal histologic
     alteration of  the alveolar  walls.

Pulmonary edema:  An accumulation of  excessive  amounts  of  fluid  in the  lung
     extravascular  tissue and air  spaces.

Pulmonary  emphysema:  An  abnormal, permanent  enlargement of the air spaces
     distal to  the  terminal nonrespiratory bronchiole,  accompanied by destructive
     changes  of the alveolar  walls and without obvious fibrosis.   The  term
     emphysema  may  be  modified  by words or  phrases to  indicate its etiology,
     its anatomic subtype, or any  associated airways  dysfunction.

Residual  volume  (RV):  That volume  of  air  remaining in  the  lungs after maximal
     exhalation.  The  method  of measurement should be  indicated  in the  text
     or,  when necessary,  by appropriate qualifying symbols.
                                     A-9

-------
Resistance flow (R):   The  ratio of the flow-resistive components of pressure
     to simultaneous flow,  in cm hLO/liter per sec.   Flow-resistive components
     of pressure are obtained by subtracting any elastic or inertia! components,
     proportional  respectively  to  volume  and volume acceleration.   Most flow
     resistances in  the  respiratory system are nonlinear,  varying  with the
     magnitude and direction of flow, with lung volume and lung volume history,
     and possibly with volume acceleration.  Accordingly, careful specification
     of the  conditions of  measurement is  necessary;  see AIRWAY RESISTANCE,
     TISSUE  RESISTANCE,  TOTAL  PULMONARY  RESISTANCE,  COLLATERAL RESISTANCE.

Respiratory  cycle:   A respiratory  cycle  is  constituted by the  inspiration
     followed by the expiration of a given volume of gas, called tidal volume.
     The duration of the respiratory cycle is the respiratory or ventilatory
     period, whose reciprocal is the ventilatory frequency.

Respiratory exchange ratio:  See RESPIRATORY QUOTIENT.

Respiratory  frequency  (fp):   The number of breathing  cycles per unit  of time.
     Synonymous with breathing frequency (fg).

Respiratory  quotient  (RQ,  R):   Quotient of the  volume of C0? produced divided
     by the volume of 0,, consumed by an organism, an, organ, or a tissue during
     a given  period  of Time.   Respiratory  quotients are  measured by comparing
     the composition of an incoming and an outgoing medium, e.g., inspired and
     expired gas,  inspired gas and alveolar gas, or arterial and venous blood.
     Sometimes the phrase  "respiratory exchange ratio"  is  used to  designate
     the  ratio  of C02 output to  the Op uptake by  the  lungs,  "respiratory
     quotient" beirtg restricted to the actual  metabolic CO,,  output  and 02
     uptake  by the  tissues.   With this definition, respiratory quotient and
     respiratory exchange ratio are identical in the steady state,  a condition
     which implies constancy of the 02 and COp stores.

Shunt;  Vascular connection  between circulatory pathways so that venous blood
     is diverted  into vessels  containing  arterialized  Wood  (right-to-left
     shunt,  venous admixture) or vice  versa  (left-to-right shunt).  Right-to-
     left shunt within the lung, heart, or large vessels due to malformations
     are more important  in respiratory physiology.    Flow  from  left to  right
     through a shunt should be marked with a negative sign.

Specific airway conductance  (SGaw):   Airway conductance divided by the lung
     volume  at which it was  measured, i.e.,  normalized airway conductance.
     SGaw = Gaw/TGV.

Specific airway resistance (SRaw):   Airway resistance multiplied by the volume at
     which it was measured.  SRaw = Raw x TGV.

Spirograph:   Mechanical  device, including bellows  or other scaled,  moving
     part, which collects  and stores gases and  provides  a graphical record  of
     volume changes.  See BREATHING PATTERN,  RESPIRATORY CYCLE.

Spirometer:  An apparatus similar to a spirograph but without recording facil-
     ity.
                                    A-10

-------
Static lung compliance (C,  t):   Lung compliance measured at zero flow (breath-
     holding) over linear portion of the volume-pressure curve above FRC.   See
     COMPLIANCE.

Static transpulmonary pressure  (P  .):   Transpulmonary pressure measured at a
     specified lung volume; e.g., TS+JLC is static recoil pressure measured at
     TLC (maximum recoil  pressure).

Sulfur dioxide (SCL):   Colorless gas with pungent odor, released primarily from
     burning of fossil fuels, such as coal, containing sulfur.

STPD conditions  (STPD):   Standard  temperature and pressure,  dry.  These are
     the conditions of a volume of gas  at  0°C,  at 760 torr, without water
     vapor.  A STPD volume of a  given gas contains a  known  number of moles of
     that gas.

Surfactant,  pulmonary:   Protein-phospholipid  (mainly  dipalmitoyl  lecithin)
     complex which lines alveoli (and possibly small  airways) and accounts for
     the low surface tension which makes air space (and airway) patency possible
     at low transpulmonary pressures.

Synergism:    A  relationship  in  which the combined  action  or effect  of two  or
     more  components  is  greater than the sum  of effects  when the components
     act separately.

Thoracic gas  volume  (TGV):   Volume of  communicating  and  trapped gas in the
     lungs  measured  by  body plethysmography at specific lung volumes.   In
     normal  subjects, TGV  determined at end expiratory  level  corresponds  to
     FRC.

Tidal volume (TV):   That volume of  air inhaled or exhaled with each breath
     during  quiet breathing, used only to  indicate  a subdivision  of lung
     volume.   When tidal volume is  used in gas exchange formulations, the
     symbol  VT should be used.

Tissue  resistance (R+4):   Frictional  resistance of the pulmonary and thoracic
     tissues.
                                                      2
Torr:   A  unit of pressure equal to  1,333.22 dynes/cm  or 1.33322 millibars.
     The torr  is  equal to the pressure  required  to support  a  column of mercury
     1 mm  high when the  mercury' is of standard density and  subjected to standard
     acceleration.  These standard conditions  are met at  0°C  and 45° latitude,
     where the acceleration of gravity  is 980.6  cm/sec  .  In  reading a mercury
     oarometer at other  temperatures and latitudes, corrections, which commonly
     exceed 2 torr,  must be introduced for these terms and  for the thermal
     expansion  of the measuring scale  used.   The torr is synonymous with
     pressure  unit mm Hg.

Total  lung capacity (TLC):   The sum of all  volume compartments  or the volume
     of air in the lungs after maximal  inspiration.  The  method  of measurement
     should be indicated, as with  RV.
                                    A-11

-------
Total pulmonary resistance (R,):  Resistance measured by relating flow-dependent
     transpulmonary pressure  xo airflow at the  mouth.   Represents  the total
     (frictional)  resistance  of the lung tissue  (R,.) and the airways  (Raw).

     RL = Raw+Rtr

Trachea:  Commonly known as the windpipe; a cartilaginous air tube extending
     from the larynx (voice box) into the thorax  (chest) where it divides into
     left and right branches.

Transpulmonary  pressure (P,):   Pressure  difference between  airway opening
     (mouth, nares,  or cannula opening) and the  visceral pleura! surface,  in
     cm HpO.  Transpulmonary in the sense  used includes  extrapulmonary struc-
     tures,  e.g.,  trachea  and  extrathoracic airways.   This  usage  has come
     about  for  want of an anatomic  term which  includes  all  of the airways and
     the lungs together.

Ventilation:  Physiological process  by which gas  is  renewed in the  lungs.   The
     word ventilation  sometimes designates  ventilatory  flow rate (or ventila-
     tory minute  volume)  which is  the product  of  the  tidal  volume by the
     ventilatory frequency.   Conditions are usually indicated as modifiers;
     i.e.,

               VV  = Expired volume per minute  (BTPS),
                .    and
               V,-  =, Inspired  volume  per minute  (BTPS).

     Ventilation is often  referred to as "total  ventilation"  to  distinguish it
     from "alveolar ventilation" (see VENTILATION, ALVEOLAR).

Ventilation, alveolar  (V«):   Physiological  process  by which  alveolar gas is
     completely  removed and replaced with  fresh gas. Alveolar ventilation is
     less than total ventilation because when  a  tidal volume  of  gas leaves  the
     alveolar spaces,'the last part does not  get expelled from the body but
     occupies the  dead  space,  to  be reinspired with the  next  inspiration.
     Thus the volume of alveolar gas actually  expelled  completely  is  equal  to
     the tidal volume  minus the volume of the  dead space.  This  truly  complete
     expiration volume times  the ventilatory frequency constitutes  the alveolar
     ventilation.

Ventilation, dead-space  (Vr,):  Ventilation  per minute of the  physiologic dead
     space  (wasted ventilation), BTPS,  defined  by  the  following equation:

          VD = VE(PaCD2  -  PEC02)/(PaC02 - PjC02)

Ventilation/perfusion  ratio  (V./Q):   Ratio  of  the alveolar ventilation to the
     blood  perfusion volume flow through the pulmonary parenchyma.  This  ratio
     is a fundamental  determinant of the 0~ and  C0? pressure of the alveolar
     gas and of the  end-capillary  blood.   Throughout  the lungs the  local
     ventilation/perfusion  ratios  vary,  and consequently the local  alveolar
     gas and end-capillary blood compositions  also vary.

Vital  capacity  (VC):   The maximum volume of air exhaled from the  point of
     maximum inspiration.
     ««»»wr ttamne omot «»—«5
                                    A-12

-------

-------
Environmental Protection
Agency
Information
Cincinnati OH 45268
      BULK RATE
POSTAGE & FEES PAID
         EPA
   PERMIT No. G-35
Official  Business
Penalty for Private Use, $300
                                                                                                          Special Fourth-Class Rate
                                                                                                                    Book
                                                               Please make all necessary changes on the above label.
                                                               detach or copy, and return to the address in the upper
                                                               left-hand corner.

                                                               If you do not wish to receive these reports CHECK HERE D;
                                                               detach, or copy this cover, and return to the address in the
                                                               upper left-hand corner.
                                                             EPA/600/8-84/020dF

-------