>EPA
              United States
              Environmental Protection
              Agency
             Office of Health and
             Environmental Assessmei..
             Washington DC 20460
600884020A4 .,,i
    External Review Draft c,\
              Research and Development
Air Quality Criteria    Review
for Ozone and  Other  Draft
Photochemical
Oxidants
                                              (Do Not
                                              Cite or Quote)
              Volume  IV of V
                           NOTICE

              This document is a preliminary draft. It has not been formally
              released by EPA and should not at this stage be construed to
              represent Agency policy. It is being circulated for comment on its
              technical accuracy and policy implications.

-------
                                 EPA-600/8-84-020A
Do Not                                   June 1984
Cite or Quote                      External Review Draft
         Air Quality Criteria
        for Ozone and  Other
     Photochemical  Oxidants

                 Volume  IV
                       NOTICE

This document is a preliminary draft. It has not been formally released by EPA and should not at
this stage be construed to represent Agency policy. It is being circulated for comment on its
technical accuracy and policy implications.
          Environmental Criteria and Assessment Office
          Office of Health and Environmental Assessment
             Office of Research and Development
             U.S. Environmental Protection Agency
              Research Triangle Park, N.C. 27711
                      U.S. Environmental Protection Agency
                      Region V, Library
                      230 South Dearborn Street
                      Chicago, Illinois 60604

-------
                                     NOTICE

             Mention of trade names or commercial products does not constitute
             endorsement or recommendation for use.
Environmental Fraction Agen«

-------
                                   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 1983 and early 1984.

     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.  Separate chapters are 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 proper-
ties, chemistry, 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 appreisal of effects of ozone and other oxidants; effects
observed in controlled human exposures; effects observed in field and epidemio-
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.
                                      m
0190LG/B                                                              May 1984

-------
                              CONTENTS


                                                                      Page

VOLUME I
  Chapter 1.    Summary and Conclusions 	      1-1

VOLUME II
  Chapter 2.    Introduction 	      2-1
  Chapter 3.    Precursors to Ozone and Other Photochemical
               Oxidants 	      3-1
  Chapter 4.    Chemical and Physical  Processes in the Formation
               and Occurrence of Ozone and Other Photochemical
               Oxidants 	      4-1
  Chapter 5.    Properties, Chemistry, and Measurement of Ozone
               and Other Photochemical Oxidants 	      5-1
  Chapter 6.    Concentrations of Ozone and Other Photochemical
               Oxidants in Ambient Air 	      6-2

VOLUME III
  Chapter 7.    Effects of Ozone and Other Photochemical Oxidants
               on Vegetation 	      7-1
  Chapter 8.    Effects of Ozone and Other Photochemical Oxidants
               on Natural and Agroecosystems 	      8-1
  Chapter 9.    Effects of Ozone and Other Photochemical Oxidants
               on Nonbiological Materials 	      9-1

VOLUME IV
  Chapter 10.  Toxicological Effects of Ozone and Other
               Photochemical Oxidants 	     10-1

VOLUME V
  Chapter 11.  Controlled Human Studies of the Effects of Ozone
               and Other Photochemical Oxidants 	     11-1
  Chapter 12.  Field and Epidemiological Studies of the Effects
               of Ozone and Other Photochemical Oxidants 	     12-1
  Chapter 13.  Evaluation of Integrated Health Effects Data for
               Ozone and Other Photochemical Oxidants 	     13-1
                                       iv

-------
Chapter 10:  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.
Northrop Services, Inc.
Environmental Sciences
P.O. Box 12313
Research Triangle Park, NC  27709
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
0190LG/B
                         May 1984

-------
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. Srnolko
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
                                      VI
0190LG/B                                                              May  1984

-------
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
  and 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
University of California
Santa Barbara, CA  93106
Stress
Dr. George J. Jakab
Department of Environmental
  Health Sciences
Johns Hopkins School of Hygiene
  and Public 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
0190LG/B
                                     vn
                         May 1984

-------
                              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

Dr. William F. McDonnell
Health Effects Research Laboratory
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
Baltimore, ND  21205

Ms. Beverly  E. Til ton
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
0190LG/B
                                     vm
                         May 1984

-------
                                  CHAPTER 10
        TOXICOLOGICAL EFFECTS OF OZONE AND OTHER PHOTOCHEMICAL OXIDANTS

                               TABLE OF CONTENTS
LIST OF TABLES 	
LIST OF FIGURES 	
LIST OF ABBREVIATIONS 	

10.1  INTRODUCTION 	

10.2  REGIONAL DOSIMETRY IN THE RESPIRATORY TRACT 	
      10.2.1  Absorption in Experimental Animals 	
              10.2.1.1  Nasopharyngeal Absorption 	
              10.2.1.2  Lower Respiratory Tract Absorption 	
      10.2.2  Ozone Dosimetry Models 	
              10.2.2.1  Modeling Nasal Uptake 	
              10.2.2.2  Lower Respiratory Tract Dosimetry Models ...
      10.2.3  Predictions of Lower Respiratory Tract Ozone
              Dosimetry Modeling 	
              10.2.3.1  Illustration of Dosimetry Simulations 	
              10.2.3.2  Comparisons of Simulations to Experimental
                        Data 	
              10.2.3.3  A Use of Predicted Doses 	

10.3  EFFECTS OF OZONE ON THE RESPIRATORY TRACT 	
      10.3.1  Morphological Effects 	
              10.3.1.1  Sites Affected 	
              10.3.1.2  Sequence in which Sites are Affected
                        as a Function of Concentration and
                        Duration of Exposure 	
              10.3.1.3  Structural Elements Affected 	
              10.3.1.4  Considerations of Degree of Suscepti-
                        bility to Morphological Changes 	
      10.3.2  Pulmonary Function Effects 	
              10.3.2.1  Short-Term Exposure 	I	
              10.3.2.2  Long-Term Exposure 	
              10.3.2.3  Airway Reactivity 	
      10.3.3  Biochemically Detected Effects 	
              10.3.3.1  Introduction 	
              10.3.3.2  Antioxidant Metabolism 	
              10.3.3.3  Oxidative and Energy Metabolism 	
              10.3.3.4  Monooxygenases 	
              10.3.3.5  Lactate Dehydrogenase and Lysosomal Enzymes
              10.3.3.6  Protein Synthesis 	
              10.3.3.7  Lipid Metabolism and Content of the Lung ...
              10.3.3.8  Lung Permeability 	
              10.3.3.9  Proposed Molecular Mechanisms of Effects ...
      10.3.4  Effects on Host Defense Mechanisms 	
              10.3.4.1  Mucociliary Clearance 	
              10.3.4.2  Alveolar Macrophages 	
   XT
   xii
   xiii

   10-1

   10-3
   10-4
   10-4
   10-5
   10-6
   10-6
   10-6

   10-9
   10-10

   10-14
   10-14

   10-15
   10-15
   10-15
   10-38
   10-39

   10-42
   10-48
   10-48
   10-53
   10-57
   10-64
   10-64
   10-65
   10-80
   10-83
   10-87
   10-90
   10-95
   10-97
   10-100
   10-107
   10-107
   10-112
0190PT/A
5/1/84

-------
                        TABLE OF CONTENTS (continued).
10.5
10.6
              10.3.4.3  Interaction with Infectious Agents
              10.3.4.4  Immunology 	
      10.3.5  Tolerance 	
10.4  EXTRAPULMONARY EFFECTS OF OZONE
      10.
      10.
   4.1
   4.2
                                      Effects
      10.4.3
      10.
      10.
   4.4
   4.5
      10.4.6
Central Nervous System and Behavioral
Cardiovasculcir Effects 	
Hematological  and Serum Chemistry Effects ...
10.4.3.1  Animal Studies - In Vivo Exposures
10.4.3.2  In Vitro Studies 	
10.4.3.3  Changes in Serum 	
10.4.3.4  Interspecies Variations 	
Reproductive and Teratogenic Effects 	
Chromosomal and Mutational Effects 	
10.4.5.1  Chromosomal Effects of Ozone 	
10.4.5.2  Mutational Effects of Ozone 	
Other Extrapulmonary Effects 	
10.4.6.1  Liver 	
10.4.6.2  The Endocrine System 	
10.4.6.3  Other Effects 	
EFFECTS OF OTHER PHOTOCHEMICAL OXIDANTS
10.5.1  Peroxyacetyl Nitrate 	
10.5.2  Hydrogen Peroxide 	
10.5.3  Complex Pollutant Mixtures 	
SUMMARY
10.6.1
10.6.
10.6.
           2
           3
Introduction	
Regional Dosimetry in the Respiratory Tract 	
Effects of Drone on the Respiratory Tract 	
10.6.3.1  Morphological Effects 	
          Lung Function 	
          Biochemically Detected Effects of Ozone
          Host Defense Mechanisms 	
          Tolerance 	
              10
              10
              10
              10
             3.2
             3.3
             3.4
             3.5
      10.6.4
        Extrapulmonary Effects 	
        10.6.4.1  Central Nervous System and Behavioral
        10.6.4.2  Cardiovascular Effects 	
                                                              Effects.
              10.
              10.
              10.
              10.
           6.4.3
           6.4.4
           6.4.5
           6.4.6
          Hematological and Serum Chemistry Effects
          Reproductive and Teratogenic Effects 	
          Chromosomal and Mutational Effects 	
          Other Extrapulmonary Effects 	
      10.6.5  Effects of other Photochemical Oxidants
10.7  REFERENCES

      APPENDIX A
10-119
10-127
10-130

10-138
10-138
10-143
10-144
10-145
10-151
10-155
10-156
10-157
10-160
10-160
10-172
10-174
10-174
10-181
10-187

10-188
10-188
10-189
10-192

10-199
10-199
10-200
10-201
10-201
10-203
10-205
10-210
10-212
10-213
10-213
10-214
10-214
10-216
10-217
10-218
10-219

10-221

 A-l
0190PT/A
                                                                   5/1/84

-------
                                LIST OF TABLES
Table

10-1  Morphological effects of ozone 	
10-2  Effects of ozone on pulmonary function:  short-term exposures ..
10-3  Effects of ozone on pulmonary function:  long-term exposures ...
10-4  Effects of ozone on pulmonary function:  airway reactivity 	
10-5  Changes in the lung antioxidant metabolism and oxygen
      consumpti on by ozone 	
10-6  Monooxygenases 	
10-7  Lactate dehydrogenase and lysosomal enzymes 	
10-8  Effects of ozone on lung protein synthesis 	
10-9  Effects of ozone exposure on lipid metabolism and content of
      the 1ung 	
10-10 Effects of ozone on lung permeability  	
10-11 Effects of ozone on host defense mechanisms:   deposition and
      clearance 	
10-12 Effects of ozone on host defense mechanisms:   macrophage
      alterati ons 	
10-13 Effects of ozone on host defense mechanisms:   interactions with
      infectious agents 	
10-14 Effects of ozone on host defense mechanisms:   mixtures 	
10-15 Effects of ozone on host defense mechanisms:  immunology 	
10-16 Tolerance to ozone 	
10-17 Central nervous system and behavioral  effects of ozone 	
10-18 Hematology:  animal-~iji vivo exposure  	
10-19 Hematology:  animal--i_n vitro exposure 	
10-20 Hematology:  humna—j_n vitro exposure  	
10-21 Reproductive and teratogenic effects of ozone 	
10-22 Chromosomal effects from iji vitro exposure to high ozone
      concentrati ons	
10-23 Chromosomal effects from ozone concentrations at or below
      1960 ug/m3 (1 ppm)	
10-24 Mutational effects o'* ozone 	
10-25 Effects of ozone on the 1 iver	
10-26 Effects of ozone on xhe endocrine system, gastrointestinal
      tract, and uri ne 	
10-27 Effects of complex pollutant mixtures  	
   Page

   10-16
   10-49
   10-54
   10-58

   10-66
   10-84
   10-88
   10-91

   10-96
   10-98

   10-109

   10-113

   10-122
   10-125
   10-128
   10-132
   10-139
   10-146
   10-152
   10-153
   10-158

   10-161

   10-163
   10-167
   10-175

   10-182
   10-193
0190PT/A
5/1/84

-------
                                LIST OF FIGURES

Figure                                                                     Page

10-1  Predicted tissue dose for several trachea! 03 concentrations
      for rabbit and guinea pig 	     10-11
10-2  Tissue dose versus morphometric zone for rabbit and guinea
      pig and tissue dose versus airway generation for human at a
      trachea! 03 concentration of 510 pg/m3 (0.26 ppm) 	     10-12
10-3  Intracellular compounds active in antioxidant metabolism of
      the 1 ung 	     10-65
                                      xii
0190PT/A                                                                 5/1/84

-------
                             LIST OF ABBREVIATIONS
A-V
ACh
AChE
AM
AMP
ATPase
ATPS

BTPS

CC
Cdyn
CHEM
CL
C
CMP
CMS
CO
COHb
COLD
COMT
co2
CPK
CV
DL'
DLCO
DNA
E
ECG, EKG
EEC
ERV
FEF
   max
FEF
Atrioventricular
AcetyIcholine
Acety1choli nesterase
Alveolar macrophage
Adenosine monophosphate
Adenosine triphosphatase
ATPS condition (ambient temperature and pressure, saturated
with water vapor)
BTPS conditions (body temperature, barometric pressure,
and saturated with water vapor)
Closing capacity
Dynamic lung compliance
Gas phase chemiluminescence
Lung compliance
Static lung compliance
Cytidine monophosphate
Central nervous system
Carbon monoxide
Carboxyhemoglobi n
Chronic obstructive lung disease
Catechol-o-methyl-transferase
Carbon dioxide
Creatine phosphokinase
Closing volume
Diffusing capacity of the lungs
Carbon monoxide diffusion capacity of the lungs
Deoxyribonucleic acid
Elastance
Electrocardiogram
Electroencephalogram
Expiratory reserve volume
The maximal forced expiratory flow achieved
during an FVC.
Forced expiratory flow
0190PT/E
                                     xm
                                          May 9, 1983

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

                       Mean forced expiratory flow between 200 ml and 1200 ml  of
                       the FVC [formerly called the maximum expiratory flow  rate
                       (MEFR)].
FEF25-75%              Mean forced expiratory flow during the middle half of the
                       FVC [formerly called the maximum mid:expiratory flow  rate
                       (MMFR)].
FEF75%                 Instantaneous forced expiratory flow after 75% of the FVC
                       has been exhaled.
FEV                    Forced expiratory volume
FIVC                   Forced inspiratory vital capacity
fR                     Respiratory frequency
FRC                    Functional residual capacity
FVC                    Forced vital capacity
G                      Conductance
G-6-PD                 Glucose-6-phosphate dehydrogenase
Gaw                    Airway conductance
GMP                    Guanosine monophosphate
GS-CHEM                Gas-solid chemiluminescence
GSH                    Glutathione
GSSG                   Glutathione disulfide
Hb                     Hemoglobin
Hct                    Hemcitocrit
HO-                    Hydroxy radical
H20                    Watfcr
1C                     Inspiratory capacity
IRV                    Inspiratory reserve volume
IVC                    Inspiratory vital capacity
K                      Average mucous production rate per unit area
LDH                    Lactate deyhydrogenase
LDrn                   Lethal dose (50 percent)
LM                     Light microscopy
LPS                    Lipopolysaccharide
MAO                    Monamine oxidase
                                     xiv
0190PT/E                                                         May 9, 1983

-------
                       LIST OF ABBREVIATIONS (continued)
MAST
max Vr
max V02
MBC
MEFR
MEFV
MetHb
MMFR or MMEF
MNNG
MPO
MVV
NBKI
(NH4)2S04
N02
NPSH
AN2, dN2
°2
o2-
°3
P(A-a)02
PABA
PAco2
PaC02
PAN
PA°2
PaO.
PEF
PEFV
PG
PHa
PHA
PL
PMN
Kl-coulometric (Mast meter)
Maximum ventilation
Maximal aerobic capacity
Maximum breathing capacity
Maximum expiratory flow rate
Maximum expiratory flow-volume curve
Methemoglobin
Maximum mid-expiratory flow rate
N-methy1-N'-ni trosoguani di ne
Myeloperoxidase
Maximum voluntary ventilation
Neutral buffered potassium iodide
Ammonium sulfate
Nitrogen dioxide
Non-protein sulfhydryls
Nitrogen washout
Oxygen
Oxygen radical
Ozone
Alveolar-arterial oxygen pressure difference
para-aminobenzoic acid
Alveolar partial pressure of carbon dioxide
Arterial partial pressure of carbon dioxide
Peroxyacetyl nitrate
Alveolar partial pressure of oxygen
Arterial partial pressure of oxygen
Peak expiratory flow
Partial expiratory flow-volume curve
Prostaglandin
Arterial pH
Phytohemaggluti ni n
Trarispulmonary pressure
Polymorphonuclear leukocyte
0190PT/E
                                      xv
                                          May 9, 1983

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

PPD                    Purified protein derivative
Pgt                    Static transpulmonary  pressure
PUFA                   Polyunsaturated fatty  acid
R                      Resistance to flow
Raw                    Airway resistance
RBCs                   Red blood cells
Rcol-|                  Collateral resistance
R|_                     Total pulmonary resistance
RQ, R                  Respiratory quotient
Rt-                    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                    Thyroid-stimulating hormone
TV                     Tidal volume
UFA                    Unsaturated fatty acid
UMP                    Uridine monophosphate
UV                     Ultraviolet photometry
                                      xvi
0190PT/E                                                         May 9, 1983

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

V.                     Alveolar ventilation
V./Q                   Ventilation/perfusion ratio
VC                     Vital capacity
VCOp                   Carbon dioxide production
Vr,                     Physiological dead space
Vrv                     Dead-space ventilation
Vr,    .                 Anatomical dead space
Vr                     Minute ventilation
Vr                     Expired volume per minute
Vr                     Inspired volume per minute
 L
V,                      Lung volume
Vm3V                   Maximum expiratory flow
 lUaX
VCL                    Oxygen uptake
V00, Oo,               Oxyqen consumption
12s
   I                   Irradiated iodine
5-HT                   5-hydroxytryptamine
6-P-GD                 6-phosphogluconate dehydrogenase
                                      xv ii
0190PT/E                                                         May 9, 1983

-------
                           MEASUREMENT ABBREVIATIONS
g                      gram
hr/day                 hours per day
kg                     ki1ogram
kgm/mi n                ki1ogram-meter/mi n
L/min                  liters/min
pptn                    parts per million
mg/kg                  milligrams per kilogram
    o
mg/m                   milligrams per cubic meter
min                    minute
ml                     milliliter
mm                     millimeter
    3
pg/m                   micrograms per cubic meter
pm                     micrometers
uM                     micromole
sec                    second
                                    xvi n
0190PT/E                                                          May 9,  1983

-------
     10.   TOXICOLOGICAL  EFFECTS OF OZONE  AND  OTHER PHOTOCHEMICAL OXIDANTS

10.1  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 have been studied,  and many  of the effects
of ozone are  reasonably well described.  In  general, emphasis has been placed
on recent studies  at 1960 ug/m3 (1 ppm) of ozone or less.   Higher concentra-
tions have been cited  when the data add to  an understanding of mechanisms.
Concentrations of  1 ppm or greater  cannot be studied ethically  in man because
of the toxicity of even short-term exposures.
     Primary emphasis has been  placed on 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 these  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.
     Eirphasis  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.   Technical considerations  for  inclusion  of a specific
study included an  evaluation of the exposure  methods; the analytical method
used to  determine the  chamber  ozone concentration;  the  calibration of the

0190PT/B                           10-1                            5/1/84

-------
ozone monitoring equipment and the analytical  methods used (wherever possible);
the  species,  strain,  age  and  physical  characteristics of the  the  subject
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 judgments
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  DO
are next 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 protective role 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 beyond those now
well documented in the respiratory system.
     A brief discussion of the available literature on the effects of other oxi-
dants 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  in  this area,
but  its  brevity does  not  necessarily reflect  a  general  lack of  importance.
0190PT/B                           10-2                             5/1/84

-------
     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.
10.2  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 the goal  is dosimetry, which is the specification of the
quantity of inhaled material,  in this instance ozone (00, 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 03 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  Oo  ab-
sorption in the  species and in different-sized, 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
research,  and to  test  our understanding of  03  absorption in the respiratory
tract.

0190PT/B                           10-3                            5/1/84

-------
10.2.1  Absorption in Experimental  Animals
     There have been very few experiments in which measurements of the regional
uptake of (L  or  other reactive gases have  been  determined.   Of the several
results published, only  one  is concerned with the uptake of 0, in the lower
respiratory tract; the others deal  with nasopharyngeal uptake.
10.2.1.1  Nasopharyngeal Absorption.   Nasopharyngeal  removal  of  03 lessens
the quantity  of  0.,  delivered  to the lung  and must  be  accounted for when
estimating the 0,  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 0^  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.  Their data, however, indicate that the trachea! concentra-
tion  increases with  increased  nose or mouth concentrations.  They also demon-
strated that  the concentration  of 0,  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  Qg in the nasopharynx of  acutely
and chronically  exposed  dogs.   Beagles chronically exposed (18  months) to 1 to
3 ppm  of 03 under  various daily exposure regimes had  significantly higher tra-
cheal  concentrations of  0, than animals tested after 1 day of  exposure to cor-
responding regimes.  Moorman et al.  (1973)  suggested that the  differences were
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,

0190PT/B                           10-4                             5/1/84

-------
they had significantly lower trachea! values than those continuously exposed.
The average tracheal  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 0^ 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 tracheal 0- concentrations significantly greater than
those observed with acute exposure.
     Nasopharyngeal  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 tracheal
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 tracheal  and chamber concentrations  is  in  agreement with Yokoyama  and
Frank (1972).
10.2.1.2  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 03 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 10.3).
     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
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
03 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.
0190PT/B                           10-5                            5/1/84

-------
10-2.2  Ozone Dosimetry Models
10.2.2.1  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  (MAS, 1977) concluded that the  model predicts 99 percent penetration
for 0~.   This  is  much  more  than that measured by Yokoyama and Frank (1972) or
by Miller  et al.  (1979).   NAS (1977) discusses several  possible  reasons  for
the differences,  but considers the major factor 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
assumption of  quasi-steady-state  flow,  mass balance,  and that  the flux of a
trace  gas  at the  air-mucus  interface  is  proportional  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  0^
uptake  data of Yokoyama and Frank (1972) and concluded that the uptake coef-
ficient  (average  mass  transfer coefficient) for (k, as well as for the other
gases  considered, increases with increasing air flow rate.
10.2.2.2   Lower Respiratory Tract Dosimetry Models.   There are  two models for
which  published results are available.   The model  of McJilton  et al.  (1972)
has never been formally published; however, its features have  been discussed
in detail  (National  Academy of Sciences, 1977; Morgan and Frank, 1977)  and
simulation results  for 0,  absorption in  each generation of the  human lower
respiratory tract are  available (National Academy  of  Sciences, 1977; Morgan
and Frank, 1977).  A detailed description of the formulation of the mathemati-
cal model  of  Miller and co-workers is found in Miller (1977) and  an  overview
is given  in  Miller (1979).    Results  of  simulations of the respiratory tract
absorption of  03  in humans,  rabbits,  and guinea  pigs are in Miller  (1977,
1979),  and Miller et al. (1978).
 0190PT/B                           10-6                            5/1/84

-------
     Because  both  models  were developed to  simulate  the local  absorption of
03, they have much in common.  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-, and others
in  the  lower  respiratory tract (see Scherer et al., 1972; Paiva, 1973; Chang
and Farhi, 1974; Yu, 1975; Pack et al., 1977; Bowes et al., 1982).  The approx-
imation is appropriate for 0., as well.
     McJilton et al.  considered  only molecular diffusion for CL transport in
the  liquid  lining  the airway (National Academy of Sciences, 1977; Morgan and
Frank,  1977).   In  addition  to this  process,  Miller et al. (1978)  included the
effect  of  chemical  reactions on CL transport in the liquid lining.  By using
                                  «J
•their respective  assumptions,  equations based on mass balance considerations
were developed  for (L transport in the  lining.   The liquid-phase equations
were coupled  to the  gas-phase  equations based on physical  considerations:
continuity of 0, transfer across the gas-liquid interface was required and the
gas  and liquid  phase interfacial concentrations were  related  by  Henry's  law
constant.
     Airway or  morphometric  zone models such as  those  of Weibel  (1963) and
Kliment  (1973)  were  used in  both models  to  define  the  lengths,  radii,  surface
areas,  cross-sectional  areas,  and  volumes of the  airways and  air spaces of
each generation or zone.   A sinusoidal  breathing  pattern was  used;  however,
dimensions were held constant throughout the breathing  cycle.  The  physical
properties of the liquid lining were assumed to be those  of water.  The lining
thickness depended  on generation or zone, being  thicker in  those  with lower
numbers than  in those with higher numbers.
     The flux of 03 from the lumen to the liquid lining  was defined in terms
of  a mass-transfer coefficient which depended on  gas-phase  and  liquid-phase
transfer coefficients.   McJilton et al.  (National Academy of Sciences, 1977)
made the assumption  that the mass transfer  was  liquid-phase controlled and
estimated the transfer coefficient from empirical data on the physical proper-
ties (not  chemical) of  the  liquid  lining and of 03-   In the  Miller model
(Miller, 1977)  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.  It  was combined  with the liquid-phase

0190PT/B                           10-7                            5/1/84

-------
mass transfer coefficient (which depended on the chemical  and physical  proper-
ties of the  liquid and of 0~) to obtain the overall transfer coefficient.  No
assumption was necessary or made about which phase, if either, controlled the
transfer.
     Ozone concentration  in  tissue and  at  the liquid-tissue interface was
assumed to be  zero  (Miller,  1977;  National  Academy of Sciences, 1977).  The
National Academy of  Sciences  (1977)  and Miller (1977) believe this boundary
condition means that  CU  reacts  (chemically) instantaneously when  it reaches
the tissue.  Miller  et al.  (1978)  define the tissue dose as that quantity of
0,  per  unit  area reacting with the tissue  at the liquid-tissue interface.
     There are major  differences  in  the two models.   Ozone is known to react
chemically with constituents of the liquid lining.  This factor was not included
in  the  McJilton  et  al. model.   Chemical reactions were taken into account in
the mucous-serous  lining In  the  Miller model.   The  inclusion  resulted in
significant differences between the tissue dose pattern curves in the tracheo-
bronchial  region  predicted  by  the two models.   The  McJilton  et al.  model
predicts a dose curve (equivalent  to a  tissue  dose curve because of no  mucous
reactions) in  the conductive  zone  that has its maximum  at the trachea and
decreases distally  to the thirteenth or  fourteenth generation (see Figure 7-5
in  National Academy of Sciences, 1977).   By contrast,  the Miller model  predicts
the tissue dose to be a minimum at the trachea and to increase distally to the
pulmonary region.
     Miller  and  co-workers  took  into  account the reaction  of  0.,  with the
unsaturated  fatty acids  (UFA)  and ami no acids  in the mucous-serous lining.
Reactions of 0~ with other components (such as carbohydrates) were not  included
in  the model because of insufficient information  (Miller, 1977; Miller  et al.,
1978).  The  CL-UFA  and  CL-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 specification of the production
rate  of  the  UFA  and amino 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  03, and  the assumption  that  the  production rate decreased distally
(Miller, 1977).
     Although  the instantaneous reaction scheme is a  good preliminary approach
to  treating  CL reactions in the mucous-serous  lining, its use is not completely

0190PT/B                           10-8                            5/1/84

-------
justifiable.   Second-order rate  constants  of 0- with some of the mucous UFA
indicate that although they are large (Razumovskii and Zaikov,  1972), they are
less than the diffusion-limited rates necessary for the instantaneous reactions
and  scheme.   Experimental  evidence  (Mudd  et al., 1969)  suggests  that the
reactions of  0-  with  amino acids are very  rapid.   Rate constants for these
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 Miller model  also 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 et al.  did  not take  this factor  into  account
(Morgan and Frank, 1977).  However, this may not be an important difference in
the  two  models.   Pack et al.  (1977) and Engle  and Mack!em (1977)  reported
results that  indicate an insensitivity of airway concentrations to the effective
dispersion coefficient.

10.2.3  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  a possible  use for dosimetry modeling.
     The following  discussion  of modeling  results of lower respiratory tract
absorption is based on  simulations using the Miller  model.  This is  because
the  model  includes  the  important effects of  0~  reactions  in the mucous-serous
lining and  because simulations of  0,  absorption  in  laboratory animals are
available.
     Simulations of CL  absorption  in different animals can be carried out by
modifying input  parameters  of the computer  program that  solves the mathema-
tical equations.  These  input parameters, which characterize an animal, include
the  number and  dimensions  of  the airways, tidal volume,  and length  of  time of
one  breath.  The airway  and alveolar dimensions of Weibel (1963) were used for
the  simulation  of  0_  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).  In general, more than one generation  in man corresponds to one
zone  in  an animal.   However,  each zone is  characterized by  the  number of
airways and their  lengths  and diameter.  The zone models were used because

0190PT/B                           10-9                            5/1/84

-------
they were the only complete (tracheobronchial and pulmonary regions) "airway"
models available at  the  time.   Also, similar zone models,  such as that of
Landahl (1950),  have been used successfully in human  aerosol  deposition studies.
     To illustrate  simulation  results,  three aspects of  the  simulations by
Miller and  co-workers are  considered:   1) the effect  of various tracheal
concentrations on the  tissue  dose pattern (tissue dose as a function of zone
or generation) in guinea pigs and rabbits; 2) the similarity between the dose
patterns of guinea pigs,  rabbits, and humans; and 3)  the effect on simulations
due to the uncertainty of the knowledge of the production rates of the mucous-
serous components that react with (L and of the knowledge of the mucous-serous
thickness.
10.2.3.1  Illustration of Dosimetry  Simulations.   Figure 10-1  is  a  set of
plots  of  the  tissue  dose for one breath  versus zone  for  various  tracheal  0.,
concentrations for rabbit and guinea pig simulations.   (Ambient concentrations
would  be  roughly twice the  tracheal  concentration according to  Miller  et al.,
1979.)  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 maximum dose of 03.  Although the model predicts  uptake of 03 by respiratory
tissue (zones 6,  7,  and  8) for  all  tracheal  concentrations studied  (62.5  to
         3
4000 ug/m ), the penetration of 0., to the tissue in the airways lined by mucus
depended  on the  tracheal  concentration and the specific animal species.  For
example,  as illustrated  in  Figure 10-1 for  the tracheal  0- concentration  of
         O                                                 -5
1000 ug/m , no 03 reaches the airway tissue of the rabbit until zone 3, whereas
00  is  predicted  to  penetrate to  the guinea  pig airway tissue  in all  zones.
                                                                             3
However, at the two  lowest tracheal concentrations plotted, 250 and 62.5 jjg/m  ,
no  penetration occurs until zones 4 and  6,  respectively, for  both animals.
     The  similarity  of the  predicted dose patterns in rabbits  and guinea pigs
extends to  the  simulation of 0,  uptake  in humans.   Figure 10-2 compares the
                                                                            3
tissue dose for  the three  species  for  a  tracheal  concentration of 500
 (0.26 ppm).  Because of the different airway models used, the plot of the data
 for  the  laboratory animals differs from that of Figure 10-1.  In Figure 10-2
 the  guinea  pig and rabbit tissue dose  are  plotted  in  the  form of a  histogram.
 This  a"i lows a plot of  the human data,  even though there  is more than one
 generation  corresponding  to one  zone.
 0190PT/B                           10-10                                 5/1/84

-------
   1Q5
£
i
O  10'6
u.
O
IU

8
Q
UJ
D
(A
£  10-7
   108

              O--D-
                                   TRACHEAL
                                   O, CONC.

                                  ^g/m3  ppm

                               • Q40062.041
                               AA1000   0.510
                               • O 250   0.128
                               •O  62.5  0.032
I    i    I   I   I    i   I    i    I
              0

         TRACHEA
       2   3
      BRONCHI i
 5  6
JBRON-
                               1CHIOLES
 7   8 MODEL ZONE
AD.JA.S MORPHOMETRIC ZONE
        Figure 10-1. Predicted tissue dose for several tracheal
        O3 concentrations for rabbit (open symbols) and
        guinea pig (closed symbols). See text for details.
        (A.D. = alveolar  duct; A.S. = alveolar sac).
        Source: Adapted from Miller et al. (1978).
                              10-11

-------
IU -




1
£ 106
.Q
N
E
u
1
n
o
u.
O
LU
O
O
LU
3
| 107









108
;' 1 • 1 i 1 •
-

•*•


— Q
^^
-»*-,


—
_






—
~
"™
•"•
«.



MM


1 1 I 1 1 1 i
	 '

•
« | 1 | 1 | 1-
fr \m 0 	 0 .
/ \ . 	 .
/ \ / \ ~ - -\
T V
0 ' \
V7"1 . ^
i \
1 » -^1
Ol 1 \
• ' 1 ', Z
IT - -"

1

/
/
1
/
/
/
1










1 1 l 1 1 1 .
\
\
\ _
\
\
\
•
\ ""
\
\
\
\
\
\
\
\
\
\ -
\
\
i 1 . 1 . 1 k





























AIRWAY
0 2 4 6 B 10 12 14 18 18 20 22 GENERATION
™A: BRONCHI BRONCHIOLES A.D. 1 A.S. MORPHOMETRIC
uncA I 1 ZONE
Figure 10-2. Tissue dose versus morphometric zone for rabbit (O)
and guinea pig (•); and tissue dose versus airway generation for
human (-•-) at a tracheal O3 concentration of 500 /jg/m3 (0.26
ppm). See text for details. (A.D. = alveolar duct; A.S.  =  alveolar
sac)
Source: Adapted from Miller et al. (1978).
                             10-12

-------
     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, 0., penetrates to the
tissue everywhere in  the  pulmonary region (zones > 5 and generations > 16);
however, 0,  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.
     The characteristics of the ozone tissue dose pattern for the three species,
as  discussed  here,  are general to  the  simulations  presented by Miller and
co-workers (Miller, 1977; Miller et al., 1978; Miller, 1979):  1) the maximum
tissue dose  occurs  in the first surfactant-lined zone or  generation;  2) 0-
penetrates to tissue  everywhere in  the  pulmonary region  (zones > 5), decreas-
ing distally  from the maximum; and 3) the onset of 0- penetration to mucous-
lined tissue, as well  as  dose in general, depends on trachea! Q~ concentra-
tion, animal  species, and the breathing pattern.   These general  characteristics
of tissue dose pattern are independent of the two airway models  used.   Whether
or  not  this  would be  so  if more  recent and detailed  airway models  (e.g.,
Phalen et al., 1978) were used or if (L uptake were simulated in other species
would be of interest in understanding CL absorption.
     Sensitivity studies  were carried  out with two  major  parameters,  the
average mucous production rate per unit area (K ) and the mucous-serous thick-
ness (Miller, 1977;  Miller  et al.,  1978; Miller, 1979).   The parameter KQ is
related to the  chemical  reaction  scheme used by Miller and co-workers; as K
increases, so does the  reactivity.  The  study showed that if K  overestimated
the correct  value,  there  would be no difference in the predicted respiratory
bronchiolar (zone 6 or generation 17) dose for any of the three  species studied.
If  KQ underestimated  the  correct  value, the results were not as clear; they
depended on the tracheal concentration and species.   The guinea  pig and rabbit
were the least sensitive species;  they required a twenty-five and a fifty-fold
                                         3
underestimate of KQ  for 325 and 750 ug/m  tracheal 0, concentration, respec-
tively,  before significant  variations in respiratory bronchiolar tissue dose
occurred.   Significant differences occurred in the human dose for a ten- and a
twenty-five-fold underestimate (corresponding to 325 and 750 ug/m ,  respective-
ly).  Decreasing KQ  also  allowed  more 03  to penetrate  to the tissue in the
conducting airways, and  increasing  RQ resulted in less 03 penetrating to the
conducting airway tissue.   In  the  other sensitivity study,  a 25 percent  in-
crease in the mucous lining thickness decreased the conducting airway dose but
had no  effect on respiratory  region  tissue  dose (see Figures 2 and  8 and
discussion in Miller, 1977).
0190PT/B                           10-13                                5/1/84

-------
10.2.3.2  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 0» in the  lower  respiratory  tract.   Yokoyama and Frank
(1972) observed 80  to  87  percent uptake of 0- by the lower respiratory tract
of dogs.   Miller et al. (1978) predict a 60 percent  uptake for humans; however,
because of differences between  two species, comparing the two 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 10.3.1.1).   Of interest
is that for the animals simulated by Miller and co-workers,  the maximum tissue
dose  of  0^ occurs  in  this  junction and decreases distally.   Thus,  in the
pulmonary region,  the  model results are  in qualitative agreement with  experi-
mental observations.
     Damage is also observed in the trachea and bronchi  of animals (see Sections
10.3.1.1.1.2  and  10.3.1.1.1.3).  In  the  animals  modeled, the Miller  model
either predicts significantly  less  tissue dose in these  regions 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 conducting airways (see Figure 10-2).
A  factor complicating  an  understanding of the above is the possibility of 03
reaction products  being toxic.   Other factors confounding interpretation are
cell sensitivity and animal differences.  Also, much of the reported damage in
the trachea and bronchi is  associated with the cilia  of ciliated cells, which
in the Miller model are not part  of  the tissue.   The cilia extend into the
hypophase  (perciliary) portion  of  the mucous-serous  layer, and the  Miller
dosimetry model does not  distinguish the cilia of  the  ciliated cells as  a
separate component of this layer.  Thus,  zero or relatively low  predicted
tissue dose should  not be interpreted as predicting no cilia damage.   Likewise,
the frequent  reporting of cilia being damaged following 0_ exposure should not
be interpreted  necessarily as an indication of 0., tissue dose since the defini-
tion of tissue  applied does not  include cilia.
10.2.3.3  A Use of  Predicted Doses.  Miller (1979) discusses how doses predicted
by the model  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.   The procedure  is  to  simulate tissue

0190PT/B                           10-14                                5/1/84

-------
dose  for  several  species for  the  same time (Miller, 1979)  for  a  range of
trachea! CL  concentrations.  The doses  for  a specific zone or generation,  for
each  species,  are  plotted versus  tracheal  or ambient 0-  concentration.   By
using  such  plots  and information on  nasopharyngeal  removal  (Miller et al.,
1979), the ambient concentration necessary to produce the same dose in differ-
ent species  can be estimated (see Miller, 1979).  Also, the  relative  quantity
of OT  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.
10.3  EFFECTS OF OZONE ON THE RESPIRATORY TRACT
10.3.1  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 10-1),  it appears  likely  that many of the postexposure
biological processes of animals could also occur in  man.
10.3.1.1   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 submucosa 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

0190PT/B                           10-15                                5/1/84

-------
                                                    TABLE  10-1.  MORPHOLOGICAL  EFFECTS OF OZONE
Exposure
Ozone . duration
concentration Measurement ' and
ug/m3 ppm method protocol Observed effect(s)
Species Reference
    196


    392
o
 i
0.1


0.2
UV,
NBKI
7 days,    Two of six fed with "basal" vitamin E diet had increased cen-
continuous 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.
Rat     Plopper et al., 1979
196
392
392
686
980
1568
392
686
392
980
1568
0.1
0.2
0.2
0.35
0.5
0.8
0.2
0.35
0.2
0.5
0.8
UV,
NBKI
MAST,
NBKI
UV,
NBKI
UV,
NBKI
uv,
NBKI
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
Five of six fed E-deficient "basal" diet had centriacinar AMs
and bronchiclar 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
surface area. Lung weight and alveolar number did not
change. Decrease in lung tissue elasticity. Parenchyma
appeared "normal" by LM.
Respiratory bronchiolitis at all concentrations. Increased
AMs. Bronchi olar epithelium both hyperplastic and hypertrophic.
Increased alveolar type 2 cells. Random foci of short, blunt
cilia or absence of cilia (LM, SEM, TEM).
All exposed monkeys had LM & EM lesions. Trachea and bronchi
had areas of shortened or less dense cilia. RBs had AM
accumulation and cuboidal cell hyperplasia. Alveoli off RBs
had AM accumulations and increased type 2 cells. RB walls of
the 0.35-ppm group were often thickened due to mild edema and
cellular infiltration.
Exposed groups gained less weight. Focal areas of missing or
damaged cilia in trachea and bronchi. TB nonciliated (Clara)
cells were shorter and 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.
Rat
Rat
Monkey
(Rhesus
and
Bonnet)
Monkey
(Bonnet)
Rat
Chow et al . , 1981
Bartlett et al . , 1974
Dungworth et al., 1975b
Castleman et al. , 1977
Schwartz et al . , 1976

-------
                                            TABLE 10-1.  MORPHOLOGICAL EFFECTS OF OZONE  (continued)
Ozone
concentration Measurement '
ug/m3
392
980
1568








392
980
1960



ppm method
0.2 UV,
0.5 NBKI
0.8


1
i




0.2 NO,
0.5 NBKI
1.0



Exposure
., duration
b j
and
protocol
20, 50, or
90 days;
8 hr/day








4 days,
3 hr/day,
exercised in
a rotating
case alter-
nate 15 min


Observed effect(s)c
Epithelial changes and PAM accumulations at 90 days were
similar to 7-day exposures, but less severe. 0.5- and 0.8-ppm
groups had increased centriacinar PAMs at all times. 0.2 ppm
and controls could not be separated by "blind" LM examination,
nor were there distinguishing EM 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. Nonci Hated cells were flattened
lumenal surfaces that occasionally occurred in clusters. Proximal
alveoli of 20- and 90-day 0.8-ppm groups had thicker blood/air
barriers.
Exercised control mice have significantly smaller body weights.
Both unexercised and exercised mice exposed to 0.5 or 1.0 ppm
had smaller body weights and larger lung weight. Exercised
mice exposed to 0.2 ppm also had larger lung weights. Other
pathology not studied.



Species Reference
Rat Boorman et al., 1980










Mouse Fukase et al . , 1978 i
(male,
5 weeks
old, l
ICR-JCL)!

 392    0.2      CHEM
1960    1.0
9800    5.0
7, 14, 30, 60, Short-term exposures produced a slight degree of tonsil  epithelial
90 days; con-  detachment.   Cell infiltration below the epithelium was  slight.
tinuous        Long-term exposures caused slight edema of the lacunar epithelium
               which was destroyed or detached in places.  Lymphocyte infiltra-
               tion also occurred.

10 days, con-  Tonsil epithelium had a high degree of detachment.   Cell satura-
tinuous        tion occurred below the epithelium.   Some protrusion of  the tonsil
               into the oral cavity.

3 hr           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.
                                                                                                                    Rabbit   Ikematsu, 1978

-------
                                                  TABLE 10-1.  MORPHOLOGICAL EFFECTS OF OZONE   (continued)
Exposure
Ozone b duration
concentration Measurement ' and
ug/mj
490
510
980
1960
588
ppm method
0.25 NO
0.26 MAST,
0.50 NBKI
1.0
0.3 NBKI
protocol
6 weeks ,
12 hr/day
4.7-6.6 hr,
endotracheal
tube
16 days,
3 hr/day
Observed effect(s)c Species Reference
Centriacinar alveoli had more type 1 and 2 alveolar epithelial Rat
cells and macrophages. Type 1 cells were smaller in volume,
covered less surface, and thicker. Type 1 cell damage and
occasional sloughing seen by TEM.
Desquamation of ciliated epithelium. Focal swelling or Cat
sloughing of type 1 cells.
SEM, but not LM, showed swollen cilia with hemispheric Rat
extrusions and surface roughness. Some adhesion of
Barry et al. , 1983
j Boatman et al . , 1974.
i „ — 	 	
Sato et al. , 1976a
o
 I
I—>
CO
severely injured cilia occurred.   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
and attachment of round bodies.  All responses were pronounced
in vitamin E-deficient rats.   Some rats had chronic respiratory
disease.
588



588


686
980
1372
1470
1960
686
980



0.3 NBKI



0.3 UV


0.35 ND
0.50
0.70
0.75
1.00
0.35
0.50



28 weeks,
5 days/week,
3 hr/day

6 weeks,
5 days/week,
7 hr/day
1, 2, 4, 5, 6
or 8 days, con-
tinuous


4 days, contin-
uous, followed
by 0.50, 0.70,
0.75 or 1.00
for 1-4 days
No morphological differences noted between vitamin E-defi- Rat ; Sato et al., 1980 '
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 Sherwin et al., 1983'


Dividing cells were labeled with tritiated thymidine and Rat Evans et al., 1976b
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 03 (0.35 ppm) did not maintain
tolerance. Groups exposed to higher initial concentration
(0.50 ppm) demonstrated tolerance.

-------
TABLE 10-1.   MORPHOLOGICAL EFFECTS OF  OZONE   (continued).
Ozone
concentration Measurement '
pg/nr1 ppm method
784 0.4 MAST
784 0.4 NBKI
980 0.5 NO *
D
J 1
980 0.5 UV,
or or NBKI
1568 0.8
980 0.5 UV,
NBKI
980 0.5 UV,
1568 0.8 NBKI
Exposure
, 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 24 hr/
day
7, 21, and
35 days,
continuous
7, 28, or
90 days,
continuous,
8 hr/day
Observed effect(s)c
All (exposed and control) lungs showed some degree of inflam-
matory infiltrate possibly due to intercurrent 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 framework of the alveoli was
difficult. Small pulmonary arteries had thickened tunica medias,
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.
Principal lesion was a "low-grade respiratory bronchiolitis"
characterized by "intraluminal accumulations of macrophages
and hypertrophy and hyperplasia of cuboidal bronchiolar
epithelial cells." Conducting airway lesions not apparent
by LM, but parallel linear arrays of uniform shortening and
reduction of density of cilia by SEM. Kulschitzky-type cells
appeared more numerous in exposed.
Species
Rabbit
(New
Zealand)
Rabbit
Rat
(young
males)
Rat
Mouse
(Swiss-
Webster;
60 days
old; 35-40 g)
Monkey
(Bonnet)
Reference
P'an et al. , 1972
Boatman et al . , 1983
Stephens et al. , 1974b
Brummer et al. , 1977
Zitnik et al., 1978
Eustis et al. , 1981

-------
                                               ABLE  10-1.   MORPHOLOGICAL  EFFECTS  OF  OZONE   (continued)
ro
o
Ozone
concentration Measurement '
ug/fnj ppm method
980 0.5 UV,
1568 0.8 NBKI
980 0.5 UV,
to to NBKI
3920 2.0
>
Exposure
u duration
and
protocol Observed effect(s) Species Reference
7 days, All exposed monkeys had lesions. Lesions similar in 0.5- Monkey Mellick et al., 1975,
8 hr/day 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 eosinophilis and debris.
Nonciliated cuboidal bronchiolar cells were larger, more
numerous, and sometimes stratified. Proximal alveolar
epithelium thickened by increased numbers of type 2 cells.
Progressive decrease in intensity of lesions from proximal to
distal orders of RBs.
7 days, Elevated collagen synthesis rates and histologically Rat Last et al . , 1979
24 hr/day discernible fibrosis was present at all levels of 03.
0.5 ppm Minimal or no thickening of walls or evidence of
fibrosis. Increased number of cuboidal cells and
macrophages present.
          0.5
          to
          1.5
                        14 days
                        and
                        21 days,
                        24 hr/day
                            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 0,
CHEM,
NBKI
                                  6 months,      Only  03  caused  pulmonary  lesions.   Only  LM histopathology,
                                  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
                                                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 10-1.  MORPHOLOGICAL EFFECTS OF OZONE  (continued)
       Ozone
   concentration
   ug/m3   ppm
       Measurement
          method
                  a.b
              Exposure
              duration
                and
              protocol
                              Observed effect(s)
Species
Reference
    980
0.5
           0.5 0,
uv,
NBKI
o
 i
ro
3, 50, 90,    H2S04 did not potentiate effects of 03 alone.   Fixed lung
or 180 days;  volumes were increased at 180 days, but decreased at 62 days
continuous,   postexposure.  After 50, 90, 94 180 days all 03 exposure
plus 62 days  rats had "bronchiolization of alveoli" or formation of an RB
postexposure, between the TB and AOs.   Centriacinar inflammatory cells were
24 hr/day     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 mild
              interstitial thickening of alveolar septa in the centriacinar
              region (LM, SEM, TEM).
                                                                                                                      Rat    Moore  and Schwartz, 1981
   1058
   1725
0.54
0.88
ND
   1058    0.54
2, 4, 8, 12,  Severe loss of cilia from TB after 2 hr.   TB surface more
or 48 hr      uniform in height than controls.   Necrotic ciliate cells in
              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
              minimal 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.   Macrophages
              accumulated in proximal alveoli.   Endothelium appeared
              relatively normal.

              Repair started at 20 hr.   Type 2 cells divide, cuboidal
              epithelium lines proximal alveoli whi :'e 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, SEM,  TEM).

6 months,     No mention in either the results or discussion of  the 6 months
24 hr/day     at 0.54-ppm group.
Rat    ^tephens et al. ,  1974a

-------
                                               TABLE  10-1.  MORPHOLOGICAL  EFFECTS  OF  OZONE   (continued)
o
 i
Ozone
concentration Measurement '
ug/m3
1058
1725



1764



490


1176
2548



1372

1568


1372

ppm method
0.54 ND
0.88



0.9 03
•«-
0.9 N02

0.25 03
+
2.5 N02
0.6 1 NO
1.3



0.7 UV,
to NBKI
0.8


0.7 ND

Exposure
. duration
b and
protocol
4 hr
to
3 weeks


60 days



6 months


1 or 2 days,
6 hr/day or
7 hr/day


7 days ,
continuous



24 hr,
continuous

r~
Observed effect(s)
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 N02 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.
In situ cytochemical studies of lungs from 03 exposed and
control rats. Ozone-exposed rats had increased acid phosphatase,
both in lysosomes and in the cytoplasm, in nonciliated bronchiolar
(Clara) cells, alveolar macrophages, type 1 and 2 cells, and
fibroblasts.
Exposure end: General depletion of cilia from TB surface.
Nonciliated cells were shorter and contained


Species
Rat
(month
old)


Rat
(month
old)

Rat
(month
old)
Mouse
(young)



Rat




Rat



Reference
Freeman et al. , 1974
!










Bils, 1970




Castleman et al. , 1973a




Evans et al . , 1976a

                                                 Post exposure:
                                                 (0-4 days)
fewer dense granules, less SER, and more free
ribosomes.

TB returned towards normal.

-------
TABLE 10-1.   MORPHOLOGICAL EFFECTS OF OZONE  (continued)

Ozone
concentration
(jg/ra3 ppm
1568 0.8









0
i
ro
CO


1568 0.8





1568 0.8







Exposure
. duration
Measurement ' and
method protocol
UV, 6, 10, 20
NBKI days
exposure,
24 hr/day
or
20 days
exposure +
10 days
postexposure







UV, 7 days,
NBKI continuous
with samples
at 6, 24,
72 and
168 hr.
UV, 4, 8, 12,
NBKI 18, 26, 36,
50, and post
48 and
168 hr,
continuous





Observed effect(s)c Species Reference
SEM 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 Lum et al., 1978
changed after 03 exposure; fewer ciliated and more non-
ciliated secretory cells.



Degeneration and necrosis of RB type 1 cells predominates Monkey Castleman 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
macrophages increased during exposure, but marked clusters
of K cells at 26-36 hr.

-------
                                                'ABLE 10-1.  MORPHOLOGICAL EFFECTS OF OZONE  (continued)
Ozone
concentration
pg/m3 ppm
1568 0.8










Exposure
. duration
Measurement ' and
method protocol
UV, 3 days, 1st Exposure:
NBKI continuous
0, 2, 6, 9,
16, and 30
days post-
exposure;
2nd 3 days
continuous
after 6, 13,
and 27 days Postexposure:
postexposure 6 days:


Observed effect(s)c
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.




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





Most obvious lesions were not present. TB epithelium
o
i
ro
          had usual pattern.  Clumps of macrophages had cleared
          from the lumen.  Most proximal alveoli lined by normal
          type 1 and 2 cells.

30 days:  Lungs indistinguishable from controls.
2nd 3-Oay 6 or 27 days after the end of the 1st. Lesions
Exposure: same as 1st exposure.
1666 ,0.85 NO
Similar
exposure
regimen
for 14 ppm
N02, but not
mixtures.
1960 1.0 NBKI
1, 2, 3
days,
continuous
~60 weeks,
~5 days/week,
6 hr/day
(268 expo-
posures)
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- j
Dawley; ',
1, 5, 10,
15, 20, 25,
30, 35, and
40 days old)
Mouse, Stokinger et al., 1957
Hamster,
Rat

-------
                                                      TABLE  10-1.  MORPHOLOGICAL EFFECTS OF OZONE  (continued)
Ozone
concentration
(jg/m3
1960
to
1960
1960
1960
5880
ppm
1.0
to
3. 0
1 0
1.0
1.0
2.0
3.0

Measurement3'
method
NO

A
B
C
D
E
Expasure
duration
and
protocol
18 months,
8-24 hr/
day
= 8 hr/day
= 16 hr/day
= 24 hr/day
= 8 hr/day
= 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.
                                                            C    1 ppm, 24 hr/day:  More extensive fibrosis of
                                                                 centriacinus.  Thickened AO walls.  More "extra"
o
1
ro
en


RB and

AO.

D & More fibrosis. Epithelial hyperplasia and squamous
E metaplasia.
1960
3920









1-0 CHEM,
and NBKI
o n


mixtures
with
H2S04




2 or 7 days, Results:
6 hr/day 03 alone:
1 ppm:





2 ppm:
03 plus
H2S04
Rat Cavender et al. , 1977
Lesions limited to centriacinus. and
Hypertrophy and hyperplasia of TB epithelium. Guinea
Centriacinar alveoli had increased type 2 pig
cells, increased macrophages, and thickened
walls. Some edema in all animals.
Lesions less severe at 7 days than at ? days.
This adaptation was more rapid in rats than
guinea pigs.
Same plus loss of cilia in bronchi.
No additive or synergistic morphological
changes.
Measurement method   MAST - Kl-coulometric (Mast meter), CHEM = gas phase chemiluminescence),  NBKI = neutral  buffered potassium iodide,
UV = UV photometry; ND = not described.

Calibration method   NBKI = neutral  buffered  potassium  iodide.

Abbreviations used:  LM = light microscope, EM = electron microscope; SEM = 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 endoplasmic reticulum,  RER =  rough endoplasmic reticulum.

-------
and  glands  associated with  the  sense  of smell.  Significant morphological
differences exist  among  the  various animal species  used  for 0- exposures  as
                                                              »J
well as between most of them and man (Schrieder 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 "scrubbing"  has been studied (Yokoyama and Frank, 1972; and Miller et
al., 1979).
     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:
Morphology  (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  bron-
chiolar epithelial  cells,  which  in  some species can  appropriately be called
"Clara" cells.  Subepithelial  tissues  are sparse and do  not contain glands.
     The sensitive cell of the upper and lower conducting airways is the cili-
ated cell, which is primarily responsible for physical clearance or removal of
inhaled foreign  material  from  conducting airways of  the  respiratory  system
(see Section 10.3.4).   The effects of (L on this sensitive 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  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).
0190PT/B                           10-26                                5/1/84

-------
10.3.1.1.1  Ai rways
     10.3.1.1.1.1    Upper airways (nasal cavity, pharynx, and larynx).   The
effects of 0- on the upper extrathoracic airways have received little attention.
The effect  of  upper airway scrubbing on the  level  of 0- 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
ambient concentrations between 196 and 3920 pg/m  (0.1 and 2.0 ppm).   Earlier,
Yokoyama and  Frank (1972)  studied nasal  uptake  in dogs.  They  found  uptake to
                                                                           3
vary with flow  rate as well as with  0^ concentration.   At 510 to 666 (jg/m
(0.26 to 0.34 ppm) of 0^, 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 ±
2.7 percent.  At 1529  to 1568 ug/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.
     The effects of 392, 1960, and 9800 pg/m3 (0.2,  1.0, and 5.0 ppm) of 03 on
the tonsils, the primary lymphoreticular structures  of the upper airways, were
studied.   In  palatine  tonsils  from rabbits exposed to 392 |jg/m  (0.2 ppm) of
0- continuously for 1 and 2 weeks, Ikematsu (1978)  reported epithelial  detach-
ment and disarrangement and a slight  cellular  ^filtration.   The significance
of these observations  to the function of immune mechanisms in  host defense is
unknown.
     10.3.1.1.1.2  Trachea.  Tracheal  epithelial lesions have been described
                                                             3
in several  species following exposure to less than  1960 (jg/m  (1 ppm) of 0,..
Boatman et  al.  (1974)  exposed anesthetized, paralyzed cats to 510,  980, or
         3
1960 pg/m   (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 tracheal concentrations than in usual exposures.  They reported desquama-
                                         3
tion of ciliated epithelium at  1960 pg/m   (1 ppm)  of 0~, but  none at 510 or
980 jjg/'m3 (0.26 or 0.5 ppm).
0190PT/B                           10-27                                5/1/84

-------
     In rats exposed to  960 or 1568 ug/m   (0.5  or  0.8 ppm) of 0-, 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
                                                                           '3
electron microscope  (SEM)  and  were not obvious  in rats exposed to 392 ug/m
(0.2 ppm) of GO  for the  same times.   Cavender et al.  (1977),  when using only
light microscopy  (LM), studied tracheas from rats and guinea pigs exposed  for
7 days to 1960 or 3920 ug/m   (1 or 2 ppm) of 03 or sulfuric acid (H2$04)  or
both.  The  article  does  not state the hours of  exposure  per  day.  Tracheal
lesions, which consisted  of reduced  numbers of  cilia and goblet  cells, were
reported only for guinea  pigs exposed to 0.,.   Animals exposed  to both pollutants
had lesions  similar to those exposed to 03 alone.
     By using SEM and transmission electron microscopy (TEM),  Castleman et al.
(1977) described  shortened  and less  dense cilia in  tracheas from bonnet mon-
                                3
keys exposed to  392 or 686 ug/m   (0.2 or  0.35 ppm), 8 hr/day  for  7 days.
Lesions occurred  as  randon 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.
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 156o ug/m  (0.5  or 0.8 ppm) of
03  in  the same exposure  regimen (Dungworth et al.,  1975b;  Mellick et al.,
1977).
                                          3
     After  mice  were  exposed to 1568 ug/m  (0.8 ppm)  of 03 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 surface of the  tracheas  was
examined by SEM  by  Ibrahim et  al.  (1980).   Short and  normal-length cilia were
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
0190PT/B                           10-28                                5/1/84

-------
clusters of nonciliated  cells  were still present.  Earlier,  Penha  and Wer-
thamer (1974)  observed metaplasia of the tracheal  epithelium from mice exposed
                                        3
to high concentrations of  03 (4900 ug/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.
     10.3.1.1.1.3  Bronchi.  Bronchial  lesions  were studied  in  many  of the
same animals  as  those whose tracheal lesions are described above and were
reported as generally similar to the tracheal lesions.  At  low concentrations
(Castleman et  al., 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 pg/m  (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 Mel lick
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.  Sato et al.  (1976a,b)
studied bronchi from vitamin E-deficient and control rats exposed to 588 pg/m
(0.3 ppm) of  03  3 hr/day for up to 16 days.  Using LM, they did  not see bron-
chial  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) exposed vitamin E-deficient
                                 3
and supplemented  rats to 588 ng/m  (0-3 ppm) of 0- 3 hr/day, 5 days/week for 7
months  following  which  they  did not see clear  differences  with  SEM or TEM.
     10.3.1.1.1.4  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
bronchroles).   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

0190PT/B                           10-29                                5/1/84

-------
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  (Castleman  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 03 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.
     Terminal bronchiolar lesions  in rats due to inhalation of <  1960 pg/m  (<
1.0 ppm)  of  DO  for  2 hr to 1 week have  been  des  ribed  by Stephens  et al.
              •j
(1974a),  Evans   et  al.   (1976a,c), Schwartz  et  al.  (1976),  and others
(Table 10-1).   Ciliated  cells are  the most  sensitive 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., 19766.;  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 granularity.   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

0190PT/B                           10-30                               5/1/84

-------
and precursors  of  basal  bodies indicative of  ciliogenesis  (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 genera-
tion,  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/m  (0.54 or 0.88  ppm)  of  03  for  4  hr  to  3 weeks.
In addition to the centriacinar accumulations  of macrophages and the hyperplasia
of  the distal  airway  epithelium,  they  reported  an increase  in connective
tissue elements and collagen-like strands which formed bridges across alveolar
                            3
openings.  In  the  1725-ug/m  (0.88-ppm)  03 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                                                        i
and one of 1690 |jg/m  (0.9 ppm) nitrogen dioxide,  or to a mixture of 490 ug/m
(0.25  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  expo-
sures  (Freeman  et  al.,  1972).   Freeman et al.  (1974) did  not  report emphysema
in rats exposed to 03 alone, only in those rats exposed to the 0.9/0.9 mixture.
The topic of emphysema is discussed later (Section 10.3.1.4.2).
     Results differ in three studies of  long-term (3- to 6-month) exposures of
                    3
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
using  only LM  to study  effects in rats exposed for  6 months to  980 (jg/m3  (0.5
ppm) for 6 hr/day,  Cavender et al.  (1978) did r t find exposure-related lesions.
Using  SEM and  TEM  to  supplement the  LM  technique,  Boorman et al.  (1980)  and
Moore  and Schwartz (1981)  reported  significant bronchiolar lesions following
exposure to  980  or 1568 ug/m3 (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
luminai 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 bronchiolar 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.
0190PT/B                           10-31                                5/1/84

-------
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.  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 morphology persisted, the changed
segment  reduced in length after the 180-day-exposed rats had breathed filtered
air for  62 postexposure  days.   The addition of HpSO. to these concentrations
of 0.,  for  the  same exposure times did not potentiate the lesions seen in the
    O
03-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
                                                                            3
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 03 levels (980
or 1568  pg/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.
     Guinea  pigs were  exposed  by  Cavender  et  al.  (1977,  1978)  continuously to
1 or 2 ppm of 0,. for 2 or 7 days in acute  studies and to 980 ng/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-c, .rway  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
HUSO,  aerosol  and 03-
10.3.1.1.2  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.
 0190PT/B                            10-32                                 5/1/84

-------
     10.3.1.1.2.1   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 0,  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).
Short-term  exposures  of  monkeys to 392, 686,  980,  or  1568  ug/m   (0.2,  0.35,
0.5,  or 0.8 ppm) of 03 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 07 concentration.
                                                       o
     The pathogenesis  of  the lesions  due  to  1568  ug/m  (0.8 ppm) of 03 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, 8,  and 12 hr of
exposure.    Type  1  cell necrosis, which  resulted  in  bare  basal  lamina,  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
stem cells  in  renewal  epithelium, and both contributed to the hyperplasia seen
at the  latter  exposure times.   The inflammator^  exudate included both fibrin
and a  variety  of leukocytes in the early  phases.   In the latter  phases, the
inflamnatory 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.
     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 al. ,  1981).  After  the  90-day exposure, the  percent of
cuboidal respiratory  bronchiolar epithelial  cells  was  90 percent rather than

0190PT/B                           10-33                                5/1/84

-------
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  seven-fold  those  of controls.   This study did not include  a
postexposure period.
     In an earlier study, Freeman et al. (1973) exposed female beagle dogs  to
         2
1960 ug/m  (1 ppm) of  0~  8, 16,  or 24 hr/day for  18 months.   Dogs exposed to
         3
1960 ug/m  (1 ppm)  of 03 for 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
and proximal alveoli.   Epithelial  hyperplasia was seen  sporadically  in  the
bronchiolar-ductal  zone.  Other dogs  in that study exposed  to  3920  or 5880
ug/m  (2 or 3 ppm)  of 0- 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.
     10.3.1.1.2.2  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  al.,  1974b; Schwartz et al., 1976; Mellick et al.,
1977).  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  th.ckening  of  the interalveolar
septa.  In animals with respiratory bronchioles (e.g., dog, monkey) the alve-
oli involved at low concentrations are those opening directly into the respira-
tory 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
                                            3
al., 1977).   In monkeys exposed tp 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.
0190PT/B                           10-34                                5/1/84

-------
     Type 1 cell  damage  and loss have been reported in rats after a 2-hr ex-
                   3
posure to 392  (jg/m  (0.2 ppm) of 0~ (Stephens et al., 1974a).   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
        3
980 ug/m  (0.35 or 0.5 ppm) of 0, (Evans et al., 1976b).   Although the labeling
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 follow-
                                     3
ing 90 days  of exposure to 1568 ug/m   (0.8  ppm) by Boorman et al.  (1980).
They interpreted  these intermediate cells  as due  to  delay or arrest of the
transformation 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
                   3
exposed to  490 ug/m  (0.25  ppm) of  03 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.  Sherwin
et al.  (1983)  found increased numbers of lactate dehydrogenase (LDH)-positive
cells,  presumed to be type 2 alveolar epithelial cells, by automated LM morph-
                                               3
ometry of lungs from  mice exposed to 588  ug/m   (0.3  ppm)  of 03 for  6 weeks.
Moore and Schwartz (1981)  reported nonciliated bronchiolar cells  lined some
alveoli opening into  the transformed airways located between terminal bronchi-
oles and alveolar ducts of rats exposed to 1568 ug/m  (0.8 ppm)  of 03 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
                                3
early as after 4  hr of 1568 ug/m   (0.8  ppm) of  J,  exposure  (Castleman et  al.,
1980).   In  rats,  the  numbers  of inflammatory cells per centriacinar alveolus
                                                                             q
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 \ig/m  (0.5 ppm)  of 03 and after 62 days  of filtered
air following  180  days of exposure.  The  addition of H?S04 aerosol  did not
result in larger  increases.   In  monkeys,  the  intensity  of the  response was
greater than  in rats  exposed  to  the same  03  concentration (1568  ug/m3,  0.8
ppm) in the  same  regimen (8 hr/day) for 7,  28,  or 90 days (Eustis  et al.,
0190PT/B                           10-35                                5/1/84

-------
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).
     The interalveolar septa of centriacinar alveoli  are thickened following
exposure to ambient  concentrations  of  0~.   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
in rats exposed to 1568 pg/m  (0.8 ppm) of 0- 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.  This study  is especially
important, because it  is quantitative and concerns rats exposed only 8 hr/day
            3
to 1568 ug/m  (0.8 ppm) of 0,.
                                                                3
     Moore and  Schwartz  (1981),  after exposing rats to 980 ug/m   (0.5 ppm) of
03  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
dogs exposed to 1960 to 5880 ug/m3 (1 to 3 ppm) of 03 8 to 24  hr/day for 18
months.
     In several biochemical studies, Last and colleagues have shown that 0~ is
collagenic.  In one of these (Last et al., 1979), Me biochemical observations
were correlated with  histological  observations of slides stained  for collagen
and reticulin.   Elevated collagen synthesis rates were found at all concentra-
tions and  times studied.   Mildly increased amounts of  collagen  were seen
morphologically in centriacinar alveolar duct septa from most  rats exposed to
        3
980 ug/m   (0.5  ppm)  of 0, 24 hr/day for 14 or 21 days.  More  severe  lesions
                                              3
were seen in rats exposed to 1568 to 3920 ug/m  (0.8 to 2.0 ppm) of 0, 24 hr/day
for 7,  14, or 21 days.
     Only one  published  report  addressed the biologically important question
of  the  morphological  effects that follow multiple sequential exposures to 03
with several  days  of clean air  interspersed between  0- exposures (i.e., a

0190PT/B                           10-36                                5/1/84

-------
multiple episode exposure regime).  Plopper et al. (1978) exposed rats to 1568
    3
ug/m  (0.8 ppm) 0- continuously for 3 days, held them in the chambers breathing
filtered air until postexposure day 6, 13, or 27, when they were again exposed
            3
to 1568 pg/m  (0.8 ppm) of 0, 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  0_ 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 the  first exposure.
10.3.1.1.3  Vasculature, blood,  and lymphatics.  Although edema is the apparent
cause of death  due  to inhalation of high concentrations of 0-, there is very
                                                                             3
little morphological  evidence of pulmonary vascular damage due to < 1960 ug/m
U 1-0 ppm) of  Oo exposure.   Bils  (1970) reported capillary endothelial damage
                                                              3
in mice less than 1 month of age exposed for 7 hr to 1960 ug/m  (1 ppm) of 03>
but this experiment  has  not been confirmed by others.  Boatman et al. (1974)
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 03 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 endothelium  in both groups
exposed to  784 ug/m   (0.4  ppm)  of  0,  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  ^30  or 1764 ug/m  (0.5 or 0.9
                                      3
ppm) of 0,  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 0_ 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.
0190PT/B                           10-37                                5/1/84

-------
                                                                           3
     The earlier report of  arterial  lesions in rabbits exposed to 784 ug/m
(0.4 ppm) of 03 6 hr/day, 5 days/week for 10 months by P'an et al.  (1972) has
not been confirmed by other investigators,  because no other long-term exposure
of rabbits has been  reported.   Possibly, susceptibility to arterial lesions
was specific to this animal  species,  or these rabbits may have had  intercurrent
infectious 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."
     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  0-, blood capillary  endothelial  damage has not been  reported, and
             <3
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.
10.3.1.2  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) found minimal type 1 alveolar epithelial cell damage
                                 3
after a 2-hr exposure to 392 ug/m  (0.2 ppm) of 0- and more significant damage
                               3
2 hr after exposure to 980 ug/m  (0.5 ppm) of 03-   Boatman et al. (1974) found
lesions  in both the  conducting  airways  and paren  iyma  of  cats exposed to 510,
                  3
980, or  1960 ug/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).
Severa'  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, Mel lick
0190PT/B                           10-38                                5/1/84

-------
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 |jg/m  (0.5 or 0.8 ppm)  of 0- 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 pg/m  (0.2
and 0.35 ppm)  of  03 8 hr/day for 7  days.   In  the centriacinar  region, this
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  0~.
                                                  •3
     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  al., 1974a,b; Castleman et al., 1980).    Repair  of the damage,
as indicated by DMA 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 dayi 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).
10.3.1.3.   Structural Elements Affected
10.3.1.3.1  Relative sensitivity  of cell  types.   The  ciliated cell  of  airways
and the type 1 (squamous) alveolar epithelial  cell are the cells most sensitive
to the effect  of  inhaled C<3  (Stephens et  al.,  1974a,b; Schwartz  et al. ,  1976;
Mel lick et  al.,  1977;  Castleman  et  al.,  1980; Eustis et  al. ,  1981).   The
amount of damage  to  these sensitive cells varies with their position  in the
0190PT/B                           10-39                                5/1/84

-------
respiratory system.   As mentioned  earlier,  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 cells in terminal  bronchioles of animals without  respiratory
bronchioles (i.e.,  rats)  are severely damaged by even low concentrations of 03
(Stephens  et al., 1974a;  Schwartz  et al.,  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 al.,  1977).
     In  a  similar  manner, type 1  alveolar  epithelial  cells located in the
centriacinar region  are  subject  to damage  by  low  concentrations of 03, 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; Castle-
man et al., 1980).
     Although type 2 alveolar epithelial  cells are relatively resistant to 03,
some in  centriacinar locations develop mild  lesions  detectable  with the  TEM
(Castleman  et  al.,  1980).  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 03  lesion
(Evans et  al.,  1976b).
     Although  nonciliated cuboidal  bronchiolar cells  are more resistant to 03
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 03 damage (Schwartz  et al.,  1976).   Following exposure  to 1960
ug/m3  (< 1 ppm) of  03, their  height is  reduced and their luminal  surface is
more granular  (Schwartz  et al.,  1976).   The red ction in height  appears to be
due to  a loss  of smooth endoplasmic reticulum  (Schwartz  et  al.,  1976).
     Several investigators  report that type  3  alveolar  epithelial cells,  the
brush  cells, are not damaged by  less than 1 ppm of 03  (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 much less sensitive 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).   In  the earlier  reports,  damaged endothelial
cells  were those located  immediately deep to denuded  basal  lamina and  resulted

0190PT/B                           10-40                                 5/1/84

-------
from sloughing of  type  1 epithelial cells in the centriacinar 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 CL 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).   Only  one morphometric
evaluation of the  centriacinar interalveolar septa is in the open literature
(Boorman et  al., 1980),  and  it does  not provide  detail  concerning specific
cell types and extracellular materials.
     Mucous-secreting cells  in conducting airways appear relatively resistant
to 0.,.   Boatman  et al.  (1974), in studies of cats exposed to < 1960 ug/m  (g
1.0 ppm) of  0,, 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
mucous  cells from  monkeys exposed  to  392 or 686 ug/m   (0.2  or  0.35 ppm) of 03
8 hr/day for 7 days.  Mellick  et  al.  (1977) mention that mitochondria! 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 ug/m  (0.5 or 0.8 ppm)
of 0-  8 hr/day   for 7 days.   Schwartz et al. (1976),  who  reported mucigen
droplets being released  from the  apical surfaces of  mucous  cells and mucous
droplets trapped among cilia, did not find charjes suggesting damage to organ-
el les 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.  No reports  of damage  to conducting airways other
than to ciliated and mucous cells  were found.
10.3.1.3.2  Extracellular elements (structural  proteins).  Although physiologic
and biochemical  changes  following 0- exposure suggest changes in the extracell-
ular 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.   Three studies
provide morphological evidence of  mild fibrosis (i.e.,  local  increase  of
0190PT/B                           10-41                                5/1/84

-------
collagen) in centriacinar interalveolar septa following prolonged exposure to
           3
< 1960 ug/m  (< 1 ppm) of 03  (Last et a!.,  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).   Additional evidence
of more  collagen or  changes in collagen location  is  in  the report  of dogs
exposed to 1960 or  5880  ug/m  (1 or  3 ppm)  of 03 for 18 months  (Freeman et
al., 1973).
10.3.1.3.3  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
(^ 1.0 ppm) of 0,  for  short or long exposure periods (Schwartz et al.,  1976;
Cavender  et al. , 1978;  Mel lick 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  0-  for several hours
to 1 week.  Interstitial  edema is  not reported following longer-term (i.e.,
                                        3
weeks to  months) exposure to  < 1960 ug/m  (< 1 ppm) or less  (Cavender et al.,
1977; Eustis et al.,  1981;  Boorman et al.,  1980;  Moore  and Schwartz,  1981;
Zitnik  et al.,  1978).   Biochemical  indicators of  edema are described in
Section 10.3.3.
10.3.1.4  Considerations of Degree of  Susceptibility to Morphological Changes
10.3.1.4.1  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 respr .d  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.
     10.3.1.4.1.1  Vitamin E deficiency.  Rats maintained on vitamin E-deficient
diets tended to develop  more morphological  lesions following 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
                                                                  3
average  U. S.  human adult intake were  exposed  to 196 or  392 |jg/m   (0.1 or
0.2 ppm)  of OT  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
0190PT/B                           10-42                                5/1/84

-------
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 pg/m  (0.1
ppm) Oo for the same period, LM revealed no increased centriacinar macrophages.
      O
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
                    o
exposed to  196  pg/m  (0.1 ppm)  of 0-j  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
they were  exposed either to filtered  air or to 196 yg/m   (0.1 ppm)  of 03
continuously for 7 days.   The morphology of six rats from each diet and exposure
group was studied  using  SEH,   None of  the  filtered-air  control  animals had
lesions.  Of the rats exposed to 0^,  five of the six on the vitamin E-deficient
                                  O
diet, four  of  six  on the  ceficient 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  0., 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 pg/m  (1.3 ppm) of 0, 3 hr daily for 16 consecutive days or
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 03 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, Og-exposed rats.  They did find mild centriacinar
0-  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 M9/m  (0-9 ppm)  of 03 for  1,  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
0190PT/B                           10-43                                5/1/84

-------
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 10.3.3.
     10.3.1.4.1.2   Acje___at  sta_rt__of__gxp_os.ure.    Although most  exposures  use
young adult  experimenter animals,  there are a  few  reports  of  exposures of very
young animals  (i  ». , either before weartinc or  very soon thereafter).
     Bartlett  ^t  al.  (197'--) exposG-5  3- to 4-week-old rats to 392 pg/m   (0.2
ppm) of Q,  for 30 days.  Lung  volumes,  • ut not  body weights,  were significantly
greater in the exposed  rats.   Light w>;:"o$>cepy  of  paraffin sections  of conven-
tionally  fixed lungs  did  net  rovudl  drr';:eri:ices  between  exposed and  control
rats in the  lung  parenchyma cr '.en; i •>;• bro.,-:,n"o seas  with  the exception of two
control animals which had  'lesi>.'.v-.  -j'= "•*,yp1r&i  in wine  pneumonia."  Morphometry
was done  on thick sections art by  nar.d with  a  razor blade from the dorsal and
lateral surfaces  of aJr-d'^'Osj h.-j.^b  >. a'.".",*-jr  ;han  c;; the paraffin sections of
conventionally fix-^d  'uncjs. Hor,.i'.o,T,et.ry tiaincj  these  nonrandom samples revealed
significantly  increased mean ah/ec'-a" chord "engths and alveolar surface area,
but no difference  in alveolar  nu^bess.
     Freeman et al.  (1974: 9*pr,s«.G irontrrolri  rats to  ]058 or 1725 |jg/m'  (0.54
or 0.88 ppm) of 0^ for  period-. ;f iVorr. 4  irr  to 3 weeks.   In addition to the
centriacinar accumulation? of iTi<:.C'Ophoget> and hyperp'lasia of  distal  airway
epithelium  seen by ethers  foT'.ovM.c  exposures  of young adult  animals, they
reported  an  increase in connect-t  c!.  H974) studied month-old rats exposed
             3                               3
to 1764 jjg/m  (0.3  ppiiij  of 0-   *•••  t:}u ir^/in'  (0.5  ppm)  nitrogen dioxide  combined.
                             ,';
After  (50  days  of expos-jr?, ^r,fJ -^'-..>. vvi5  the  yoss and microscopic appearance
of advanced  experimental e,nphy-:.tyi of \,he type xbey earlier described following
nitrogen  dioxide  exposure  (frecmsr; ei. al. , 1372).   Although others have  reported
larger  fixed  lung  volumes in   .:?xpo~,ee;  young  adult rats  (Moore and Schwartz
1981), reports of emphysema !V;'uvr.nc; 0,, exposures are  uncommon and are  discus-
sed in the  next subsection of   chir. document.
0190PT/B                            10-44                                 5/1/84

-------
     Stephens et al.  (1978)  exposed rats at 1,  5,  10,  20, 25, 30, 35, and
                            3
40 days of age  to  1666 ug/m  (0.85 ppm) of  0,  continuously for 24, 48, or
72 hr.   Rats exposed  to  0~ before weaning at 20 days of age developed little
or no  evidence  of  injury,  as evaluated by  light  and  electron microscopy.
Following weaning at 20 days of age, centriacinar lesions increased progressive-
ly, 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
ug/m  (0.25 ppm) of 0, 12 hr/day for 6 weeks.    They observed  persistence of
the centriacinar damage to type  1 epithelial cells  and  increased centriacinar
macrophages.   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 discus-
sed 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.  (1983) are discussed in Section 10.3.2.
     Bils (1970) studied the effects of 1176 to 2548 [jg/m3 (0.6 to 1.3 ppm) of
03 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 0^, a  conclusion
not supported by more recent studies,  which deal mostly with  other species.
Bils did  note the  lesions were more severe in the 4~day-old mice than in the
1- or 2-month-old mice.
     10.3.1.4.1.3   Effect of pneumonectomy.   Two  to four  weeks  following
pneumonectomy of rabbits,  the centralateral  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 pneuinonectomized and control rabbits to 784 (jg/m  (0.4 ppm)
of 0-, 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  0~  exposure and that no difference  existed  between males and
females in this  response.
10.3.1.4.2  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

0190PT/B                           10-45                                5/1/84

-------
                                 3
following exposure to < 1960 |jg/m  (< 1 ppm) of 0- 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.s  1980;  Wright et al., 1983)  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).   Thus, a restudy of these three reports  in the 1978
document appears appropriate.
     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/
week for 14.5  months to  a mean concentration of slightly more than 1960 pg/m
(1 ppm) of 0-.  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 (CIBA, 1959)  or enlarged air spaces accompanied  by
destructive changes in alveolar walls (World Health Organization,  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).   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 the rats,  the  mild degree  of  emphysema  in those
exposed "did not  exceed the emphysema found in the unexposed control rats."
In the  hamsters,  "mild to moderate"  emphysema was present in exposed animals,
but not 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 presentation at the seventh Aspen Conference on
Research in Emphysema,  reviewed the lesions in the hamsters from this exposure
and described  a  "destructive  process"  that resulted in contraction of inter-
alveolar septa not associated with enlargement of air spaces.
0190PT/B                           10-46                                5/1/84

-------
     The earlier 0.,  criteria  document (U.S.  Environmental Protection Agency,
                  «j
1978) cites Stephens  et  al.  (1976),  a "long abstract" that appears not to be
refereed.   This  brief article  states "rats exposed  continuously  for long
                                      3                             3
periods (3-5 months) to 28,200 ug/m (15.0 ppm) N02 or  1568 ug/m  (0.8  ppm)
0- develop a  disease that closely resembles emphysema" but does not  provide
 O
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 NO^-exposed animals.
The  fourth reference (Freeman et al., 1973) is  to an  exposure of dogs  to  1960
to 5880 |jg/m   (1 to  3 ppm)  of 0., 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 group of month-old  rats exposed continuously for
                     3
3 weeks to 1725  ug/m  (0.88 ppm)  of  03,  half  of which died and had  "grossly
inflated, dry lungs."  In this same study, they also exposed month-old rats to
                       3
a mixture  of  1690  yg/m  (0.9 ppm) of N02 and 03 continuously for  60 days, at
which time the lungs were grossly enlarged, and "both grossly and  microscopical-
ly,  the  appearance  of the lungs was  characteristic of advanced experimental
emphysema" of the type they earlier reported followed N02 alone at much higher
concentrations (Freeman et al., 1972).
     The third citation in the 0- criteria document (U.S.  Environmental Protec-
tion  Agency,  1978)  is to  P'an  et al. (1972).   These investigators  exposed
rabbits  to 784 ug/m  (0.4  ppm) of 03 6 hr/day, 5 days/week  for 10  months.
Tissues were  apparently fixed by immersion rather than infusion via the trachea,
which Is not  in accord with the American Thoracic Society's diagnostic standard
for  emphysema and  makes  emphysema lesions much more  difficult  to  accurately
evaluate.  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
congested  that visualization of the mural framework of the alveoli was difficult."
This  is  more  reaction than  reported  in  other  species  exposed  to this  compara-
tively low 0, concentration.  The rabbits were  not specified pathogen  free nor
            O
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.
0190PT/B                           10-47                                5/1/84

-------
10.3.2  Pulmonary Function Effects
10.3.2.1  Short-Term Exposure.  Results of short-term 0~ exposures of experi-
mental animals are shown in Table 10-2.   These studies were designed to evalu-
ate the acute changes in lung function associated with 0., 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  0- exposure of the lung periphery have
been examined in dogs by Gartner et al.  (1983a,b).   A fiber-optic bronchoscope
with an outside  diameter  of 5.5 mm was wedged into a segmental airway and a
continuous flow of 196 or 1960 ug/m  (0.1 or 1.0 ppm) of 03 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
         3
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
(1.0 ppm) of  0,,  but decreased again  to control  levels during continued expo-
                 3
sure to 196 ug/m   (0-1  PPm) °f  63-  Based on these observations, the authors
reported that tolerance appears  to develop in the collateral airways to locally
                                                       3
delivered 03 at concentrations of 196 but not 1960 (jg/m  (0.1 but not 1.0 ppm)
of 03.
     Amdur et al. (1978) measured breathing 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 Q-.   Accelerated  respiration rates with  no  significant
                       o
changes in  tidal  volume were measured  during exposures  to all 0^ 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
to 804 or 1568 ug/m  (0.41  or 0.8 ppm)  of 03> 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 0.,.   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.
0190Z4/A                           10-48                                 5/1/84

-------
                                          TABLE 10-2.   EFFECTS OF OZONE ON PULMONARY FUNCTION:   SHORT-TERM EXPOSURES
O
 i
Ozone
concentration Measurement :
ug/mj
196
1960
431
804
1568
470 to
2156
510
980
1960
666
1333
2117
2646
980
ppm method
0.1 MAST
1.0
0.22 CHEM,
0.41 NBKI
0.80
0.24 to NBKI
1.1
0.26 MAST
0.5
1.0
0.34 NBKI
0.68
1.08
1.35
0.5 NBKI
. Exposure
' duration
& protocol Observed effects0 Species
30 min Collateral system resistance increased Dog
rapidly during exposure, falling to
control levels at 0.1 ppm tout con-
tinuing to increase at 1.0 ppm of 03.
2 hr Concentrati on- dependent increase in fR Guinea pig
for all exposure levels. No change in (200-300 g)
R. , TV, or MV. Decreased C.dyn during
exposure to 0.4 and 0.8 ppm of 03.
12 hr Premature airway closure at 6 hr, and Rabbit
1 and 3 days following exposure, reflec-
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.
2.0 to Concentration-dependent increase in R. Cat
6.5 hr 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.
2 hr Increased fg and decreased TV during Guinea pig
exposure to all 0, concentrations. (300-400 g)
Increased R during exposure to 1.08
and 1.35 ppfo of 03.
2 hr Slight increase in fg and R (to 113% Guinea pig
of control values) during exposure. (280-540 g)
Reference
Gertner et al. ,
1983a,b
Amdur et al . , 1978
Inoue et al. , 1979
Watanabe et al . ,
1973
Murphy et al. , 1964
Yokoyama, 1969

-------
                                   TABLE 10-2.   EFFECTS OF  OZONE  ON  PULMONARY  FUNCTION:   SHORT-TERM EXPOSURES   (continued)
o
 i
en
o
Ozone
concentration Measurement
(jg/m3 ppm method
1470 0.75 CHEM
1960 1.0
3920 2.0


1960 1 NBKI




1960 1 NBKI


. Exposure
' duration
& protocol
Continuous
1, 3, 7 or
14 days


3 hr




6 hr/day,
7 to 8 days



Observed effects Speoes 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 ppm of 03).
Reduced TLC at air inflation pressure Rabbit Yokoyama, 1972, 1973
of 30 cm 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 MEFV curves.
            Measurement method:   MAST = Kl-coulometric (Mast meter);  CHEM - gas phase chemiluminescence;  NBKI = neutral  buffered potassium iodide.
             Calibration method:   NBKI = neutral buffered potassium iodide.
            cSee Glossary for the definition of pulmonary symbols.

-------
     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 (jg/m  (0.26, 0.50, or
1.00 ppm) of 0.,  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
           3                                                       3
to 510 |jg/m  (0.26 ppm) of 0-, after 63 min of exposure to 980 (jg/m  (0.50 ppm)
                                                3
of 0~,  and after 49 min of exposure to 1960 ug/m  (1 ppm) of 03-   Dynamic lung
compliance was  decreased during 0~ exposure  in  the  Watanabe et al. (1973)
study,  as it was  in the A«ndur 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
during 2-hr  0,  exposures.  These  investigators  found  concentration-related
                                                                             3
increases in respiration rate during exposure to 666, 1333, 2117 or 2646
(0.34, 0.68, 1.08, or 1.35 ppm) of O,.   Respiratory flow resistance was increased
(to 148 and 170 percent of pre-exposure values) in guinea pigs exposed to 2117
             3
and 2646 ug/m   (1.08 and 1.35 ppm) of 03 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  03  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

0190Z4/A                           10-51                                5/1/84

-------
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™.
     Inoue et al.  (1979)  exposed rabbits to 470 to 2156 ug/m3 (0.24 to 1.1 ppm)
of 0-  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,  1 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
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  0~-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 O^-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 HLO, 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

0190Z4/A                           10-52                                5/1/84

-------
(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
            3
to 1960 |jg/m   (1 ppm)  of 0~,  6 hr/day, for 7 to 8 days.  He found  increased
pulmonary resistance  and decreased dynamic compliance in CL-exposed animals
compared to air-exposed  control  animals.   Static pressure-volume curves and
maximum expiratory flow-volume curves were not significantly different between
the two groups.
10.3.2.2  Long-Term Exposure.   Table 10-3  summarizes results  of long-term 0~
exposures.   Raub et al.  (1983a)  exposed neonatal and 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  0^  exposure.   Peak inspiratory  flow measured  in  these
animals  during spontaneous respiration  was  significantly  lower  following
                           3
exposure to 235 or 490 ug/m   (0.12  or  0.25 ppm)  of 0~.   Lung volumes measured
at high  distending  pressures  were  significantly higher  in  neonatal animals
                   3
exposed to 490 ug/m   (0.25 ppm)  of 0-  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)  following exposure
                                            3
of young rats (3- to 4-week-old)  to 392 ug/m  (0.2 ppm) of 0,  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
                                                            3
fixative) after 180 days cf continuous exposure  to 980 ug/m (0.5 ppm)  of 0^.
Yokoyama and  Ichikawa (1974)  found no change in lung  static pressure-volume
curves in mature rats exposed to 882 |jg/m  (0.45 ppm)  of 0., 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 HpO was also reported following 0- exposure, which is
similar to the fixed lung volume  changes reported by Moore and Schwartz (1981).
Morphology and morphometry of paired lungs or lungs  from  animals  similarly
exposed to 03 is  reported in Section 10.3.1.4.
0190Z4/A                           10-53                                5/1/84

-------
                                            TABLE 10-3.   EFFECTS OF OZONE ON PULMONARY FUNCTION:   LONG-TERM EXPOSURES
o
en
Ozone
concentration
ug/mj
157
235
490
392
392
1568
3920
784
882
980
1568
ppm
0.08
0.12
0.25
0.2
0.2
0.8
2.0
0.4
0.45
0.5
0.8
Exposure
Measurement ' duration
method & protocol
CHEM 6 weeks,
12 hr/day,
7 days/week
MAST, 28 to 32 days,
N6KI continuous
UV or CHEM 62 exposures,
NBKI 6 hr/day,
5 days/week
NBKI 6 weeks,
7 hr/day,
5 days/week
MAST 6 to 7 weeks,
6 hr/day,
6 days/week
UV, 7, 28, or
NBKI 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
0_-exposed rats at high lung
vBlumes (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 and FEF25 and FEFto
were decreased at 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 quasi static compliance (not
related to concentration).
Species Reference
Rat Raub et al. , 1983
(neonate or
6-week-old
young adult)
Rat Bartlett et al. ,
(3 to 4 weeks) 1974
Rat Costa et al. , 1983
(10 weeks)
Rabbit Martin et al. ,
1983
Rat Yokoyama and
Ichikawa, 1974
Monkey Eustis et al . , 1981
(Bonnet)

-------
                                    TABLE 10-3.  EFFECTS OF OZONE ON PULMONARY FUNCTION:  LONG-TERM EXPOSURES   (continued)
o
 i
01
Ozone
concentration
ug/nr1 ppm
1254 0.64
Exposure
Measurement ' duration
method & protocol
UV, 1 year,
NBKI 8 hr/day,
7 days/week
Observed effects Species
Following 6 months of exposure, venti- Monkey
lation was less homogeneous and R, (Bonnet)
was increased. Following 12 months
of exposure, R. remained elevated
and forced expiratory maneuvers showed
small airway disfunction (decreased
FEVj and FEFl2.s). During the 3-month
recovery period following exposure,
C.st decreased.
Reference
Wegner, 1982
             Measurement method:  MAST = Kl-coulometric (Mast meter); CHtM = gas phase chemiluminescence; UV -  UV  photometry.


             Calibration method:  NBKI = neutral buffered potassium iodide.


             See Glossary for the definition of pulmonary symbols.

-------
                               PRELIMINARY DRAFT
     Costa et al.  (1983)  evaluated  lung function changes in rats exposed to
                       3
392,  1568, or 3920 ug/m  (0.2,  0.8,  or 2 ppm)  of 0.,,  6 hr/day,  5 days/week for
62 exposure days.   (This report will  not discuss effects in  animals  exposed to
2 ppm of  0~).  These  investigators  found increased pulmonary resistance (not
                                                           3
concentration related)  in  rats  exposed to 392 or 1568 ug/m   (0.2 or 0.8 ppm)
of 0-j.  Lung  volumes  measured  at high distending pressure  (VC and TLC) were
                                          3
increased following exposure to  1568 ng/m  (0.8 ppm) of 0,.  Similar changes
in lung distensibility  were  observed by Raub  et al.  (1983a), Bartlett et al.
(1974),  Moore and  Schwartz (1981),  and Martin et al.  (1983).   Costa et al.
(1983) also observed  decreased (not  concentration-related)  maximum expiratory
flows at  low  lung  volumes  (25 and 10 percent of VC)  in  rats  exposed to  392  or
1568 |jg/m  (0.2 or 0.8 ppm) of  0,.   Changes in maximum flow  at low lung volumes
                                O
indicate peripheral airway dysfunction and may be related to decreased  airway
stiffness or narrowing of the airway lumen.
     Eustis et al. (1981)  evaluated  lung function in bonnet monkeys (Macaca
radiata) exposed to 980 or 1568 ug/m  (0.5 or 0.8 ppm) of 03 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, quasistatic lung  compliance was significantly  different  in
Oo-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
were  exposed  to 1254  ug/m0 (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 (DITQ)>
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.
0190Z4/A                           10-56                                5/1/84

-------
                               PRELIMINARY DRAFT
     Following six months  of 0~ exposure, pulmonary resistance and frequency
dependence of  pulmonary compliance were  significantly  increased.   After 12
months5  the CL exposure had significantly increased pulmonary resistance and
             O
inertance (related to the pressure required to accelerate air and lung tissue),
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 CL-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 0.,-exposed and control  animals.  However, the
decrease in compliance was significantly greater in CL-exposed animals than in
control  animals.   No other significant differences were measured following the
3-month  recovery  period, although  values for 0,-exposed  animals  remained
substantially different  from  those for control  animals, suggesting that full
recovery was not complete.
10.3.2.3  Airway Reactivity.  Ozone potentiates  the effects of drugs that con-
strict airway  smooth muscle in mice,  guinea  pigs,  dogs, sheep, and  humans
(Table 10-4).   Early experimental  evidence for  hyperreactivity to  broncho-
constrictive drugs following  00 exposure was provided  by  Easton  and Murphy
(1967).   Although much of their work was done with very high Oq concentrations
                   3
(9800 to  11760 ng/m  ,  5  tc  6  ppm),  they  did show that mortality from  a single
subcutaneous injection  of  histamine was higher  in guinea pigs exposed to 980
or 1960 (jg/m  (0.5 or  1 ppm) of 0~  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.
     Abraham et al.  (1980)  evaluated  airway reactivity  in  sheep from  measure-
ments of pulmonary resistance following inhaled  carbachol aerosols.   Carbachol
causes bronchoconstrictior  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

0190Z4/A                           10-57                                5/1/84

-------
                                        TABLE 10-4.   EFFECTS OF OZONE ON PULMONARY FUNCTION:  AIRWAY  REACTIVITY
o


-I
CDl
Ozone
concentration
ug/m3 ppm
196 to 0.1 to
1568 0.8
196 0.1
1568 0.8
196 0. 1
1960 1.0
980 0.5
980 0. 5
1568 0.8
980 0.5
2156 1.1
980 0.5
1960 1-
3920 2
Exposure
Measurement3 duration
method & protocol
CHEM 1 hr
CHEM 1 hr
MAST 10-30 min
CHEM 2 hr
MAST Continuous,
13 to 16 days
of 0- 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
Subcutaneous injection of histamine 2 hr following 0,
exposure caused a greater increase in R, following expo-
sure to 0.8 ppm of 03 and a greater decrease in C.dyn
following exposure to all 0., concentrations (magni-
tude of C. changes not related to 0., concentration).
Decreased diaphragm and lung cholinesterase activity;
parathi on-treated animals had increased peak airway
resistance compared to controls, but the difference was
not statistically significant.
Bilateral vagotomy: completely Dlocked increased peri-
pheral lung resistance from 0.1 ppm of 03 but not histamine;
only partially blocked response from 1.0 ppm of 03.
Hi stamine- 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
lavage 24 hr after exposure.
Repeated exposures to 0.5 or 0.8 ppm of 03 plus aerosolized
ovalbumin resulted in greater mortality from anaphy lactic
shock produced by intravenojection of ovalbumin compared
with effects of ovalbumin injection in mice repeatedly
exposed to ovalbumin aerosols but no 0,.
«J
Increased histamine- induced mortality immediately
following exposure to 0.5 or 1.1 ppm of 03.
Increased airway reactivity to aerosolized carbachol
24 hr but not immediately following exposure to
980 ug/m3 (0.5 ppm) of 0., with no change in R, ,
FRC, C.st, or tracheal mucous velocity. Increased
R, 24 nr following exposure and airway reactivity
immediately and 24 hr following exposure (1 ppm).
Species Reference
Guinea pig Gordon and Amdur,
(200-300 g) 1980
Guinea pig Gordon et a!., 1981
Bog Gertner et al. ,
1983a,b,c
Kaplan et al . ,
1981
Sheep Sielczak et al., 1983
Mouse Qsebold et al.,
1980
Guinea pig Easton and Murphy,
1967
Sheep Abraham et al., 1980

-------
                                     TABLE 10-4.   EFFECTS OF OZONE ON PULMONARY FUNCTION:   AIRWAY REACTIVITY  (continued)
Ozone
concentration Measurement
ug/m3 ppm method
1100 to 0.56 to CHEM
1666 0.85
Exposure
duration
& protocol Observed effects Species
2 hr Abnormal, rapid, shallow breathing in conscious dogs Dog
while walking on a treadmill following 0, exposures.
Maximal 1~ to 3-hr postexposure, normal z4-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 .
Reference
Lee et al. , 1979
      1313
O
i
0.67
CHEM
                                       2 hr
Abnormal, rapid, shallow breathing during exposure to

air containing low 02 or high CQz immediately following

O- exposure.   Abnormal breathing not affected by

inhaled atropine aerosols or inhaled isoproteronol
Dog   Lee et al., 1980
aerosols but abolished by vagal cooling.
1372 to 0.7 to CHEM 2 hr
2352 1.2 j I
1960 1.0 ! UV 2 hr
4312 2.2
5880 3.0
Greater increase in R, caused by histamine aerosol Dog
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
bronchoconstrictlon) prevented 0.,-induced reactivity
to histamine.
Marked increase in airway responsiveness to ACh and Dog
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.

Lee et al . , 1977
Holtzman et al. ,
1983a,b
Fabbri et al . , 1984
       Measurement method:   MAST = Kl-coulometric (Mast meter);  CHEM = gas phase chemiluminescence;  NBKI  = neutral  buffered potassium iodide.

       See Glossary for the definition of pulmonary symbols.

-------
but not significantly  higher  than  pre-exposure values at 24 hr postexposure
                                                          3
but not immediately  following  a 2-hr exposure to 980 |jg/m  (0.5 ppm) of Ov
                                                                          O
This 0-, exposure did not affect resting end-expiratory lung volume (functional
      0                                                                     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.
     Gordon  and  Amdur (1980) evaluated airway  reactivity  to  subcutaneously
                                                                               3
injected histamine in guinea pigs following a 1-hr exposure to 196 to 1568 ug/m
(0.1 to 0.8  ppm) of  0-.  Airway reactivity to histamine was maximal 2 to 6  hr
following  0^ exposure and returned  to control levels by 24 hr following exposure.
The histamine-induced  increase  in  pulmonary  resistance was greater in guinea
                         3
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
0.,-exposed groups than  in air-exposed controls, but there were no differences
in the  the histamine-induced decreases in pulmonary compliance between any  of
                                            3
the 0» concentrations (from 196 to  1568 ug/m ; 0.1 to 0.8 ppm).
     Gordon  et al.  (1981)  studied  the effect of 0^ 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 1,568 ug/m   (0.1  or 0.8 ppm) of 03 for I hr.
After 2 hr,  brain,  lung,  and diaphragm samples were  analyzed  for cholines-
terase  activity.   Brain cholinesterase activity was  not  affected,  but lung
cholinesterase underwent a 17 percent decrease in activity at 196 pg/m  (0.1 ppm)
                                         3                                3
and  a  16  percent decrease  at  1,568 fjg/m   (0.8 ppm).   Ozone at 1,568 ug/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 resistance
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 cholinesterase
inhibition by  0~  may contribute to the 0.,-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  Oo  exposure,  should then
increase the contraction of the bronchus.   The persistence of this activity is
not known.
0190Z4/A                           10-60                                5/1/84

-------
     Lee et al.  (1977)  evaluated airway reactivity  in  0.,-exposed  dogs from
changes in pulmonary resistance induced by histamine aerosol inhalation.  Dogs
were exposed to  1372  to 2352 ug/m  (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-
                O
teronol aerosols prevented any change in resistance following histamine exposure.
This experiment showed that the increased resistance normally observed follow-
ing 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) de-
creased the response  to  histamine both before and following 0- exposure and
abolished  the  hyperreactive  airway  response.  These experiments showed that
the increased  sensitivity  to  histamine following  0, exposure was  caused  by
heightened activity of vagal  bronchoconstrictor reflexes.
     Although the work of Lee et al. (1977) suggests that stimulation of vagal
reflexes by histamine is responsible for the increased airway reactivity found
in dogs following ozone exposure, Kaplan et al.  (1981) found that local respon-
ses 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 bron-
choscope,  collateral  airflow  resistance  increased  both  before and after bila-
teral cervical vagotomy.   In  follow-up studies that used similar techniques,
Gertner et al.  (1983a,b,c) described the role of vagal  reflexes in the response
of the  lung periphery to locally administered histamine  and Ov  Collateral
                                                                          3
resistance increased  during   separate  30-min exposures  to  either 196  |jg/m
                           -fi     3
(0.1 ppm)  of 0.,  or  1.5 xlO   mg/m  of histamine.   However, although parasym-
pathetic blockade (atropine or  bilateral cervical  vagotomy)  prevented  the re-
sponses to 0~, it did not  prevent the  responses to histamine (Gertner  et al.,
                                                    3
1983b).  In addition,  a  30-min exposure to 196 ug/m  (0.1 ppm) of 0- did not
affect  the airway responsiveness  to histamine,  but when  the 0, exposure was
                      3
increased  to 1960 ug/m   (1.0  ppm)  for  10 min, histamine produced greater in-
creases in collateral  resistance  that  were  not abolished by parasympathetic
                                                        3
blockade (Gertner et al.,  1983c).   Exposure  to 1960 |jg/m  (1.0  ppm)  of Q~ for

0190Z4/A                           10-61                                5/1/84

-------
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) 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 resis-
tance, 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  the lung may  respond differently  from  the  larger conducting
airways during exposure  to  G~  and that factors in addition to vagal broncho-
constrictor reflexes can produce an increased  airway reactivity  to  histamine.
     Holtzman  et  al.  (1983a)  reported the time  course  of  0,-induced  airway
                                                     3
hyperreactivity in dogs exoosed to 1960 and 4312 (jg/m  (1.0 and 2.2 ppm) of 0^
for 2 hrs.  Airway responsiveness to acetylcholine in 7 dogs increased markedly
                                                           3
1 hr, to  a lesser extent, 24 hr after exposure to 4312 ug/m  (2.2 ppm) of 0-,
returning to control levels by 1 week after exposure.   Ozone-induced increases
in airway responsiveness to  histamine  were similar  following  exposure to
          3
1960 ug/m   (1.0 ppm)  of  03,  but  data were reported for only  2 dogs.   The
authors suggested that the time course of the 0- 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-
                                                           3
mation  in dogs 1 hr following a 2-hr exposure to 4116 ug/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 a!.,  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
                                                     3
acetylcholine  following  a  2-hr  exposure to  5880 ug/m   (3.0 ppm) of 0,.   The
                                                        3
results of trachea!  lavage in sheep exposed to  980 jjg/m  (0.5 ppm) of 03 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
 0190Z4/A                           10-62                                5/1/84

-------
(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,
suggesting an  association between an enhanced inflammatory  response  and 0.,-
induced  bronchial  reactivity reported previously  in  sheep (Abraham et al.  ,
1980).   The  relationship  between  airway  inflammation and  airway reactivity  in
other studies at lower concentrations of 0~ is not well understood.
     Increased drug-induced  bronchoconstriction  is not the only indicator of
airway hyperreactivity following 0- 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 11).   Lee et al.
(1979, 1980) showed that  abnormal,  rapid,  shallow breathing  in conscious dogs
                                                          3
immediately  following  2-hr  exposures  to  1100  to  1666 (jg/m  (0.56 to 0.85 ppm)
of Oo  was  a  hyperreactive airway response.   This  abnormal  breathing  pattern
was elicited by  mild exercise, histamine aerosol  inhalation, or breathing air
with reduced oxygen (0^) or elevated carbon dioxide (C0?) concentrations.   The
rapid, shallow breathing observed in dogs following 0., 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
following 03 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  0, exposure  showed  damage  to the
respiratory epithelium  (Section  10.3.1).   Damage to the epithelium overlying
sensory  receptors  may  be  responsible  for the  increased  receptor reactivity  to
mechanical  stimulation  (increased ventilation with exercise, low 0?,  or high
C02)  or  chemical  (histamine)  stimulation (Nadel,  1977;  Boushey et  al.,  1980).
The rapid, shallow breathing  observed  in experimental animals during  0., expo-
sures  (Amdur et  al.  1978; Yokoyama, 1969;  Murphy  et  al., 1964)  may also be
related  to increased sensory  neural activity  coming from  the lungs, not to  an
indirect effect  of  changes  in airway diameter or lung distensibility as pre-
viously speculated.
0190Z4/A                           10-63                                5/1/84

-------
     In their  study of allergic lung  sensitization,  Osebold et al.  (1980)
showed additional functional evidence  for  epithelial  disruption caused by 0,
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 ug/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  0~  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
aerosolized ovalbumin, 34 percent of the 0--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  ug/m   (0.5 ppm)  of 0~ 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 03 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
(1970)  observed  a similar increase  in  the  allergic response of sensitized
guinea  pigs  to inhaled  antigen  following  30 min of  exposure  to 2 ppm 0.,.

10.3.3  Biochemically Detected  Effects
10.3.3.1  Introduction.   This section will   include some studies involving con-
centrations above  1960 ug/m   (1 ppm)  of 03, 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-  are  similar across  the  classes  of  metabolism.
Lung permeability is discussed in this section, because it  is typically detected
0190Z4/A                           10-64                                5/1/84

-------
biochemically.   The final subsection presents hypotheses about the molecular

mechanism(s) of action of 0_,  relying on  the data presented earlier by metabolic

class.
10.3.3.2  Antioxidant Metabolism.   Antioxidant metabolism of  the lungs  is

influenced by 0_ exposure.  As  shown  in  the schematic diagram below  (Figure
               O
10-3),  this system  consists  of a number of  enzymes.  As shall be discussed, 03
has been  shown to produce several reactive oxidant species  iji  vitro from  com-

pounds  found in the lung, as well  as  in other  organs.   It is reasonably certain

that 0.( can produce such reactive species in the lung  after  jj\ 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 0_.
                     -G.6.p	v—NADP + -*-v^ GSH
                          G-6PD        GSH       GSH
                         HMP shunt     reductase   peroxidase
-A—GSSG**-A*-»» I
                     6_PG .*_/^-»*NADPH-/V— GSSG *•*-**-»» ROH
          Figure 10-3. 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 =  glutathione; GSSG  -
          glutathione disulfide; [O] = oxidizing moiety [i.e., hydrogen
          peroxide, free radical, lipid peroxide]; GSH peroxidase =
          glutathione peroxidase; GSH reductase  = glutathione reductase;
          and ROH = reduced form of [O]).
          Source: U.S. Environmental Protection Agency (1978).
                                                                   3
     Typically, following exposures to levels of 0_ below 1960 pg/m  (1 ppm),
                                                  «3
the  activities  of  most enzymes in this system  are  increased  (Table 10-5).

Whether this  increase  is  due  to direct mechanisms (e.g., de  novo  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



0190Z4/A                           10-65                                 5/1/84

-------
                                    TABLE  10-5.  CHANGES IN THE LUNG ANTIOXIDANT METABOLISM AND OXYGEN CONSUMPTION  BY  OZONE
o
 i
en
cr>
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
. 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
1 8 hr/day for
7 days
Observed effect(s)c Species Reference
With vit E-deficient diet, increased levels of GSH and Rat Chow et al., 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-PO 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 succinate-cytochrome C
reductase activity fairly proportional to 03 level.
Similar results for intermittent exposure groups.
         1568    0.8
Continuous       Increased rates of 02 consumption, reaching a peak
for 1 to 30      at day 4 and remaining at a plateau for the re-
days             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  10-5.  CHANGES IN THE LUNG ANTIOXIDANT METABOLISM AND OXYGEN CONSUMPTION BY OZONE  (continued)
o
 i
Ozone
concentration Measurement3'
ug/m3
392
686
980
1568



392
980
1568




392
OOA
980
1568


392
980
-| f\ff\
1960

392
OOA
3OU
1960

2352-
1 c f\T)
-LO , U / L.


ppm method
0
0
0
0



0.
0.
0.




0.
0.
0.


0.
0.
1.

0.
.
1.

1.
.


2 I
35
5
8
i
i

2 NBKI
5
8




2 ! MAST,
5 NBKI
8


2 NBKI
5
0

2 NO

0

2-



k Exposure
duration and
protocol
8 hr/day for
7 days





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



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


3 hr/day for
4 days


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 ppm.
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 Mouse
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 in 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 10-5.   CHANGES  IN  THE  LUNG ANTIOXIDANT METABOLISM AND OXYGEN CONSUMPTION BY OZONE  (continued)
O
 I
Co
Ozone
concentration
ug/m3 ppm
392 0.2
980 0.5
1568 0.8
1568 0.8
3920 2
7840 4
392 0.2
980 0.5
1568 0.8
627 0.32
882 0.45
. 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) Species
All 03 levels: increase in NPSH levels; increased Rat
activities of G-6-PO, 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 Monkey
enzymes .
Loss of GSH; loss of SH from lung Rat
raitochrondrial and nricrosomal frac-
tions and inhibition of marker
enzyme activities from these
fractions.
Concentrati on- related increase in 02 consumption. . Rat
In both vit E-supplemented and nonsupplemented Mouse
groups: increased G-6-PO activities and GSH
levels; decreased ACHase activities.
Mice: increased levels/activities of TSH, NPSH, Mouse,
GSH peroxidase, GSH reductase, G-6-PD, 6-P-GD, 3 strains
Reference
DeLucia et al . ,
1975a
j
j
Mustafa et al . , i
1973
Moore et al. , 1980
Mustafa et al . ,
1982
          882
0.45
                         UV
                                           8  hr/day
                                           for 7 days
isocitrate dehydrogenase, cytochrome c oxidase,
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.

03 and 4.8 ppm of N02 alone produced no significant
effects but 03 + N02 produced synergistic effects:
increased total and nonprotein sulfhydryls; increased
activities of succinate oxidase and cytochrome c
oxidase.
                                                                                                                      of rats
Mouse     Mustafa et  al. ,
          1984

-------
                             TABLE 10-5.  CHANGES IN THE LUNG ANTIOXIDANT METABOLISM AND OXYGEN CONSUMPTION BY OZONE (continued)
en
10
Ozone
concentration
ug/m3 ppm
882 0.45
980 0.5
1372 0.7
1568 0.8
1470 0.75
1568 0.8
1568 0.8
1568 0.8
3920 2
. Exposure
Measurement ' duration and
method protocol
NO 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
Continuous
for 7 days
NBKI Continuous for
3 days
I Continuous for
10 to 20 days
8 hr
Observed effect(s)c Species; Reference
Increased SOD activity at days 3 and 5, but not Rat
days 2 and 7 of exposure.
Bhatnagar et al. ,
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 monkeys.
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 vitamin E
levels in diet.
Increase in activities of GSH peroxidase, GSH Rat
reductase, G-6-PD, 6-P-GD, and pyruvate kinase
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
and G-6-PD; levels of NPSH; general protein synthesis;
and rate of mitochondrial succinate oxidation.
Decrease to control values 6 to 9 days postexposure.
Re-exposure using same regimen (6, 13, or 27 days
postexposure) resulted in similar elevations.
At the higher concentration: increase in the lung Rat
mitochondrial 03 consumption in oxidation of 2-
oxyglutarate and glycerol-1-phosphate and the number
of type 2 alveolar cells which are rich in mito-
chondria. No change in malonaldehyde. At the
lower concentration: increase in 02 consumption.
Chow and Tappel ,
1973
i
Chow et al., 1976b;
Mustafa et al . ,
1973

-------
                          TABLE 10-5.   CHANGES  IN  THE  LUNG ANTIOXIDANT METABOLISM AND OXYGEN CONSUMPTION BY OZONE (continued)
Ozone
concentration
ug/m3 ppm
1568 0.8
2940- 1.5-
7840 4
1568 0.8
. Exposure
Measurement ' duration and
method protocol
I < 24 hr
< 8 hr
I 10 days
Observed effect(s)c
NPSH level unaffected at 0.8 ppm of 0
or 1.5 ppm for < 8 hr; decreased at 2
8 hr or 4 ppm for 6 hr. At 4 ppm of
decreased level of GSH; no change in
Species
3 for < 24 hr or Rat
ppm of 03 for
03 for 6 hr,
GSSG level.
Lung SH levels unchanged. Increase in G-6-PO and Rat
and cytochrome c reductase activities. No change
in malonaldehyde levels.
Reference
DeLucia et al . ,
1975b
OeLucia et al. ,
1972
        3920
o
 I
                                          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.
3920- 2- 30 min In vitro: decrease in SH levels; increase in
5880 3 malonaldehyde levels.
1568 0.8 NO Continuous Increased activity of superoxide dismutase.
for 7 days
1568 0.8 UV 92 hr 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
                                                          High mortality in 7- and 12-day-old rats.
        1568
               0.8
                       UV
                                          Continuous       Diet was constant vitamin E and deficient or sup-
                                          for  5 days       plemented (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  10-5.   CHANGES  IN THE  LUNG ANTIOXIDANT  METABOLISM  AND  OXYGEN  CONSUMPTION  BY  OZONE  (continued)
Ozone
concentration
ug/m3 ppm
1568 0.8



. Exposure
Measurement ' duration and
method protocol


Observed effect(s)
UV Continuous Diet was constant vitamin E and 0 ppm of Se or 1 ppro
for 5 days in
in
GSH
Se. +Se: increased G-6-PD, 6-P-GD; no change
GSH reductase or GSH peroxidase. -Se: decreased
reductase.


Species
Mouse





Reference
Elsayed et al . ,
1983


Both diet groups had increase in TSH and NPSH, and
lung Se levels after 03.
1568 0.8



1764 0.9




1764 0.9
NBKI continuous Vitamin E partially prevented increased activities
for 7 days of
of
Ho
G-6-PD, 6-P-GD, and malic enzyme. Activities
phosphofructokinase and pyruvate kinase increased.
effect on aldolase and malate dehydrogenase.
CHEM 96 hr Trend towards decreased activities of GSH reductase
GSH
by
peroxidase, G-6-PO before 18 days of age, followed
increases thereafter. For G-6-PD: no change at 5
Rat



Rat
(5-180
days old)
Chow and Tappel ,
1973


Tyson et al . ,
1982

and 10 days of age; decrease at 15 days, and Increase
at
MAST > 96 hr 96
25 and 35 days.
hr: No effect below 20 days of age; G-6-PD in-
creases thereafter up to 35 days, after which (40




anc
at
mal
age
50 days old) it decreases. When exposure started
25 or 32 (but not 10 to 15) days of age, the maxi-
increase in G-6-PD occurred at about 32 days of
under continuous exposure conditions.

Rat
(10-50
days old)




Lunan et al . ,
1977




Measurement method:   MAST = Kl-coulometric (Mast meter);  NBKI = neutral  buffered potassium iodide; CHEM = gas solid chemiluminescence;
                      UV = UV photometry;, I = iodometric;  ND = 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.

-------
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 10-5).   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 CL 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  0, levels (3920  and 7840 ug/m ,  2 and 4  ppm) and
short-term (8 or  24 hr/day,  7  days)  exposures to  lower 0~  levels (392, 980,
         3
1568 ug/m ,  0.2,  0.5,  0.8 ppm) on rats.   For  nonprotein sulfhydryl levels
(principally glutathione, GSH), decreases in the level of GSH were progressive
                                                3
with time of  exposure  (2 to 6 hr) to  7840 ug/m  (4  ppm).  For glutathione
disulfide (GSSG), 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
                                      3
7-day continuous exposure to 1568  ug/m  (0.8 ppm) of 0_,  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
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 10-5) 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 ug/m  (0.2, 0.5, or 0.8 ppm) of 03 and found a

0190Z4/A                           10-72                                5/1/84

-------
concentrated-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  ug/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 10.3.1).  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
                       3
392, 980,  or 1960 ug/m  (0.2,  0.5,  or 1.0 ppm) of 0-.   Exercise enhanced the
effect.   At the  lower  03 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 ug/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 had increased by day 3 and reached a peak at 10 days and
then began to return toward control  values.  GSH peroxidase activity continued
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
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
metabolism.  Rats were exposed continuously for 3 days to 1568 ug/m  (0.8 ppm)
of 0~, allowed to remain  unexposed for 6,  13,  or 27  days,  and  then re-exposed
for  3  days to the same  0-  level  (Chow et al., 1976b).   Immediately after

0190Z4/A                           10-73                                5/1/84

-------
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 0,  toxicity  has  been
extensively studied, because it typically reduces the toxicity of 0~ 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  ug/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
                                                 3
(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 0.,  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
ug/m  (0.8 ppm)  of  07 continuously for 7 days also had changes in 6-phospho-
                                                                            3
gluconate dehydrogenase activity.  Rats  were exposed to 1372  to 31,360 ug/m
(0.7 to 16 ppm) of Qo 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
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
ug/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
0190Z4/A                           10-74                                5/1/84

-------
fed high- or low-PUFA diets. High supplemental levels (105 mg/kg) of vitamin E
acetate were protective  and delayed the LT50 to 0- by an average of 15 days.
Although these experiments demonstrate clearly the protective effect of vitamin
E  against  (L  toxicity,  they do  not  support  the  hypothesis that changes  in
fatty  acid  composition  of the lung  will  increase  (L  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
10.3.1) 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
days to 196 or 392 |jg/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
of the responses in the higher  vitamin E  group at 392 pg/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 ug/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:
0, 11, or  110  ppm (Chow  et al.,  1981).  Rats were exposed to 196 M9/m3  (0.1
ppm) of 0~ continuously for 7 days.   In the 0-ppm vitamin E group,  0~ 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.
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 10.3.1),  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.
0190Z4/A                           10-75                                5/1/84

-------
                                                                  3
     Chow and Tappel (1972) exposed rats continuously to 1372 ng/m  (0.7 ppm)
of 0- for  5  days.   The animals had been maintained  on  diets  with different
levels of  dl-a-tocopherol  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. , 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 a-tocopherol) from  the effect of 0~ (1568 ug/m  ,  0.8 ppm; 7
                                                    O
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
ug/m  (0.8 ppm)  of  0,.  continuously for 5 days.  In these mice, Se deficiency
                     •3
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 P3  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
0190Z4/A                           10-76                                5/1/84

-------
influence the effects of 0~ 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  sparring action
between Se (i.e., GSH peroxidase activity) and vitamin E might explain similar
effects of 0- 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  (jg/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 ug/m  (0.5 ppm)  of 0~ 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
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
                                                                      3
after exposures  for  8 hr/day for 7 days  to  levels as low as 392 ug/m  (0.2

0190Z4/A                           10-77                                5/1/84

-------
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
Long Evans) were compared after a 5-day continuous exposure to 882 ug/m  (0.45
ppm) of 0  (Mustafa et al., 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 the
increased activities 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  0_,  an innate difference  in species  sensitivity,  or differences in
experimental  design  (e.g., small  sample sizes, insufficient concentration-
response studies).
     Age-dependent  responsiveness to 0--induced  changes  in  GSH systems  has
been  observed.  Tyson  et al.  (1982) exposed rats (5 to 180 days old) to 1764
(jg/m   (0.9  ppm) of  0_  continuously for 96 hr, except for  suckling neonates  (5
                     O
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
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
                                          •3
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-

0190Z4/A                           10-78                                5/1/84

-------
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.
     Ten- to 50-day-old rats were exposed continuously to 1764 ug/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.
                                                                      3
     Elsayed et al.  (1982a)  exposed rats of various ages to 1568  ug/m  (0.8
ppm) of 0~  continuously  for 92 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-5-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
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 0^.  The
decreased activities observed in the neonates are reminiscent of the decreased
0190Z4/A                           10-79                                5/1/84

-------
activities that occur  at  higher 0~ 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 10-5).
     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.
                                                         3
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  Q*  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 (0~-),  a toxic  oxidant species thought to be  formed  from
07 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 mitochondrial fractions of the lungs (Mustafa
et al. ,  1977).  In a more complex  exposure regimen  in which  rats were exposed
                        3
for 3  days  to  1568 ug/m  (0.8  ppm)  and then  various days of combinations of
2940 ug/m  (1.5 ppm) and 5880 ug/m  (3 ppm) of 0~, SOD activity also increased.
Bhatnager et al. (1983) studied the time course of the increase in SOD activity
                                              3
after  continuous exposure of rats  to 882 ug/m  (0.45 ppm) of 0.,.   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.
10.3.3.3.   Oxidative and  Energy Metabolism.   Mitochondrial  enzyme activities
are typically studied  to evaluate effects on Op consumption, which is a funda-
mental parameter of cellular metabolism.  Mitochondria are cellular organelles
that are the major sites of 0?  utilization and energy production.  Many of the
enzymes  in  mitochondria  have functional sulfhydryl  groups, which are known to
be affected  by  0~,  and mitochondrial membranes  have  unsaturated fatty acids
that are also susceptible to  0-.  The patterns of 0, effects on 0? consumption,
as will  be  discussed,  are quite similar to effects on antioxidant metabolism
(Section 10.3.3.2).
0190Z4/A                           10-80                                5/1/84

-------
     Mustafa et al.  (1973) showed in rats that acute (8-hr) exposure to a high
                               3
concentration of 0-  (3920 pg/rn ,  2 ppm) decreases  (L  consumption using the
substrates  succinate,  croxoglutarate,  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 mitochondrial  fraction.   No
change in malonaldehyde  levels  was found.   When rats were exposed to high CL
                 3
levels (5880 ug/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 ug/m , 0.8 ppm)
caused an increase  in  0^ 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 activity 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 0? 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  ug/m3  (0.2, 0.5, or 0.8 ppm) of 03.
Succinate oxidase activity increased linearly with 0., concentration.  No major
differences  were apparent between continuous  and intermittent exposures.   No
statistical   analyses  were  reported.   Concentration-dependent morphological
effects were also observed (Section 10.3.1).  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 0? consumption of rats increased with in-
creasing 03  level (392, 980, 1568 |jg/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
            3
to 1568 ug/m  (0.8 ppm) of 03 continuously for 30 days, and 02 consumption was
measured as  the activities  of succinate oxidase, 2-oxoglutarate oxidase,  and
glycerol-1-phosphate oxidase.   On day  1,  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
0190Z4/A                           10-81                                5/1/84

-------
30 days of exposure.   Mitochondria!  succi nate-cytochrome c reductase exhibited
a similar pattern under similar CL levels for 7 days of exposure.  Equivalent
                                                                          3
results occurred  in  rats  during a 7-day  continuous  exposure to 1568
(0.8 ppm) of 0, (DeLucia et al . ,  1975a).
                                                        3
     In rats,  recovery  from  an  ozone-induced (1568 yg/m  , 0.8  ppm;  3 days,
continuous) increase in  succi nate  oxidase  activity occurred by 6 days post-
exposure (Chow et  al . ,  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 09 consumption.
                                                       3
After 7  days  of  continuous exposure to 196 or 392 pg/rn   (0.1 or  0.2  ppm) of
0~, the  lung  homogenates  of  rats maintained on diets with either  11 or 66 ppm
 O
of vitamin  E  were  examined for changes in 0^ consumption (succinate  oxidase
activity)  (Mustafa,  1975;  Mustafa  and Lee,  1976).  In the  11-ppm vitamin E
group,  increases in  09  consumption  occurred at both 0-. levels.  In the 66-ppm
                              3
vitamin E group, only 392 [jg/m  (0.2 ppm) of 0, caused an increase.   Mitochon-
dria  were  isolated from the lungs  and  studied.   Neither dietary group  had
03-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 mitochondrial 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 10.3.3.2),
responsiveness to effects of 0Q on 09 consumption is age-related  (Table 10-5).
                                                                3
Elsayed et al . (1982a) exposed rats of various ages to 1568 pg/m  (0.8 ppm) of
0-  continuously  for 92  hr.   The 0,- induced increase  in  the activities of
 O                                 O
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.
     Species  and strain differences were observed after  a  5-day continuous
exposure  to 882 |jg/m3  (0.45 ppm)  of 03 (Mustafa  et  al . , 1982).   Mice and
Sprague  Dawley rats  (but  net 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.
0190Z4/A                            10-82                                 5/1/84

-------
     Rats  have  also been compared to two strains of monkeys after a 7-day (8
hr/day)  exposure to various  concentrations  of 03 (Mustafa and Lee,  1976).
Rats were  exposed to 392,  980,  or 1568  ug/m3 (0.2,  0.5,  or 0.8  ppm)  of  DS  and
exhibited  increases  in succinate oxidase activity.   Rhesus monkeys exposed to
                       3
either 980 or 1568 pg/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 0.,, succinate
oxidase activity  increased  at the two higher 0- levels.  The number of animals
                                                                              o
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).
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 0^ consumption  is typically attributed to concurrent
morphological changes  (Section  10.3.1)  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 al., 1981; Schwartz et al., 1976; DeLucia
et al. ,  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 pg/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
the number of  mitochondria, rather than an increased activity within a given
mitochondrion,  in exposed lungs is the probable dominant cause.
10.3.3.4  Monooxygenases.   Multiple microsomal enzymes function in the metab-
olism of  both  endogenous  (e.g.,  biogenic amines,  hormones)  and  exogenous
(xenobiotic)  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  0- exposure (Table 10-6).
0190Z4/A                           10-83                                5/1/84

-------
TABLE 10-6.  MONOOXYGENASES
Ozone
concentration Measurement '










o
i
oo
-pa
















MQ/rn3
392
980
1568


1470
5880
19,600
1470
5880
19,600

1568


3920
1568


3920
392
1568
1568
1568

1568


ppm method
0.2 MAST
0.5 NBKI
0.8


0.75 I
3.0
10.0
0.75 I
3.0
10.0

0.8 NO


2.0
0.8 I


2.0
0.2 I
0.5
0.8
0.8

0.8 I


. Exposure
duration and
protocol
Continuous or
8 hr/day for 7
days


3 hr


3 hr



Continuous for
7 days

8 hr
Continuous
for 7 days

8 hr
Continuous or
8 hr/day

8 hr/day

Continuous for
7 days


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

Decreased activity of banzpyrene Rabbit
hydroxylase in tracheobronchial
mucosae.

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 10-6.   MONOOXYGENASES (continued)
Ozone
concentration
i — *
0
OD
ug/m3
1960
1960
ppm
1.0
1.0
b Exposure
Measurement ' duration and
method protocol Observed effect(s)
ND 90 min Decreased levels of lung cyto-
chrome P-450. Maximal decrease
at 3.6 days postexposure.
MAST 24 hr 50% decrease in benzphetamine
N-demethylase activity 1 day
Species
Rabbit
Rat
Reference
Goldstein et al . ,
1975
Montgomery and
Niewoehner, 1979
  5880    3
                          NBKI
10 min before
lung perfusion
and continuous
throughout
experiment.
postexposure; return to control
levels by 1 wk postexposure.
Stimulation of cytochrome
bs-mediated lipid desaturation.

Decreased enzymatic conversion of
arachidonic acid to prostaglandins
when using isolated ventilated per-
fused lung.
                                                                                                           Rat
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.

-------
     Monoamine oxidase (MAO)  activity  has  been investigated (Mustafa et al.,
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 |jg/m  (0.8 ppm) continuously for 7 days.
Substrates used included rramylamine, benzyl amine, 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
                                                          3
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 mitochondrial enzymes and antioxidant metabolism  (Sections  10.3.3.2
and 10.3.3.3).
     The cytochrome P-450-dependent enzymes have been studied because of their
function in drug and carcinogen metabolism.  Palmer et al.  (1971, 1972)  found
that hamsters  exposed to 1,470 ug/m  (0.75 ppm)  of 0., for 3  hr had lower
benzo(a)pyrene hydroxylase  activity.    Goldstein  et  al.  (1975) showed  that
                              3
rabbits exposed to  1,960 |jg/m  (1  ppm) of 0- 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-demethylase in
                                                      3
the  lung was  lowered by a  24-hr exposure to 1,960 ug/m  (1  ppm)  of 03 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  1 week.   Cytochrome bg-mediated lipid desaturation was
stimulated by  03 4, 7, and  14 days postexposure.  Immediately  after  exposure,
the desaturase activity was quite depressed, but this was attributed  to anorexia
in  the  rats,   and  not to 0^.  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  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 03 exposure are involved in the activa-
tion, which makes  full interpretation  of the effect of 0- on this metabolism

0190Z4/A                           10-86                                 5/1/84

-------
impossible.   The impact  of  the decrease in cytochrome  P-450  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 10.3.3.2; 10.3.3.3),
                                             3
acute exposure to  a  high 0_ level (3920 ug/m  ,  2 ppm;  8 hr)  reduced NADPH
                           O
cytochrome c reductase  activity  (DeLucia et al., 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. ,  1977).   Continuous and intermittent  exposures
were not  different.   The time  course of the response to 1568 ug/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
                                             3
exposure for 8 hr/day for 7 days to 1568 |jg/m  (0.8 ppm) of 0, (DeLucia et al. ,
1975a).  However, monkey data were not reported in any detail.
10.3.3.5  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 on Table  10-7.  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
10.3.1), such an increase in type 2 cells and alveolar macrophages and polymor-
phonuclear 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 10.4.3.
     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
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 07 continuously for
                                3
7 days.   Lower levels  (196  ug/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
0190Z4/A                           10-87                                5/1/84

-------
                                          TABLE 10-7.   LACTATE DEHYDROGENASE AND LYSOSOMAL ENZYMES
Ozone
concentration
|jg/m3
196



' 392
980
1568

980
1568
i—"
o
i
Co 1568
CO
1372

1568

1372-
1568
i
i 1568


1568


ppm
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


a h
Measurement '
method
NBKI



MAST,
NBKI


NBKI






NBKI



MAST,
NBKI

MAST,
NBKI

MAST,
NBKI

Exposure
duration and
protocol
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 in Rat
diet group receiving 0 ppm
vitamin E. Groups with 11 or
110 ppm of vitamin E had no effect.
Increased lung lysozyme activity only Rat
after continuous exposure to 0.8 ppm.


Increased LOH activity in lungs. Rat
Change in LDH isoenzyme distri-
bution at 0.8 ppm.


No change in total LDH activity or Monkey
isoenzyme 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 HH activity not Rat
affected by vitamin E (0 or
45 mg/kg diet).


Reference
Chow et al. ,



Chow et al . ,



Chow et al . ,






Oil lard et al
1972


Castleman et
1973a

Castleman et
1973b




1981



1974



1977






.,



al.,


al.,


Chow and Tappel ,
1973



Measurement method:   MAST = Kl-coulometric (Mast meter); NBKI = neutral  buffered potassium iodide.
bCalibration method:   NBKI = neutral buffered potassium iodide.

-------
LDH activity and isoenzyme pattern (relative ratios of different LDH isoenzymes)
                                                                            3
in the lungs, plasma,  and erythrocytes of 0--exposed rats (980 or 1,568 (jg/tn ,
                                      3
0.5 or 0.8 ppm) and monkeys (1568 ug/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
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
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  03-  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 03  at  concentrations of 1,372 ± 294 ug/m   (0.70 ± 0.15 ppm) for 5
days and 1,548 ± 274 pg/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 03 exposure.   Similarly,  Castleman et al.  (1973a,b) found that activity
of lung acid phosphatase  was  increased in young  rats  that had been  exposed to
1,372 to 1,568 ug/m  (0.7 to 0.8 ppm) of 0- continuously for 7 days.  Increases
in p-glucuronidase activity were not observed.  The histochemical and cytochem-
ical localization suggested  that  0.,  exposure results  in damage to the lung's
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
0190Z4/A                           10-89                                5/1/84

-------
represent an enzymatic protective mechanism  operating locally in the centri-
acinar regions  of  0~-exposed lungs.   Chow  et al.  (1974) also  observed  an
increase in lysozyme activity (lung  homogenate)  in rats  exposed continuously
            3
to 1568 ug/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 ug/m  (0.2 or 0.5 ppm) of 03.
     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.
10.3.3.6  Protein Synthesis.  The effects of 03  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 10-8.
     Hesterberg and Last (1981)  found that increased collagen synthesis caused
                                                          3
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 03 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  OT for 30  days, the  augmentation of activity seen  earlier at
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
            o                              J
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
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
0190Z4/A                           10-90                                5/1/84

-------
                                                TABLE  10-8.  EFFECTS OF OZONE ON  LUNG  PROTEIN  SYNTHESIS
Ozone .
concentration Measurement '
ug/mj
392
784
1176
ppm method
0.2 NO
0.4
0.6
Exposure
duration and
protocol
8 hr/day for
3 days

Observed
Decreased rate of
tracheal explants

effect(s)
glycoprotein secretion by
at 0.6 ppm.

Species Reference
Rat Last and Kaizu,
1980; Last and
Cross, 1978
        1568
         980
o
i
0.8
0.5
Continuous for
1 through 90 days

Continuous for
3 or 14 days and
combined with
H,S04
Decreased rate of glycoprotein secretion.
Increased rate of glycoprotein secretion.
392
980
1568




392
784
1176
1568

392
1568
3920
980-
3920

0.2 MAST
0.5
0.8




0.2 UV
0.4
0.6
0.8

0.2 UV,
0.8 NBKI
2.0
0.5- UV
2

Continuous for
7 days





8 hr/day for
1 to 90 days



6 hr/day, 5 days/
wk, 12.4 wk (62
days of exposure)
1, 2, or 3 wk


Concentration-dependent increase in lung prolyl Rat
hydroxylase activity. No effect at 0.2 ppm. Meta-
bolic adaptation suggested at 980 MS/™3 (0-5 PPm)
At 0.8 ppm, collagen and noncollagenous protein
synthesis increased; effect on proyl 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 rate of collagen synthesis; fibrosis Rat
of alveolar duct walls; linear concencentration
response.
Hussain et al . ,
1976a,b





Last et al. , 1977




Costa et al. , 1983


Last et al. , 1979



-------
                                         TABLE 10-8.   EFFECTS OF OZONE ON LUNG PROTEIN SYNTHESIS  (continued)
o
 i
Ozone . Exposure
concentration Measurement ' duration and
ug/m3
882
1568
980
1568
1568
1568
2352
2940
ppm method
0.45 ND
0.8 ND
0.5 UV
0.8 NO
0.8 NBKI
0.8 UV
1.2
1.5
protocol
Continuous
for 7 days
Continuous
for 90 days
Continuous for
up to 180 days
Continuous
for 7 days
Continuous
for 3 days
Continuous
for 7 days
Observed effect(s)
Increased collagen synthesis at 5 and 7, but
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
7 days. No effect 20 days and beyond.
Increase in protein and hydroxyprol ine content of
lungs. No change 2 mo postexposure.
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
Mouse Bhatnagar et al . ,
1983
Rat
Rat Last and Greenberg,
1980
Rat Mustafa et al . ,
1977
Rat Chow et al. , 1976b
Rat Hesterberg and Last,
1981
      Measurement method:  MAST = Kl-coulometric (Mast meter); UV = UV photometry; NBKI = neutral buffered potassium iodide; ND = not described.


      Calibration method:  NBKI = neutral buffered potassium iodide.

-------
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
exposure to ozone for I, 2, or 3 weeks at levels ranging from 980 to 3920 ug/m
(0.5 to  2  ppm).   A general correlation was  found  between  fibrosis detected
histologically and the quantitative changes in collagen synthesis in minces of
0«-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  0,  concentration to which the rats were
exposed.
     Continuous exposure of mice  for 7 days  to  882  jjg/m   (0.45 ppm)  of 0.,
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  2, 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.
     Bhatnagar et  al.  (1983)  also studied rats  exposed continuously for 90
days to  1568 ng/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.
     Effects of subchronic  0- exposure have been studied by Last and Greenberg
                                       3
(1980).   Rats were  exposed to  980 ug/m  (0.5 ppm) continuously for up to 180
days and examined at various times during exposure and 2 months after exposure
ceased.   The total  protein content of the lungs  increased  during exposure,
with the greatest increase  after 88 days of exposure.   By 53 days postexposure,
values had  returned  to  control  levels.  Hydroxyproline content of  the  lungs
0190Z4/A                           10-93                                5/1/84

-------
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 hydroxypro-
line 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 discussing their data, the authors referred
to a concurrent morphological study (Moore and Schwartz, 1981) that  showed an
increase in lung volume, mild thickening of the interalveolar septa and alveo-
lar interstitium,  and an increase in collagen (histochemistry)  in  these areas.
Different results  for collagen  levels were observed  after another  longer term
exposure (6 hr/day, 5 days/wk, 12.4 wk) to 392, 1568, or 3920 pg/m  (0.2, 0.8,
or 2.0 ppm) of 0.,  (Costa et al., 1983).   At the two lower concentrations, rats
exhibited an equivalent decrease in hydroxyproline.   At the highest concentra-
tion, an increase  was observed.   Similar findings were made for elastin levels.
     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 0,
(Chow  at  al. ,  1976b).   Recovery  had occurred by 6  days  postexposure.  No
adaptation was observed,  because  when animals were re-exposed to  the same 0«
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 in vivo  incorporation of radioactive ami no acids.   Rats  were exposed
                                    3
continuously for 7 days to 1568 pg/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  mucus glycoproteins and their secretion  by tracheal
explants have been reviewed by Last and  Kaizu (1980).  Mucus glycoprotein syn-
                                                                    3
thesis  and secretion were  measured by  the rate of incorporation of  H-glucos-
amine into mucus glycoproteins and their subsequent secretion  into supernatant
fluid of  the  tracheal  exp'lant culture medium.  This method has been found to
be a  reproducible  index of mucus production, and these authors maintain that
0190Z4/A                           10-94                                5/1/84

-------
this measurement  ex vivo following  exposure  i_n vivo is  representative  of
injuries occurring Jjn vivo.   When rats were exposed to 1568 pg/m  (0.8 ppm) of
CL 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,
            3
and 392 ug/m   (0.6,  0.4,  0.2 ppm) of  0~.   Glycoprotein  secretion decreased
only at  the higher  concentration.   Trachea!  explants from Bonnet monkeys
                               3
exposed to 0, 980, or 1568 ug/m  (0, 0.5 or 0.8 ppm) 03 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.
                                             3
     A combination  exposure to  0,  (980 ug/m ;  0.5 ppm)  and  sulfuric acid
                         3
(HLSO.) aerosol  (1.1  mg/m  )  caused  complex effects  on mucus  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 mucus  secretion  rates, but H^SO.  had  no
effect.   In  rats  exposed to the combination  of  H^SO.  aerosol  and 03, mucus
secretion significantly  increased.   After  14  days  of  continuous exposure,  the
rats receiving a  combined  exposure to both HUSO, aerosol  and 0« had elevated
values  (132 percent) over the control group of animals.   Because mucus secretion
and synthesis are intimately involved in diminishing the exposure of underlying
cells to  0«  and  removing adventitiously inhaled  particles,  alterations in  the
mucus 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.
10.3.3.7  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 10-9.  Generally, the  unsaturated fatty
                                                   3
acid content decreased in rats exposed  to  980 ug/m  (0.5  ppm)  of  03  for  up  to
6 weeks (Roehm et al., 1972).
0190Z4/A                           10-95                                5/1/84

-------
                            TABLE  10-9.  EFFECTS OF OZONE EXPOSURE ON  LIPID METABOLISM AND CONTENT OF THE  LUNG
Ozone
concentration
jjgTni3 ppm
980
H- 980
*p 1568
<-D
01 :
1960
1960
0.5
0.5
0.8
1.0
1.0
a 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
NO 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 ID vitamin £/kg diet, decreased
ethane and pentane production.
Decreased incorporation of fatty Rabbit
acids into lecithin.
Reference
Roehm et al . , 1972
Dumel in et al . ,
1978a
Dumel in et al . ,
1978b
Kyei-Aboagye
et al. , 1973
Measurement method:  NBKI = neutral buffered potassium iodide; UV = UV photometry; I = iodometric; ND = not described.

-------
                               PRELIMINARY DRAFT
     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
1960 (jg/m   (1 ppm)  of 0~ for 60  min.  The provision  of 11 (minimum vitamin
requirement) or 40 ID (supplemented level) of vitamin E acetate per kg of diet
resulted  in a  decrease  in expired ethane and  pentane after 0, exposure.
Dumelin  et  al.   (1978a)  also measured breath ethane  and  pentane in Bonnet
                                       3
monkeys exposed to 0, 980, or 1568 ug/m  (0, 0.5, or 0.8 ppm) of 0_ for 7,  28,
                                                                  o
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.
     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  (jg/m  (1 ppm) of 0~  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  0~  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.
(1971)  and  Huber  et  al.   (1971) when  using high  levels of 0., (> 9800 |jg/m ;
5 ppm).
10.3.3.8  Lung Permeability.  Table 10-10  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. (1989) studied  the  effect  of 0~ on  active transport of
                                                            3
phenol  red in the lungs of rats exposed to 1176 to 4116 ug/m  (0.6 to 2.1 ppm)
of 0-  continuously for 24 hr.  Ozone  inhibited the carrier-mediated  transport
of intratracheally instilled phenol  red from the lung to the circulation and
0190Z4/A                           10-97                                5/1/84

-------
                                                     TABLE 10-10.   EFFECTS OF OZONE ON LUNG PERMEABILITY
o
 i

<-o
oo
Ozone

concentration
ug/m-3
196
510
1000
1960
353

510
980


1000

490
980
1960
4900
1176
2156
3136
4116
1960-
2940
ppm
0.1
0.26
0.51
1.0
0.18
. Exposure
Measurement ' duration and
method protocol
CHEM, 3 hr or 72 hr
UV


3 hr; 8 hr/day
! for 5 or 10 days
0.26 3, 24, or 72 hr
0.5

|
0.51

0.25
0.5
1.0
3 hr

MAST, 6 hr
NBKI

2.5
0.6 NBKI 24 hr
1.1
1.6
2.1
1.0- MAST 2 hr
1.5


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
0.5 ppm, 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:
MAST = Kl-coulometric (Mast meter); CHEM = gas phase chemiluminescence; UV = UV photometry; NBKI =  neutral buffered

potassium iodide.
          Calibration method:   NBKI = neutral buffered potassium iodide;  UV = UV photometry.

-------
                                                I
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 03 exposure,
the appearance  of  albumin and immunoglobins  in
examined.   Reasor et  al.  (1979)  found that dogs
(1.0 to 1.5  ppm)  of 0~ had increased  albumin  a
their airway secretions.  Alpert et al.  (1971a),
                                                 airway  secretions has been
                                                 breathing 1960 to 2940 ug/m3
                                                rid immunoglobin G content of
                                                using rats exposed for 6 hr to
                                                1.0  ppm)  of 03 for 72 hr and
490 to 4900  ug/m   (0.25 to 2.5 ppm) 0~, also fdund 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.  (198J2) exposed guinea pigs to 196,
510, 1000, or  1960 ug/m3  (0.1,  0.26,  0.51, or
found increased lavage fluid protein content sampled immediately after exposure
                             3
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
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 ei concentration-related manner
                     3                                     3
from 500 to  1470 ug/m   (0.256 to 0.75 ppm).  Again  196  ug/m  (0.1 ppm) had no
                                                                             3
effects.   The lavage fluid protein content of guinea pigs exposed to 353 ug/m
                                               consecutive days was  not dif-
                                                C deficiency could be found on
                                               i guinea  pigs exposed  to 196 to
                                                et al.,  1982).   In contrast,
(0.18 ppm) of  03  for 8 hr per day for 5 or 10
ferent from air controls.   No effect of vitamir
the accumulation of  the lavage fluid protein  i
1470 ug/m3 (0.10 to  0.75  ppm) 0, for  3  hr  (Hi
                                
-------
                                        3
edema in rats from exposure to 7890 pg/m  (4 ppm)  of 0~  for 4 hr (Giri  et al. ,
1975).   Prostaglandins F?   and  E?  were markedly increased in plasma and lung
lavage of rats  exposed  to  7840 ng/m   (4 Ppro)  for up to 8 hr (Giri et al. ,
1980).   Ozonolysis of arachidonic acid i_n 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
(Pryor, 1976; Pryor  et  al. , 1976).   Acute exposure to 5880 pg/m3 (3 ppm) 03
inhibits uncompetitively rat lung prostaglandin cyclooxygenase (Menzel  et al.,
1976).
     Earlier  reports  that  prostaglandin synthesis  inhibitors  exacerbated
0~-produced edema (Dixon and Mountain, 1965; Matzen, 1957), as did methyl pred-
nisolone (Alpert et al., 1971a), tend to confuse the interpretation of the role
of prostaglandin  in  0,,-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  0~-produced  edema,  their exact role is
still unexplained.
10.3.3.9  Proposed Molecular  Mechanisms  of  Effects.   Experts  generally agree
that the  toxicity of Q~ depends on  its  oxidative properties.  The precise
mechanism of  0  '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  Association
(1970); North Atlantic  Treaty Organization  (1974); National  Research  Council

0190Z4/A                           10-100                                5/1/84

-------
(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.
10.3.3.9.1  Oxidation of polyunsaturated lipids.   The first hypothesis  is  that
Og 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 peroxidation all can react
with proteins.  In  addition  to  the  direct  oxidation  of amino  acids by 0~,
secondary  reaction  products  from  0,-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
changes of the membrane-embedded protein.  At present, methods  are not availa-
ble to differentiate  effects  on  membrane proteins from effects  on membrane
lipids.
     Some  of  the  strongest evidence that  the  toxic  reaction  of  0., 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 0.,-induced increase
in enzyme  activities  of  the  glutathione peroxidase system (Section 10.3.3.2)
and those involved in oxygen consumption (Section 10.3.3.3).   More details are
provided in Table 10-5.   Morphological  effects due to  0-  exposure are also
lessened by dietary  vitamin  E (Section  10.3.1.4.1.1).  Although  this  is not
the strongest evidence,  vitamin  E supplementation also prevented 0--induced
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
0190Z4/A                           10-101                               5/1/84

-------
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 ug/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 pg
of  vitamin E/g of tissue), but morphological  responses were very  similar.
Stephens et al. concluded  that 0--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-
10.3.3.9.2  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  0~  (Mudd et al.,  1969;  Mudd and Freeman, 1977;  Previero et al.,  1964).   In
the  following  descending order of  rate, 03  oxidizes  the ami no acids  cysteine >
methionine >  tryptophan >  tyrosine > histidine > cystine  > phenylalanine.  The
remaining  common  amino 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  i_n vitro by 03
as  shown  by  studies of avidin,  the biotin-binding protein.   Oxidation of
tyrosine  in egg albumin by 03 occurs i_n vitro, converting the Deoxidized egg
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 03,  alpha-1-protease inhibitor  lost its ability to
inhibit  trypsin,  chymotrypsin, and elastase.
 0190Z4/A                           10-102                               5/1/84

-------
     Meiners et al.  (1977)  found that 03 reacted  i_n  vitro with tryptophan,
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 0-  also  generates hydrogen peroxide.   Hydrogen  peroxide  is a
toxic substance in  itself and initiates peroxidation of  lipids (McCord and
Fridovich, 1978).   Other active Q? 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
03 (viz. ,  1960  to 7840 ug/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 |jg/ro , 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,
and mitochondrial 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 glyce ides.  Mudd and Freeman contend,
however,  that proteins  are  far more easily oxidized  by  0^ than are lipids.
Indirect evidence in support of the idea that amines  in particular are oxidized
preferentially by 0- is the protective effect of p-aminobenzoic acid (Goldstein,
B.,  et  al. ,  1972).  Rats injected  with p-aminobenzoic acid were partially
protected from  the  mortality due to high concentrations  of 0.,.   Presumably,
the added p-aminobenzoic acid is oxidized by 0- in place of proteins.  Goldstein
and Balchum (1974) later suggested that the protection of p-aminobenzoic acid,
allylisopropylacetamide,  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

0190Z4/A                           10-103                               5/1/84

-------
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  in  vitro
                        3
exposure to  78,400  ug/m  (40 ppm) 03 through  p-aminobenzoic  acid  treatment
(Goldstein, B., et al., 1972).  Amines are also efficient lipid antioxidants,
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
well.
10.3.3.9.3  Formation of toxic compounds  through reaction  with polyunsaturated
             lipids.   The effects of  0-  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 composition of the  serum and lung lipids were similar  to
those occurring after Q~ exposure (Roehm  et al., 1971a,b,  1972; Menzel  et al.,
1976).
     Peroxidation of lung lipids could lead to cytotoxic products.   Phosphati-
dyl choline  liposomes  (spheres  formed by emulsifying  phosphatidyl  choline in
water) were  lysed  on  exposure to 0, (Teige et al., 1974).  Liposomes exposed
to 03 were more active  than  03  alone  in lysing red blood  cells.  The products
of ozonolysis of phosphatidyl choline could be stable, yet toxic,  intermediates.
10.3.3.9.4   Formation of free radicals and injury mediated by pharmacologic
             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  ROOR)  rather  than oxidation  of alkenes  to higher  oxidation
states (e.g., ROM,  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.

0190Z4/A                           10-104                               5/1/84

-------
     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 03 (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,
                                3
at levels  of 03  below 1960 ug/m  (1 ppm), increases in glutathione have been
observed (Plopper  et al.  ,  1979;  Fukase et al. , 1975; Moore  et  al. ,  1980;
Mustafa et al., 1982).  The rat lung is sensitive to the  increase of glutathione
peroxidase,  glutathione  reductase,  and  glucose-6-phosphate  dehydrogenase
                                        3
activities at  levels  as low as 196 ug/m  (0.1 ppm) 03  continuously for 7 days
(Plopper et  al.,  1979;  Chow et al.,  1981;  Mustafa,  1975;  Mustafa and Lee,
1976) in vitamin E deficient rats.  With vitamin E-supplemented rats, 392 pg/m
                                                                      3
(0.2 ppm) caused similar effects.   After acute exposures  to >1960 ug/m  (1 ppm)
0-, decreases  are  observed  in  these enzyme  activities  and  glutathione levels.
Other species (mice and monkeys) exhibit similar effects  at different concentra-
tions  (Table  10-5).  These  changes at the lower levels of 03  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 al., 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
lung.   This  was  substantiated by Mustafa  (1975)  and  DeLucia  et  al.  (1975a),
who found  that the  0?  consumption per mitochondrion  was  not increased in
0,,-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 0,,-exposed rats  were due to the demands of the glutathione peroxidase
         O                                    i
system for reducing equivalents in the  form of NADPH.   Cross et al. (1976) and
DeLucia  (1975a,b)  reported  that 0- 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 03 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
0190Z4/A                           10-105                               5/1/84

-------
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 0Q in vitro.
                                                                  *5 —— _—__—_
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 0, and nitrogen dioxide.  Fairchild and
Graham ascribed  the mortality  following 0~ to pulmonary edema.   Wong and
Hochstein (1981) suggested that 0, 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 0, sensitivity.
10.3.3.9.5  Summary.   The  actual  toxic mechanism of 0- may involve a mixture
of all of these chemical mechanisms because of the interrelationships between
the  peroxide detoxification  mechanisms  and glutathione  (See Section 10.3.3.2)
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

0190Z4/A                           10-106                               5/1/84

-------
free radical reactions (vitamins E and C), and differences in membrane composi-
tion of PUFA (relative ease of attack of 03), for example.

10.3.4  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 0- 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 CL 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 CL on its integrity (Gardner,
1981; Ehrlich, 1980;  Gardner and Ehrlich, 1983).
     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
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  * Sections.  The immune system,
which defends the overall  host  against both infectious and neoplastic diseases,
is also discussed.
10.3.4.1  Mucociliary  Clearance.   The mucociliary  system  is  one of the lung's
primary defense  mechanisms.   It protects the  conducting  airways  by trapping
and quickly  removing material that has been deposited or  is being cleared from
the alveolar region.   The effectiveness of mucociliary clearance can be deter-
mined  iDy  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  distribution  of mucus-secreting
cells.  Once this defense mechanism has been altered, a buildup of both viable

0190Z4/A                           10-107                               5/1/84

-------
and nonviable inhaled substances can occur on the epithelium and may jeopardize
the health of  the  hose,  depending on the nature of the uncleared substance.
     A number  of studies  v/ith  various animal species have  reported morpho-
logical damage to  the  cells  of the tracheobronchial  tree  from acute and sub-
chronic exposure to  490  to  1960 ug/m  (0.20 to 1.0 ppm) of 0 .   (See Section
                                                             O
10.3.1.)  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  tracheal 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  al.,  1980).   In  the
same study, when the animals were  subsequently exposed  for  2  hr to  1090 ug of
H?SO. per cubic meter (0.30 urn volume median diameter),  a significant  reduction
in ciliary  beating frequency occurred.    The  magnitude  of this  effect was,
however, significantly  less  than that observed  due  to  the effect  of  H^SO^
exposure 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
                                                                             3
slight decrease in beating frequency with a simultaneous exposure to 196 |jg/m
(0.1 ppm) of 03 and 847 ug/m  H2S04 (Grose et al., 1982).   These data  indicate
that 03  appears  to partially protect against the effects of H,,S04  on  ciliary
beating frequency.   Grose et al. (1982) proposed that the ciliary cells may be
partially protected  due  to  the increase  in  mucus tracheal  secretion of glyco-
proteins (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  03 on the host's
ability  to  physically  remove deposited particles  (Table 10-11).  The  slowing
of mucus transport in both rat and rabbit  trachea as a result  of 03 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
                   O
bactericidal activity  but not  to significantly affect physical  removal of the
deposited  bacteria.   In these  studies,  mice were exposed  to an aerosol  of
32P-labeled  Staphylococcus  aureus either  after  a 17-hr  exposure  to  03 or

0190ZS/A                           10-108                                5/1/84

-------
                                 TABLE 10-11.  EFFECTS OF OZONE ON HOST DEFENSE MECHANSIMS:  DEPOSITION AND CLEARANCE
VQl
Ozone
concentration
ug/'m3
196
784
784
785, 1568,
1960

0.
0.
0.
0,
1,
ppm
1
,4
.4
.4, 0
.0
Measurement
method
CHEM
NBKI
ND
.8, UV
Exposure
duration and protocol
3
3
4
4
hr
hr
hr
hr
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
did not enhance 03 effect.
Delay in mucociliary clearance,
in alveolar clearance.
Silicosis
acceleration
Species Reference
Hamster Grose et al., 1980
Mouse Coffin and Gardner, 1972b
Mouse Goldstein, E. , et al., 1972
Rat Kenoyer et al . , 1981
   1764
0.5
MAST
1.4 hr
Increases nasal deposition and growth of
virus; no effect in the lungs.
                                                                                                              Mouse       Fairchild,  1974, 1977
980
980
0.
0
5
5
NBKI
NBKI
16
7
2
hr/day, No effect on clearance of polystyrene and
months iron particles.
months
Reduced clearance of viable bacteria.
Rabbit
Guinea
I Friberg
pig(Friberg
	 1 	
et
et
al.,
al.,
1972
1972
   980, 1960   0.5, 1.0  CHEM
                           2 hr
                                Reduced tracheal mucus velocity at 1.0 ppm.
                                No effect at 0.5 ppm.
                                                                     Sheep      Abraham et al., 1980
    784
0.57-2.03 M
                 17 hr before   Physical clearance  not  affected, but bactericidal
                 bacteria       activity affected at  0.99 ppm.  Decrease  in
                                deposition  of  inhaled organisms at 0.57 ppm.
                                                                                                               Mouse      Goldstein et  al. ,  1971a

-------
                                  TABLE 10-11.  EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:   DEPOSITION AND CLEARANCE   (continued)
o
i
Ozone
concentration Measurement
ug/m3 ppm method <
1176 0.62-4.25 M
1372 0.7 M
1372 0.7 G
1568 0.8 UV
1960 1.0 NBKI
2352 1. 2 UV
Exposure
duration and protocol
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
Mouse Coffin and
Rat Frager et
et al . , 1971b
et al . , 1974
al . , 1982
al . , 1980
Gardner, 1972b
al., 1979
     Measurement method:  ND = not described; CHEM = gas phase chemiluminescence; UV = UV photometry;  NBKI  = neutral  buffered potassium iodide;

      MAST = Kl-coulometric (Mast meter); M = microcoulomb sensor; G = galvanic meter.

-------
before a  4-hr exposure.   Concentrations  of 0~ were  1180,  1370, 1570, or
         o                                    J
1960 ug/m  (0.6,  0.7, 0.8, or 1.0 ppm).   The mechanical 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
                                            3
bactericidal  activity beginning at 1960 ug/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 03 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.
     Friberg et al.  (1972) studied the effect of a 16-hr/day exposure to 980
    3
ug/m  (0.5 ppm) of 03 on the lung clearance rate of radiolabeled monodispersed
polystyrene and iron  particles in the rabbit and  of  bacteria  in the guinea
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 0^ 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 radiolabeled monodis-
perse polystyrene  latex  spheres  were  followed after  0,  exposure.   A 4-hr
                     3
exposure to 1568 ug/m   (0.8 ppm) of 0, 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(S04)2,  H2$04, (NH4)2$04]
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  02  free  radicals,  which  have  the  capability for tissue  destruction.
Essentially no data  are available on the effects of prolonged exposure to 0,

0190ZS/A                          10-111                               5/1/84

-------
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 pptn) of  03.   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 0- 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 0- 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.
     These results were  confirmed  when Kenoyer  et al. (1981) repeated these
studies, with three different concentrations of 0~, 784, 1568,  and 1960 ug/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 GO-
     The velocity of the tracheal 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 0- (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
tracheal mucus velocity.
10.3.4.2   Alveolar  Macrophages.   Within the gaseous  exchange  region of the
lung, the  first line of defense against microovjanisms 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 10-12 illustrates the
effect  of  0- on the alveolar macrophage (AM).
 0190ZS/A                          10-112                               5/1/84

-------
TABLE 10-12.   EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:  MACROPHAGE ALTERATIONS
Ozone
concentration
ug/n
196
1960
392
490
980
980
980
980
980
1313
980
1960
1058
n3 ppm
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
method
NBKI
MAST
NBKI
NBKI
NBKI
NBKI
CHEM
NBKI
NBKI
UV
1 Exposure
duration and protocol
2.5 hr or
30 min 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 effects
Lung protective factor partially inactivated,
increasing fragility of macrophages (concen-
tration-related).
Increased number of macrophages in lungs
(morphology).
Decreased activity of the lysosomal
enzymes lysozyme, acid phosphatase,
and p-glucuronidase.
Increased osmotic fragility.
Decreased enzyme activity and increased influx
of PMNs.
Decreased red blood cell rosette binding to
macrophages.
Decreased ability to ingest bacteria.
Decreased agglutination in the presence of
concanavalin A.
Increased number of macrophages (morphological).
Species
Rabbit
Monkey
Rat
Rabbit
Rabbit
Rabbit
Rabbit
Rabbit
Rat
Mouse
Reference
Gardner et al. ,
Castleman et al
Dungworth, 1976
Stephens et al.

1971
. , 1977
, 1976
Hurst et al., 1970, 1971
Dowel 1 et al . ,
Alpert et al. ,
Hadley et al. ,
1970
1971b
1977
Coffin et al., 1968b
Coffin and Gardner, 1972b
Goldstein et al
Zitnik et al. ,
. , 1977
1978

-------
                           TABLE 10-12.  EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:  MACROPHAGE ALTERATIONS  (continued)
Ozone I
concentration Measurement '
ug/m3 ppm I method
1568 0.8 NO
1568 0.8 NO
1568 0.8 UV
1568 0.8 MAST
1960 1 CHEM
9800 5
1960 1 UV
3136 1.6 to NBKI
6860 3.5
4900 2.5 M
4900 2.5 M
Exposure
duration and protocol Observed effects Species Reference
11 and 24 days No effect on in vitro interferon production. Mouse

90 days Eightfold increase in number of macrophages at Rat
7 days, reducing to fourfold after 90 days.
7 days Decreased number of migrating macrophages Monkey
and total distance migrated.
3, 7, 20 days Increased phagocytosis. Rat
3 hr Decreased ability to produce interferon Rabbit
in vitro.

4 hr Decreased jji vitro migrational ability, as Rat
evidenced by decreased number of macrophages able
to migrate.
2 hr to 3 hr Decreased superoxide anion radical production. Rat
5 hr Loss of B-glucuronidase and acid phosphatase in Rat
RAM with ingested bacteria; decreased rate of
bacterial ingestion.
5 hr Diminished rate of bacterial killing, increased Rat
numbers of intracellular staphylococcal clumps;
lack of lysozyme in macrophages with staphylococcal
clumps.
I Ibrahim et al. , 1976
Boorman et al. , 1977
Schwartz and Christman, 1979
Christnian and Schwartz, 1982
Shingu et al., 1980
McAllen et al., 1981
Amoruso et al . , 1981
Witz et al., 1982
I
Goldstein et al., 1978b
Kimura and Goldstein, 1981
 Measurement method:   NO = not described;  CHEM = gas  phase cht'mi luminescence;  UV = UV photometry;  NBKI  = neutral  buffered potassium iodide;
 MAST = Kl-coulometric (Mast meter); M = microcoulomb sensor.
 Calibration method:   NBKI = neutral buffered potassium iodide.
cAbbreviations used:   PMN = polymorphonuclear leukocytes; RAM = pulmonary alveolar macrophage.

-------
     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, CL,  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 reserve 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
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 0-
can result in  a macrophage  response  in  several animal  species.  Exposure of
                  3
mice to 1058 ug/m  (0.54 ppm) of 0, 23 hr/day for a maximum of 34 days resulted
in  an  increased  number  of  macrophages within the proximal alveoli  of  the
alveolar ducts  (Zitnik  et al.,  1978).  These cells were  highly vacuolated and
contained many secondary 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 in the  number of pulmonary free cells after exposure
             3
to  1568  ug/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 both monkeys and rats  exposed to concentra-
                         3
tions as  low 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 approxi-
mately equal  in susceptibility to the  short-term effects of 0.,.
     Thus, the  total  available  data would  indicate that, after short periods
of 0~ insult, there is a significant reduction in  the number of free macrophages
available for pulmonary  defense, and  that these macrophages are more fragile,
0190ZS/A                          10-115                               5/1/84

-------
are less phagocytic, and  have  decreased enzymatic activity  (Dowell  et  al.,
1970;  Coffin et al., 1968b;  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 decrease in  the  ability  of macrophages  to phagocytize bacteria after
                                               3
exposure to concentrations as  low as 980 ug/m   (0.5 ppm) of 0-  for  3 hr was
demonstrated by Coffin et al. (1968b).  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 pg/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,  1,  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 enhance-
ment correlated well with a significant increase in cell  spreading of AMs from
exposed rats as compared  to  controls.   If 03 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  va1  d, 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 collec-
ted cells  adhered after 3 days of 0- 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 0, 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  0.,  exposure might  never have been tested, because  they  were
                O
discarded.
0190ZS/A                          10-116                               5/1/84

-------
     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
membrane organization.  A decrease in agglutination of rat AMs was found after
                             3
exposure to 980 or 1960 (jg/m (0.5 or 1.0 ppm) of 0~.  A decrease in concanavalin-
A  agglutinabi1ity  of  trypsinized red blood  cells obtained  from  rats exposed
for 2 hr to 1960 ug/m  (1 ppm) was also noted.  Hadley et al.  (1977) investigated
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 0.,-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 NO™ mixtures on the
activity of AMs isolated from the  lungs of mice exposed for 1, 2, and 3 months.
Only after  a  3-month  exposure to the  mixture of 196  ug/m   (0.1 ppm) of 0~ and
0.5 ppm  of  NO^ (3 hr/day, 5 days/week) did viability in macrophages decrease
significantly,  lr\ 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.
     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 H9/m3 (1-0
ppm) of  03  for 4 hr on the  migration rate of AMs.  Migration  was measured by
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 o^ 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.
     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
                                                     3
can be inactivated  by  a 2.5-hr exposure to  196 ug/m   (0.1 ppm)  of 0».  When
normal  AM's are  placed  in  fluid lavaged from 0.,-exposed animals, they showed
an increase  in  lysis  (10 percent over control).  A  similar effect  was seen
0190ZS/A                          10-117                               5/1/84

-------
when normal AM's were  placed in protective fluid  that  had been exposed J_n
vitro to 00.   The  data indicate that some of the effects of 00 on lung cells
           j                                                   o
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 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 (L-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, any perturbation of the metabolic
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  rabbit  macrophages  after a 3-hr  exposure  of  rabbits to
                                  3
concentrations as  low  as 490 |jg/m  (0.25 ppm)  of  03 (Hurst et al. ,  1970).
Such enzymatic reductions were also observed in AMs exposed i_n vitro (Hurst et
al. , 1971).  The ability of 0-,  to alter macrophages1  enzyme activity  was also
studied by means of unilateral  lung exposure of rabbits (Alpert et al., 1971b).
A significant reduction in these same three intracellular enzymes was found to
be specific to  the lung that breathed 0-  rather than a  generalized  systemic
                                                                           3
response.   These effects v/ere  concentration-related, beginning at 980 ug/m
(0.5 ppm)  of  0,,.   The extracellular release o.  such enzymes may occur either
as a result of direct  cytctoxic  damage and  leakage of intracellular  contents,
or they may  be selectively released without  any  cell  injury.   Hurst et al.
(1971) showed  that the reduction in  intracellular lysosomal enzyme  activity
observed after jm vitro exposure coincides with the release of the enzyme into
the surrounding medium.  In these studies, the  sum of the  intra- plus extracel-
lular enzyme activity  did  not equal   the  total  activity,  indicating  that the
pollutant  itself can  inactivate the  hydrolytic enzyme as  well  as alter the
cell membrane.   Recently, Witz et al. (1982) and Amoruso et al. (1981) reported
that i_n vivo 03  exposure affected the production of  superoxide anion  radicals
(Op) by rat  AMs.  This oxygen radical is important in antibacterial activity.

0190ZS/A                          10-118                               5/1/84

-------
Exposure to concentrations  above  3136 ug/m  (1.6 ppm) of 0~ for 2 hr appears
to result in a progressive decrease in 0? 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 09 production  arises  from the
                    *3       *                  ^
direct oxidative  damage  of the membrane enzyme involved in the metabolism of
0- to  0~,  or whether it  is a result of oxi dative  degradation  of membrane
lipids that may serve as a cofactor function.
     Shingu et al. (1980)  reported the effects of  0_  on  the ability of two
                                                    O
cell  types, macrophages  and tonsillar lymphocytes, to produce  interferon, a
substance that aids  in  defending  the host organism against viral  infections.
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
                                                                  3
tonsillar lymphocytes was  not  significantly  depressed  by  9800 ug/m   (5.0 ppm)
of 03.   The authors suggested that an impairment  of  interferon production
might play  an  important  role  in the ability  of the  host to combat respiratory
viral infections.  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.
10.3.4.3  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?  Measurement of the
competency of  the host's  antimicrobial mechanisms  can  best be tested  by chal-
lenging  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., mechanical clearance via the mucociliary escalator,
biological  clearance mediated  through macrophages,  and associated cellular and
humoral  immunological events)  that would normally  function  in  defending the
host against this  microbe,  the microbe, in its attempt to survive, would take
advantage of  these weaknesses.   A  detailed  description  of the infectivity
0190ZS/A                          10-119                               5/1/84

-------
model commonly used  for  ()„  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 CL.   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  concentra-
tions of 0-  used  do not cause any  mortality.   The mortality in the control
          O
group (clean air  plus exposure  to the microorganism) is approximately 10 to
20 percent and reflects  the natural  resistance  of the host to the infectious
agent.  The  difference in  mortality  between  the 0_-exposure group and  the
                                                   O
controls is concentration-related (Gardner, 1982a).
     If the  test  agent does not prevent  the  host's defense mechanisms from
functioning  normally,  there is a rapid inactivation of inhaled  microorganisms
that have been deposited in the respiratory system.  However, if the gaseous
exposure alters the  ability of these  defense  cells  to function, the number  of
microbes in  the lungs increases  rapidly  (Coffin and Gardner,  1972b;  Miller  et
al., 1978; Gardner, 1982a).   This acceleration  in bacterial growth  is 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 could  be predicted from a  positive  blood
culture (Coffin and Gardner, 1972b).
     Coffin  et al. (1968a)  treated mice with 03  (0.08 ppm for 3 hr) and subse-
quently exposed them to  an aerosol of  infect', jus Streptococcus sp.   In  this
study,  0-  increased the  animals' susceptibility to  infection,  resulting in a
        O
significant  increase  in  mortality rate in  the  0.,-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  to 03 at
196  ug/m   (0.1 ppm) in which the  bacterial  aerosol was  administered either
immediately  or 2, 4, or 6  hr after  cessation of the 0~ exposure.   In each
replicate experiment, 20 mice per group were used,  and 0_ was monitored  contin-
uously  by chemiluminescence.  For these postexposure challenges, only the 2-hr
time resulted  in  a  significant increase in mortality (6.7 percent)  over  controls.
0190ZS/A                          10-120                               5/1/84

-------
However, when the animals were infected with Streptococci during the actual CL
exposure,  a  significant  increase  in  mortality of 21 percent was  observed.
When this latter experimental regimen was used, exposure to 157 ug/m  (0.08 ppm)
of 0- resulted in a significant increase in mortality of 5.4 percent.
     The differences  in  results  among these studies may  have  been due to a
variation  in  the  sensitivity of the method of 0~ 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  0,-treatment  group  were  significantly  less
effective  in  combating the  infectious agent to the extent that, even at this
low concentration,  there  was a significant  increase  in  mortality rate  in  this
group over controls.   As  the 0- concentration  increased, mortality increased.
The effective  concentration  of 0-  that  increases the lung's  susceptibility  to
infection is lowered when another pollutant, such as N0? or H~SO.,  is combined
with 0.,.   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 tha   0~  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 10-13).
                                                     3
     When mice were  exposed  4 hr to 392 to 1372 ug/m  (0.2 to 0.7  ppm) of 03
and than challenged with virulent Klebsiella  pneumoniae, a  significant  in-
crease  in  mortality was noted at  785 ug/m  (0.4 ppm)  0, (Bergers et  al.,
1982).  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
0190ZS/A                          10-121                               5/1/84

-------
TABLE 10-13.   EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:   INTERACTIONS WITH INFECTIOUS AGENTS
Ozone
concentration
ug/m3
157
157-196
196
196
196, 588
392-1372
588
1372-1764
1960
2940
ppm
0.08
0.08, 0.
0.1
0.1
0.1, 0.3
0.2-0.7
0.3
0.7-0.9
1.0
1.5
Measurement3
method
ND
1 CHEM
CHEM
UV
CHEM
G
ND
ND
CHEM
ND
Exposure
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,
Observed effects
j Increase in mortality to Streptococcus pyogenes.
Significant increase in mortality during 02
exposure (Streptococcus pyogenes).
Increased mortality to Streptococcus pyogenes.
Increased susceptibility to bacterial infection
(Streptococcus pyogenes).
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
(Pasturella haemolytica).
Increased susceptibility to infection
(Streptococcus pyogenes).
Increase in Mycobacterium tuberculosis lung titers.

No effect on resistance to Mycobacterium
Species
Mouse
Mouse
Mouse
Mouse
Mouse
Mouse
Sheep
Mouse
Mouse
Mouse
Reference
Coffin et al. ,
Miller et al . ,
Ehrlich et al . ,
Aranyi et al . ,
Illing et al. ,
Bergers et al. ,
Abraham et al. ,

1968a
1978
1977
1983
1980
1982
1982
Coffin and Blommer, 1970
Thomas et al . ,
Thienes et al . ,
1981b
1965

-------
(0.4 ppm) of 0., a nearly threefold lowering of the bacterial LD™ value computed
from the three challenge doses of bacteria was found for the CL-exposed group,
indicating a significant increase in mortality.  In the same study, the authors
                          3
also found that.  1372  pg/m  (0.7 ppm)  of 0,  for  3  hr resulted  in  a  decreased
ability  of  the lung  to clear  (bactericidal)  inhaled  staphylococci.   This
finding  is  consistent with  that of Goldstein et  al.  (1971b)  and confirms
previous studies  reported  by Miller and Ehrlich  (1958)  and Ehrlich (1963).
     Chiappino and Vigiani  (1982) also reported  that 0,  (1.0 ppm) potentiated
pulmonary infection in  rats.   In this  study,  the  investigators wanted  to  know
in what  way  0., modified the  reactions to  silica in specific  pathogen-free
(SPF) rats.   Silica-treated animals were exposed to 0~  (8  hrs/day, 5  days  a
week for up to 1 year) and housed in either an SPF environment or in a conven-
tional  animal house,  thus being exposed to the bacterial flora normally present
there.    The  SPF-maintained,  0~-exposed  rats showed a  complete  absence of
saprophytes and pathogens  in the lungs, whereas  the microbial  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 0, 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 0^ exposure have
also been tested in sheep.  In these studies, sheep were infected by an inocu-
lation of Pasturella  haemolytica either 2 days  before  being  exposed  to 588
    3
ug/m  (0.3 ppm) of 0., or  2  days  after  the 0.,  exposure.   In  both cases,  the  0,,
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  al. ,
1982).    Unfortunately,  only a small number of animals  were used in each 0-,
treatment (n = 3).
     Thomas et al. (1981b)  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  03  3  hr/day,  5 days/week  for  up
to 8 weeks,  initiated 7  or 14  days  after the  infectious  challenge  with
Mycobacterium tuberculosis H37RV, resulted in significantly increased  bacterial
lung titers,  as  compared  with controls.  Exposure to lower concentrations of
0190ZS/A                          10-123                               5/1/84

-------
0, did not  produce  any significant effects.   In an earlier study, Thienes et
                                                o
al. (1965)  reported  that  exposure to 2940 ug/m  (1.5 ppm) of (k  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 10-14 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  03 and NO,, mixtures.   In the  earlier studies (Ehrlich et al. ,
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/m  (1.5 to 5 ppm)  of NO  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  5 months  to  mixtures of  196 ug/m   (0.1 ppm)  of 0, and  940 ug/m  (0.5
                                                        «3
ppm) of N02 and challenged with bacterial  aerosol.   The 1- or  2-month exposure
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 0, and N09 mixture
                                                            O       £.
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
                                                                            o
superimposed 3-hr daily peak  (5 days/week) containing a mixture of 196 \jg/m
(0.1 ppm)  of 03 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 not statistically 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
mixture of 03,  SO,,,  and  (NH.^SO..   The concentrations were 200 ug/m  of 0.,,
13.2 mg/m   of  SO,,,  and 1.04 mg/m  of (NhOpSO,.  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 OT was the major constituent of the mixture affecting the host's suscepti-
bility to infection.   These  investigators also measured a number of specific
0190ZS/A                          10-124                               5/1/84

-------
TABLE 10-14.   EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:   MIXTURES
Ozone



concentration Measurement Exposure ^
ug/m3 ppm Pollutant
98 0.05 03 +
3760 N02
98-196 0.05-0.1 03 +
100-400 N02 +
1500 ZnS04
196 0.1 03 +
241-483 : H2S04
196 0.1 03 +
S 900 H2S04
i

rn j
196 0.1 ' 03 +
1090 H2S04

196 0.1 03 +
980 N02
196 0.1 03 +
980 N02

196 0.1 03 +
13200 S02 +
1040 (NH4)2S04




method duration and protocol
CHEM 3 hr

CHEM 3 hr


CHEM 3 hr

CHEM 3 hr +
2 hr



CHEM 3 hr +
2 hr

CHEM 3 months

CHEM 1-6 months


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




Observed effects Species
Exposure to mixtures caused synergistic Mouse
effect after multiple exposures.
Additive effect of pollutant mixtures Mouse
with infectivity model.

Increased susceptibility to Streptococcus Mouse
pyogenes.
Sequential exposure resulted in signifi- Mouse
cant increase in respiratory infection.
Neither alone produced a significant
effect.

Reference
Ehrlich et al. , 1977

Ehrlich, 1983


Grose et al. , 1982

Gardner et al. , 1977




Sequential exposure resulted in signifi- Hamster 1 Grose et al . , 1980
cant reduction in ciliary beating
activity.
Significant decrease in viability of Mouse
alveolar macrophages.
At 3 and 6 months, susceptibility to Mouse
pulmonary infection increased sig-
nificantly. Delayed clearance rate.
Highly significant increase in Mouse
susceptibility to infection. Effects
attributed to 03. Increased bactericidal
rate. Mixture showed greater growth
inhibition in leukemia target cells and an
increase in blastogenic response to PHA,
Con-A, and al loantigens.


Ehrlich, 1980

Ehrlich, 1980


Aranyi et al. , 1983







-------
                                        TABLE 10-14.  EFFECTS OF OZONE ON HOST DEFENSE MECHANISMS:  MIXTURES   (continued)
o
 I
CT>
Ozone
concentration
flgTifr3 ppm
216-784 0.11-0.4
3760-13720 2-7.3
980 0.5
11-3000
1568 0.8
3500
3500
3500
Pollutant
03 +
NO 2
03 +
H2S04
0, H-
Fe2(S04)2 +
H2S04 +
(NH4)2S04
Measurement 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 et al., 1978
secretion; synergism reported;
effect reversible in clean air.
Exposure to mixtures produced same Rat i Phalen et al., 1980
effect as exposure to 03 alone.
j  Measurement method:
  Abbreviations  used:
                         CHEM = gas phase chemi luminescence; UV = UV photometry; A = amperometric.
                         PHA = phytohemagglutinin; Con-A = concanavalin-A

-------
host defenses  to  determine  if host response  to  the  mixtures  was  significantly
different  from that of the controls or  of the  O^-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 0--alone and the control animals.
     Previous  studies  (Gardner et al., 1977) indicated that a  sequential expo-
sure to  Oo followed by H^SO^  significantly increased  streptococcal  pneumonia-
induced mortality rates in  mice.   Ozone  and  H2SO. had  an additive  effect when
exposed  in this sequence.   However, the  reverse  sequence did  not affect inci-
dence of  mortality.   Grose  et al.  (1982)  expanded  these  studies and showed
                                                        3
that a  3-hr exposure to the combined aerosol of  196 ug/m   (0.1 ppm) of  0-  and
                3
483  or  241 ug/m  H^SO^ also significantly increased the percent  mortality,  as
compared to control.
     The  effects  of exercise  on the response to low  levels  of 0., were also
                                                                      3
studied by using  the infectivity model.  Mice were exposed to  196 ug/m  (0.1 ppm)
of 03 and 588  ug/m  (0.3  ppm) of  0, for  3 hr while  exercising.   Each exposure
level yielded  mortality rates that were significantly higher than those observed
in the  0~  group that was not exercised (Illing et al., 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.
10.3.4.4   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
10-15).
                              o
     The  effects  of 2900 \jg/m  (1.48 ppm) of 0, 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
0190ZS/A                          10-127                               5/1/84

-------
                                       TABLE 10-15.   EFFECTS OF OZONE ON HOST DEFENSE  MECHANISMS:   IMMUNOLOGY
Ozone
concentration Measurement Exposure
pg/m3
196
H- 980, 2940
o
t
C»;980, 1568
i
980-2900
I
1150
ppm method duration and protocol
0.1 UV 5 hr/day,
5 days/wk for
90 days
0.5, 1.5 NO 4 hr
0.5, 0.8 Mast Continuous
3-4 days
0.5-1.48 CHEM 3 hr
0.59 NO 36 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
Levamisole 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 ppm for 5 days. Hemagglutination antibody
titers increased over control.
Impaired resistance to toxin stress.
Immunosuppression.
Species Reference
Mouse Aranyi et al., 1983
Mouse Goldstein et al., 1978a
Mouse Osebold et al . , 1979, 1980
Gershwin et al . , 1981
Guinea pig Thomas et al . , 1981b
Mouse Campbell and Hilsenroth,
1976
Measurement method:   ND = not described;  CHEM = gas phase  chemiluminescence;  UV  =  UV  photometry;  Mast = Mast meter.
Abbreviations used:   PHA = phytohemagglutinin; con-A = concanavalin-A;  LPS  =  lipopolysaccharide;  IgE = immunoglobulin-E;  IgA = immunoglobulin-A.

-------
with Mycobacteriurn tuberculosis, the cutaneous sensitivity to PPD was signifi-
cantly affected  by  0».   The diameters  of  the erythemas from the 03-exposed
animals were significantly smaller during the 4 to 7 weeks after the infectious
challenge,  indicating  depressed cell-mediated immune response.  Exposure  to
980 ug/m3 (0.5 ppm) of 0_ 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  (200 ug/m  of 0-  5 hr/day,  5 days/week for 90 days).   No
alteration  in  the  response to  alloantigens or the  B-cell  mitogen lipopoly-
saccharide  (IPS)  was  noted,  but a  statistically  significant suppression in
blastogenesis to the T-cell mitogens (PHA and con-A) was detected.   The authors
suggested  this  indicated that  the  cellular mechanisms for  recognition  and
proliferation to  LPS  and alloantigens  were  intact,  but  0, might  have been  in-
                                                        O
terfering 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).
     Campbell and Hilsenroth (1976) used a toxoid immunization-toxin challenge
approach  to  determine  if continuous exposure to  1150 ug/m   (0.59 ppm) of 0.,
                                                                           O
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 03-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  stephylococci  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
tested,  2940 and 980  ug/m3  (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.
0190ZS/A                          10-129                               5/1/84

-------
     The possibility that 03 may be responsible for the enhancement of allergic
sensitization has  important  implications  for human health effects.  Gershwin
et al.  (1981) reported  that  0-  (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.  In  general,  the  number of IgE-containing cells  correlated  positively
with levels of anaphylactic 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).  These authors showed that a signi-
ficant increase  in  anaphylactic sensitivity  occurred when antigen-stimulated
and  Og-exposed  animals  were  compared  to controls (Osebold et  al., 1980).
Significantly greater numbers  of animals were allergic  in experimental groups
which 0- exposure  ranged from  0.8 ppm to 0.5 ppm for 3  days.  Further studies
are needed to determine the threshold level  of these  effects.

10.3.5  Tolerance
     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 11, 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
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 forms,
tolerance and adaptation, are  not yet fully  understood  in any concise manner.

0190ZS/A                          10-130                                5/1/84

-------
Further research  is needed to  focus  on the specific structural/functional
response being addressed in each specific experimental situation and to relate
this to the  exposure  concentration and profiles of 03-   An attempt to relate
the modified  responses  seen  in animals to what  has  been described in human
subjects will be made in Chapter 13.
     In animal  oxidant  toxicity studies, the term tolerance  classically is
defined as the  phenomenon  wherein a previous exposure to a nonlethal dose of
0« will provide some protection against a subsequent  exposure to a dose  of  CL
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 morpho-
logical changes.  This protective phenomenon seen with oxidants was originally
described by Gildemeister  (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
         O
early  as  1956,  Stokinger et al.  presented  data  clearly indicating that an
animal could  also become tolerant to the lethal  effects of 03-  Such tolerance
has also been reported by many investigators, including Matzen (1957a),  Menden-
hall and Stokinger  (1959),  Henschler (1960), and Fairchild (1967).   The obser-
vation of this  tolerance phenomenon in experimentally exposed animals has led
to the speculation  that it may also be a mechanism for protecting environmental-
ly exposed humans.
     The previous  criteria document  for 0_  and  other photochemical  oxidants
cited  various studies  (Table 10-16)  that examined 03 tolerance and presented
some  evidence indicating possible mechanisms  of  action.   Review  of these
earlier data  reveals  that  pre-exposure to a certain  concentration of 03 can
protect test  animals from the acute lethal effects of a second exposure to 03-
This protection has been  attributed to a significant reduction  in pulmonary
edema  in the pre-exposed  animals.  Table 10-16  lists the  key studies on 03-
     Because  0-  has a marked proclivity to  reduce the ability of alveolar
              <3
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 10.3.4.3),  the pre-exposed (tolerant)
animals were  only partially  protected from  the  aerosol  infectious challenge

0190ZS/A                          10-131                               5/1/84

-------
TABLE 10-16.   TOLERANCE TO OZONE
Ozone
(ug/m3)
pre-
exposure
196-1960
196
490
i— '
0
1
t-1 980
OJ
ro
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
(ug/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
1, 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.
lolerance 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 al., 1971a
with methylprednisolone.
Edema as measured by recovery of 1321
in pulmonary lavage fluid.
Tolerance to edema effects of 0., did not Mouse Gregory et al., 1967
develop in thymectomized animals but
developed in sham-operated animals, in- i
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 0, 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 a local effect and seen Alpert and Lewis, 1971
only in the pre-exposed lung.

-------
                                                            TABLE 10-16.   TOLERANCE TO OZONE  (continued)
Ozone
(pg/rn3)
pre-
exposure
980
t— '
V 1470
i—1
OJ
CO
1600
1960
1960
Ozone
Ozone (ug/m3)
(ppm) Length of after
pre- pre- latent
exposure exposure period
0.5 3 hr 5880 or
0.75 3 days 7840
0.8 4 hr 2352
1 1 hr ND
1 1 hr 3920
Ozone Length of
(ppm) exposure
after after
latent latent
period period
3 and 22 3 hr
4.0 8 hr
1.2 4 hr
ND ND
2 1 hr
Observed effect(s) Species Reference
With unilateral lung exposure technique, Rabbit Gardner et al., 1972
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 Chow, 1976
peroxidase, glutathione reductase, glucose-6- Chow et al., 1976b
phosphate dehydrogenase and levels of reduced
glutathione in lungs of tolerant animals, as
compared to nontolerant animals.
Pre-exposure to 03 caused complete tolerance to Rat Frager et al., 1979
delay in mucociliary clearance at 3 days, but
not 13 days.
All animals X-irradiated to 800 R. 60% of Mouse Hattori et al . , 1963
03-pre-exposed mice survived. 100% of
controls died.
Tolerance to allergic response to inhaled Guinea Matsumura et al . , 1972
acetylcholine. pig
ND = not described.

-------
(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 ug/m  (0.10 ppm)  of 03,  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 al.,  1972),  studies  were conducted to evaluate the effects of tolerance at
the cellular  level.  These  studies indicated  that the initial Q3 exposure  did
induce tolerance against pulmonary  edema in the exposed lung;  however, there
was no protection afforded against the cytotoxic effects of  03  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 polymorphonuclear  leukocytes—PMN)
showed that there was no protection against these cellular defense mechanisms.
     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
00 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 0., 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  10.3.1.2, and they  indicate
that with  continuous exposure  to 03,  the lung attempts  to initiate the repair
of  the 03  lesion, resulting in  somewhat reduced  or less than  expected total

0190ZS/A                         10-134                               5/1/84

-------
 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 03 (Evans  et  al.,  1976a;  Dungworth  et  al.,  1975a;  Schwartz  et al.,  1976).
 Thus, there is  also  the possibility that this reparative-proliferative  response
 relines the airway epithelium with cells that have a biochemical armamentarium
 more  resistant  to oxidative  stress (Mustafa  et al.,  1977; Mustafa  and Lee,
 1976).
      Another  suggestion  is that with  03,  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 al.,
 1978).  As the airways  become  more obstructed, the 03 molecules  are less likely
 to penetrate  to lumen.  This  may result in a filtering system that removes the
 0, from the inhaled  air before  it reaches the alveoli.
      More recently,  there  have  been additional  studies examining the protective
 effects  of this tolerance  phenomenon in animals.  Frager et al. (1979) studied
 the possibility of tolerance  to 0- in mucociliary clearance.  Exposure of rats
 to 1.2  ppm of 03 following particle deposition caused a substantial  delay in
 mucociliary clearance.   This 0- effect could be eliminated by  a pre-exposure
            3
 to 1600 [jg/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 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 0^.  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 (Nilding  and  Hi!ding,  1966) or the  occurrence of some other temporary
 change  in the regenerating ciliated cell.
0190ZS/A                          10-135                               5/1/84

-------
     Tolerance to 0_  has  also  been studied by using a variety of biochemical
indicators to measure the  extent to which a  pre-exposure  to  0~  protects or
                                                               O
reduces the host  response  to  a subsequent exposure.   Chow  et al.  (1976a,b)
compared a variety  of metabolic  activities of the lung immediately after an
initial 3-day continuous  exposure to 1600 ug/m  (0.80 ppm) of 0  with the
                                                                 O
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 rou-
tine.   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.
     A  number of  biochemical  effects of 0-  gradually subside in intensity
despite continued exposure  to  the gas.   An example is prolyl  hydroxylase,  a
key enzyme  in collagen  synthesis that serves as  a sensitive indicator of
disturbances in collagen metabolism.  The  increase in  superoxide anion  caused
by  0-  stimulates  collagen synthesis.  This is  reflected  in an  increase  in
prolyl hydroxylase activity.  Bhatnagar et al. (1983)  have shown that there is
also an increase in superoxide dismutase, which quenches free radicals, thereby
restricting hydroxylase  activity and collagen formation.
     Gertner et al. (1983b) presented data showing that  the  development of
adaptation and tolerance is rapid and mediated through the vagus nerve.  These
investigators used  a  bronchoscope to study the  response  of the  periphery of
the lung to  0~  exposure.   Measurements of collateral  resistance (Rcoll) were
used to monitor the response to  0~.   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
0_  exposure ceased, the  Rcoll  returned to  normal.   Subsequent exposure  to 0.1
ppm of  0_  did not  increase Rcoll,  indicating that some protection existed.
These investigators 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.
0190ZS/A                          10-136                                5/1/84

-------
     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 0« 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.
     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 in the face of effective tolerance for
edema  production.   This  suggests that the  chemotactic  effect  of 0., may  be
                                                                   «J
separable from  the edemagenic effect.  This  may explain why chronic morpholo-
gical  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 al. , 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 0~  exposure  but did after
challenging with  hi stamina,  indicating that the parasympathetic system  may
play a  role in  response  to 03  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

0190ZS/A                          10-137                               5/1/84

-------
changes, such as an  increase in mucus thickness, that may, in effect, prevent
03 from 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 with great potential  for possible
future harm.
10.4  EXTRAPULMONARY EFFECTS OF OZONE
10.4.1    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 11), and the possible
implications of these and other symptoms as indications of low-level 0., effects,
few recent  reports  were found on behavioral  and central nervous system (CNS)
effects of 03 exposure in animals.  Table 10-17 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 03 by measuring their reaction to a 30-min exposure on one
side of an  annular plastic  mouse chamber.  A concentration-related avoidance
of the 03 side was  reported at  1176  to 16,660  ug/m   (0.60 to 8.50 ppm) of  Og.
However,  the  study  had  serious shortcomings,  including a  lack of position-
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 0.,.   Mice were individually exposed to 0^ 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 ug/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 ug/m  (10 ppm)  of  CL,  all  animals turned off
100 percent of the  trials with an average latency of  approximately 10 sec.
0190ZS/A                          10-138                               5/1/84

-------
                                         TABLE 10-17.   CENTRAL NERVOUS SYSTEM AND BEHAVIORAL EFFECTS OF OZONE
Ozone
concentration
ug/m3
110
98-1960
196-3920
o 235-1960
^ !
GO
392-980
588-1372
ppm
0.056
0.05-1.0
0.1-2.0
0.12-1.0
0.2-0.5
0.3, 0.7
Measurement3' Exposure
method duration and protocol
d 93 days, continuous
MAST 45 min
CHEM 6 hr
CHEM 6 hr
NBKI 6 hr
MAST 7 days, continuous
Observed effects(s)c Species Reference
No overt behavioral changes. Cholines- Rat Eglite, 1968
terase activity inhibited at 75 days of
exposure, returning to control levels
12 days after termination of exposure.
Motor activity progressively decreased Rat Konigsberg and
with increasing 03 concentrations up to Bachman, 1970
0.5 ppm. Slight increase in frequency
of 3-min intervals without motor activity.
Linear and/or monotonic decreases in Rat Weiss et al . ,
operant behavior during exposure. 1981
Wheel running activity decreased Rat Tepper et al.,
monotonically with increasing 03 con- 1982
centration. Components of running
were differentially affected at low
vs. high 03 concentrations.
Wheel running activity decreased 50%. Mouse Murphy et al . ,
1964
Running activity decreased 60% during
                                                                           first 2 days, returning to control
                                                                           levels during the next 5 days of expo-
                                                                           sure; running activity decreased 20%
                                                                           when 0.3 ppm exposure was followed
                                                                           immediately by 0.7 ppm 03 exposure.
                                                                           Adaptation with continued exposure
                                                                           was apparent.
980-19600 0.5-10.0
                              CHEM
                                               60 s
                                                                           Exposure terminated by nose pokes with
                                                                           increasing frequency .TS Q~ concentra-
                                                                           tion increased.
Mouse      Tepper  et  al.
           1983
980       0.5
                              NU
                                                  mm
                                                                           Elevation of simplp and choice reactive
                                                                           time.
Nonhuman   Reynolds  and
primate    Chaffee,  1970

-------
                                        TABLE  10-17.   CENTRAL  NERVOUS  SYSTEM AND  BEHAVIORAL  EFFECTS  OF  OZONE   (continued)
c
I
Ozone
concentration
ug/m3
980-1960
1176-
16,660
1960
1960-
5880
1960
1960-
5880
ppm
0.5, 1.0
0.6-8.5
1.0
1-3
1
1-3
Measurement ' Exposure
method duration and protocol
NBKI 1 hr
e 30 min
MAST 7 days, continuous
NBKI
ND 18 months,
8 hr/day
18 months;
8, 6, 24 hr/day
ND 18 months,
8 hr/day
Observed eftects(s)
Evoked response to light flashes in the
visual cortex and superior collicus
decreased after exposure.
Avoidance behavior increased with
increasing 03 concentration.
Reduction in wheel running activity; no
effect of Vitamin E deficiency or
supplementation.
COMT activity decreased at 2 ppm, MAO
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 EEC patterns after
9 months, but not after 18 months of
exposure.
Species Reference
Rat ' Xintaras et al . ,
1966
Mouse Peterson and
Andrews, 1963
Rat Fletcher and
Tappel, 1973
Dog Trams et al . ,
1972
Dog Johnson et al.,
1974
    Measurement method:  MAST  =  Kl-coulometric  (Mast meter); CHEM =  gas  phase  chemiluminescence; NBKI =  neutral  buffered potassium  iodide;  ND = not
    described.
    Calibration method:  NBKI = neutral buffered potassium  iodide.
   Abbreviations  used:  COMT  =  catechol-o-methyltransferase; MAO =  monamine oxidase;  EEC = electroencephalogram
    Spectrophotometric method  with  dihydroacridine.
    KI titration with sodium thiosulfate.

-------
     Studies by Murphy et a "I.  (1964) demonstrated that wheel-running  activity
                                                                               3
decreased by approximately 50 percent when mice were exposed to 392 to 980 ug/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 pg/m  (0.30  ppm)
of 00.   Running activity gradually  returned to pre-exposure values  during the
                                              3
next 5 days of continuous exposure  to 588 pg/m   (0.30 ppm) of  Ov  If  the  same
                                           3
mice were subsequently exposed to 1372 pg/m  (0.70 ppm) of 0- for an additional
7 days,  running activity v/as  depressed  to 20 percent of pre-exposure values.
                                                                             3
Partial recovery was described during the final  days of exposure to 1372 pg/m
(0.70 ppm), and  complete  recovery occurred several  days  after exposure was
terminated.  However, partial  tolerance was seen when  the  air-control  mice
were subsequently exposed to 392 pg/m  (0.20 ppm) of 0- for 7 days.  Konigsberg
and  Bachman  (1970)  used a  capacitance-sensing device to record  the motor
activity of rats during a 45~min exposure to 98,  196, 392, 980, and 1960 pg/m
(0.05, 0.10, 0.20,  0.50,  and  1.0 ppm)  of 0~.   Compared with control rats,
motor activity following 0,  exposure progressively decreased with increasing
                                3
0, concentrations up to 980 pg/m  (0.50 ppm).   No greater reduction was obtained
            o
at 1960 pg/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 0, concentration.
     A detailed  microanalysis 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 OT-   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 pg/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 Q~.  An increase in
the time interval between running bursts primarily accounted for the decreased
                                   3
motor activity at the low (235 pg/m  , 0.12 ppm) 0, concentration.   Postexposure
increases  in wheel running were seen following this  low 0,, concentration.   At
                                    3
higher 0-  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.
0190ZS/A                          10-141                               5/1/84

-------
     Effects of  03  on  behavior  were further investigated by  Weiss  et al.
(1981) in their  studies on  the operant behavior of rats during CL 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
                                                                      3
rats were exposed for  6 hr  to CL 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
                                                                      3
afternoon.   Ozone-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
                               3
monotonic from 196 to 3920 ug/m  (0.10 to 2.0 ppm).   Analysis of the distribu-
tion  of  responses  during 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  contrast  to  other types of toxicants,  0, did not disrupt
the temporal pattern  that characterized response during each  fixed-interval
presentation.   Based on tha 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 0.,.  One-hour air perfusion following
the 0,,  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, includ-
ing O.,, 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
1960, 3920, or  5880 ug/m  (1, 2,  or  3 ppm)  of 03-   In  8  hr/day exposures,
reported  decreases   (35 percent)  in the catecholamines  norepinephrine and

0190ZS/A                           10-142                               5/1/84

-------
epinephrine  were not  statistically  significant,  although  0-  exposure at
         3
3920 ug/m   (2 ppm)  caused  a  statistically  significant  decrease  in  catechol-o-
methyl-transferase  (COMT)  activity.    In  contrast,  monamine  oxidase (MAO)
activity was  significantly elevated  at 1960 ug/m  (1  ppm)  of 0,,  but  not at
                 3                                                             3
3920 or 5880 ug/m   (2 or 3 ppm) of 0~.  Increasing daily exposures to 1960 ug/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.  (1974)  measured electroencephalographic (EEC)
patterns in the  same dogs and noted alterations in EEC patterns after 9 months
                                 3
of exposure  to  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
EEC 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 about chronic  0, exposure.

10.4.2  Cardiovascular Effects
     Very few reports  on the cardiovascular  effects of  0.,  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 0.20 ppm
of 0.3  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 03 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  03 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
combined,  in rats.   The rats were exposed  to 0.60 ppm  ozone of 0.,  5  hr/day for
                               3
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.

0190ZS/A                          10-143                               5/1/84

-------
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
                                   3
exposed to  392, 1568,  or 3920 ug/m  (0.2, 0.8,  or  2  ppm)  of 0., 6 hr/day,  5
days/week for 62 exposure days as part of a more extensive evaluation of lung
function  (Section 10.3.2).  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
and 1960 ug/m  (0.3 and 1.0 ppm) of 0- on the pulmonary gas-exchange region of
dogs ventilated through an  endotracheal tube.   Pulmonary capillary  blood flow
and arterial 0?  pressure  (Pa02) were decreased 30 min following  exposure  to
both 0., concentrations, and arterial pH (pH ) was decreased following exposure
      o     o                              a
to 1960 ug/m   (1.0 ppm) of Ov   Decreases in pulmonary capillary blood  flow
                                                       3
persisted 24 hr  following  exposure  to 588  and 1960 ug/m   (0.3 and 1.0 ppm)  of
03 and as  long as 48 hr following  exposure  to  1960  ug/m   (1.0 ppm) of  0.,.
Persistent decreases in pH  and Pa00 were observed 24 hr following exposure to
         3
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  0.,  exposure
can cause both acute and delayed changes in cardiopulmonary function.

10.4.3  Hematological and  Serum 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
0- or some reaction product of 03 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  i_n vitro  and i_n vivo
0.,  exposure.   Hematological 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.
 0190ZS/A                          10-144                                5/1/84

-------
10.4.3.1   Animal  Studies  -  In Vivo  Exposures.   The effects  of  jm vivo  03
exposure in animals,  including  studies  reviewed in the previous 0^ criteria
document (U.S.  Environmental Protection  Agency,  1978),  are summarized in Table
10-18.
     Effects of 0- on the blood were first reported by Christiansen and Giese
(1954)  after they  detected  an  increased resistance to  hemolysis of RBCs from
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  03 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 03 concentration 8 hr/day for 7 days.
     In more recent  studies, Clark  et al. (1978) investigated the  biochemical
changes in  RBCs of squirrel monkeys exposed  to  1410 ug/m  (0.75 ppm) of 0,
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 ug/m  (1
or  3 ppm)  of  0~ 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 0.,.
     Alterations in RBC morphology have been previously observed in 0,,-exposed
laboratory  animals and man  (Brinkman  et al. , 1964; Larken  et  al. ,  1978).
Similar observations  have recently  been  made in monkeys exposed to 1254 ug/m
(0.64 ppm)  of  03   for  8 hr/day over a  1-yr  period (Larkin et al. ,  1983).
Ultrastructural  SEM  studies of RBC's following exposure  to  03 demonstrated

0190ZS/A                          10-145                               5/1/84

-------
                                                     TABLE 10-18.  HEMATOLOGY:  ANIMAL  —  IN  VIVO  EXPOSURE
o
 I
CTV
Ozone
concentration Measurement
ug/m3
110
118
235
470
941
392
392
392
392-
1960
490
980
1372
627
784
ppm method
0.056 c
0.6 UV
0.12
0.24
0.48
0.2 ND
0.2 UV
0.2 UV
0.2- UV
1.0
0.25 UV
0.50
0.70
0.32 UV
0.4 ND
Exposure
duration and
protocol
93 days
2.75 hr
4 hr
8 hours/day,
5 days/week,
3 weeks
60 min
1-4 hr
2.75 hr
6 hr
± dietary
vitamin E
G hr/day,
5 days/week,
6 months
Observed effect(s) Species
Decreased whole blood chol inesterase, Rat
which returned to normal 12 days after
exposure ceased.
RBC survival decreased at 0.06, Rabbit
0.12, and 0.48 ppm; no concentra-
tion-response relationship.
Increased osmotic fragility and Rabbit
spherocytosis of RBC's.
Increased serum glutamic pyruvic Mouse
transaminase and hepatic ascorbic
acid. No change in blood catalase.
Small decrease in total blood sero- Rabbit
tinin.
Plasma creatine phosphokinase Mouse
activity altered immediately and
15 min postexposure; no effect
30 min postexposure. No change
in plasma histamine or plasma
lactic acid dehydrogenase.
RBC survival decreased at 0.25 ppm Sheep
only.
Increased erythrocyte G-6-PD and Mouse
decreased AChE (both diets).
Increased plasma vitamin E
(both diets).
No change serum trypsin inhibitor Rabbit
capacity.
Reference
Eglite, 1968
Calabrese et al . ,
1983a
Brinkman et al . ,
1964
Veninga, 1970
Veninga, 1967
Veninga et al . ,
1981
Moore et al. ,
1981a
Moore et al . , 1980
P' an and Jegier ,
1971

-------
                                           TABLE 10-18.  HEMATOLOGY:  ANIMAL  —  IN VIVO  EXPOSURE   (continued)
o
 I
Ozone
concentration
ug/m3
784
784
784
980
980
980
1254
1470
ppm
0.4
0.4
0.4
0.5
0.5
0.5
0.64
0.75
Exposure
Measurement duration and
method protocol
NO 6 hr/day,
5 days/week,
10 months
ND 10 months
ND 6 hr/day,
5 days/week,
10 months
UV 2.75 hr
MAST Continuous,
23 days
NBKI 8 hr/day,
7 days
i
UV 8 hr/day,
1 yr
ND 4 hr/day,
4 days
Observed effect(s)
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
Rabbit
Rabbit
Rabbit
Sheep
Rat
Monkey,
rat
Monkey
Monkey
Reference

Jegier, 1973
P'an and
1972
P'an and
1976
Moore et
1981b
Menzel et
Jegier,
Jegier,
al.,
al., 1972
Chow et al. , 1975
Larkin et
Clark et
al., 1983
al., 1978

-------
                                            TABLE  10-18.   HEMATOLOGY:   ANIMAL  —  IN  VIVO  EXPOSURE   (continued)
o
 i
CO
Ozone
concentration
ug/m3 ppm
1568 0.8




1568 0.8
1568 0.8
1666 0.85

1686 0.86
1960 1.0
1960 1.0


Exposure
Measurement duration and
method protocol
NBKI 7 days




NBKI 8 hr/day,
7 days
NBKI Continuous,
29 days
MAST 4 hr

NO 8 hr/day,
5 days/week,
6 months
UV 4 hr ± vitamin E
ND 30 min


Observed effect(s) Species
Increased activity of GSH pero- Rat
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 G-6-OP,
catalase, and superoxide dismutase
and in levels of thiobarbituric acid
reactants, methemoglobin, hemoglobin,
and reticulocytes.
No change in total lactate dehydro- Monkey
genase activity or isoenzyme pattern
in plasma or erythrocytes.
Increased lysozyme activity by Rat
day 3.
Increased Heinz bodies in RBC's Mouse
(decreased with continual exposure).
Increased infestation and mor- Mouse
tality after infection with
Plasmodium berghei. Increased
acid resistance of erythrocytes.
Decreased filterability. No pro- Mouse
tection by vitamin E. No lipid
peroxidation.
Increased resistance to erythrocyte Mouse
hemolysi s.


Reference

Chow and Kaneko,
1979




Chow et al . ,
Chow et al . ,
Menzel et al

Schlipkoter
Bruch, 1973
Dorsey et al
Mizoguchi et
1973;
Christiansen
Giese, 1954




1977
1974
. , 1975a

and
. , 1983
al. ,
and


-------
                                    TABLE 10-18.   HEMATOLOGY:   ANIMAL — IN VIVO EXPOSURE  (continued)
Ozone
concentration
ug/ni3
1960-
3920
1960
5880
2940
11,760
15,680
7840
9800
19,600
ppm
1.0
2
1.0
3.0
1.5
6.0
8.0
4.0
5.0
10.0
Exposure
Measurement duration and
method protocol
CHEM 2 or 7 days
CHEM 4 hr
UV 3 days
4 days
4 days
UV 2, 4, or 8 hr
CHEM 3 hr
I 1 hr/week,
6 weeks
Observed effect(s)
No changes.
No effects on oxyhemoglobin affinity,
2,3-DPG concentrations, heme-02
binding.
No effect on SOD, GPx, K+ influx
ratios (all levels). Increased Hb,
Hct, echinocytes II & III (6 & 8 ppm);
echinocytes correlated with petechia
in lungs, indicative ot vascular
endothelial damage.
Increased plasma concentrations
of PGF2a and PGE2.
Inhibition of serum monoamine oxi-
dase.
Production of serum antibodies that
reacted with ozonized egg albumin
but not native ovalbumin.
Species Reference
Rat, Cavender et al . ,
guinea pig 1977
Rabbit Ross et al. , 1979
Rat Larkin et al . , 1978
Rat Giri et al. , 1980
Rat Suzuki , 1976
Rabbit Scheel et al., 1959
Measurement method:  NO = not described; CHEM = gas phase chemiluminescence; UV - UV photometry; NBKI = neutral buffered potassium iodide;
MAST = KI - coulometric (Mast meter); I = iodometric.

Abbreviations used:  RBC = red blood cell; G-6-PO = gl ucose-6-phosphate dehydrogenase; AChE = acetylcholinesterase; GSH = reduced
glutathione; GSH peroxidase = glutathione peroxidase;  GSH reductase = glutathione reductase; FA = fatty acid; LDH = lactic dehydrogenase;
2,3-DPG = 2,3-diphosphoglycerate; SOD = superoxide dismutase;  GPx = glutathione peroxidase; K = potassium, Hb = hemoglobin; Hct = hemato-
crit; PGF2a = prostaglandin F2a; PGE2 = prostaglandin  E2.

Spectrophotometric method using dihydroacridine

-------
reduced numbers  of normal discocytes and  increased  numbers of knizocytes,
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 Oo~exposed animals.
                                                                               3
Moore et al. (1981a) reported reduced RBC survival  in sheep exposed to 490 ug/m
(0.25 ppm) 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 ug/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
vitamin E-deficient rats to 1568 ug/m  (0.8 ppm) of 0., 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
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
blood  of 9-month-old C57L/J  mice exposed to 627 ug/m   (0.32 ppm)  of 0~ 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 0~  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 ug/m  (0.3 ppm), 1372 pg/m  (0.7 ppm),
or  1960 ug/m  (1.0 ppm)  of 0- 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-pm 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  0,. and incubation
of  RBCs for 6 hr  resulted  in  a significant increase in filtration time of RBCs
0190ZS/A                          10-150                               5/1/84

-------
from 0,-exposed mice,  and a lack of protection  by  dietary vitamin E.  The
                                                            3
hematocrit of  vitamin  E-deficient mice exposed to 1960 ug/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.
10.4.3.2  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 10-19.
     The  effects  of  in  vitro  0- exposure on human RBCs  have  been  evaluated  by
                        ..      ^
using a number of different end points, such as increases in complement-mediated
cell damage (Goldstein, B., et al.,  1974),  formation of Heinz bodies (Menzel et
al.  , 1975b),  decreases in  RBC  native protein fluorescence (Goldstein and
McDonaugh, 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 0~ exposure.   (See Table 10-20 for a summary
of  the human J_n vitro studies.)
     Kesner et al. (1979) demonstrated that 0.,-treated phosphol ipids inhibited
RBC membrane ATPase. .  Addition  of semicarbazide to  O^-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  0? through the  solution at 4  ml/min  (2.5 pM/min  of 0~)
for 1 or  2  min.   Semicarbazide (40 uM) or PABA  (40  uM)  inhibited  the  cross-
linking of 03-exposed spectrin.   The inhibition of AChE and hexokinase activities
of  RBC ghosts  exposed  to  CL was also partially prevented by  these  two  agents,
        +                J
as  was K   influx  into  whole RBCs.  The authors  attributed  the inhibition of
ATPase  to oxidation of phospholipids with subsequent cross-linking of membrane
protein by lipid  peroxidation products.   Because the  reaction of  ozonolysis

0190ZS/A                          10-151                               5/1/84

-------
                                                    TABLE 10-19.   HEMATOLOGY:  ANIMAL — IN VITRO EXPOSURE
en
ro

Exposure
Ozone Measurement duration and
concentration method protocol
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 aminotriazole.
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 10-20.  HEMATOLOGY:  HUMAN - IN VITRO EXPOSURE
o
 i
en
OJ

Ozone3
concentration
980 ug/m3 (0.5 ppm)
1960 pg/m3 (1.0 ppm)
03-treated phospholipids
4 pM/min
Methyl ozonide
10-4-2xlO-3 M
750 nM/min
106 nM/min
300 nM/min
0-9.8 uM/g of Hb
0.84 pM/min
78400 pg/m3 (40 ppm)
Measurement
method
CHEM
NO
I
NO
NBKI
NBKI
NBKI
NBKI
NBKI
Exposure
duration and
protocol
0.5-2 hr
5, 10, 15, and
20 min
1 min
30 min
14.3 or 43.0
nMol of 03 per 106
cell equivalent
5, 10, 20, 30,
40 and 50 min
NO
0-2 hr
2 hr
Observed effect(s)
Decreased agglutination of RBCs by
concanavalin A.
Decreased ATPase 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
a1-proteinase inhibitor.
Decreased glyceraldehyde-3-PD.
Decreased ATPase.
No statistics.
Decreased GSH. No effect on
Hb or on glucose uptake.
Increased complement-mediated cell
damage.
Species
Human
Human
(RBC ghosts)
Human
(RBC ghosts)
Human
Human
Human
Human
(RBC ghosts)
Human
(RBCs,
RBC ghosts)
Human
Reference
Hamburger et al. , 1979
Kesner et al. , 1979
Kindya and Chan, 1976
Menzel et al. , 1975b
Koontz and Heath,
1979
Johnson, 1980
Freeman et al . ,
1979
Freeman and Mudd,
1981
Goldstein, B. , et al. ,
1974
        1,960  pg/m3  (1.0  ppm)      NBKI
20 and 60 min        Decreased native protein fluore-

                     scence.  No statistics.
Human        Goldstein et al.,

(RBC ghosts) 1975

-------
                                      TABLE 10-20.  HEMATOLOGY:  HUMAN - IN VITRO EXPOSURE  (continued)
Ozone3
concentration
40 nM/rain
t — i
0
i — »
en
-p>
2.5 uM/min
2.5 uM/min
L Exposure
Measurement duration and
method protocol
I 4 min
I 20, 40, and 60
rain
I 20, 40, and
60 min
Observed effect(s)
Decreased ATPase activity; lost 40%
membrane sulfhydryls. Lipid per-
oxidation and protein crosslinking
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 benzoic acid prevented
these 03 effects.
Species Reference
Human Chan et al., 1977
(RBC ghosts)
Human Verweij and
Steveninck, 1980
Human Verweij and Steveninck,
(RBC ghosts) 1981
bNot ranked by concentration; listed by reported values.
 Measurement method:   NO = not described; CHEM = gas phase chemiluminescence; NBKI = neutral buffered potassium iodide; I = iodometric.

-------
products  with semicarbazide  and PABA during 03  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  i_n vitro
reaction  of  03 with sulfhydryl  groups of human RBC membrane, proteins, and
cytoplasmic  contents.   After exposure of  RBCs  to  03 in  0^  at 20 ml/min  (0.84
nMol/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 [jg/m  (0.5 ppm)  of 03 continuously for 23 days.
     Although  jjn 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 i_n  vivo 03 toxicity in  man  is diffi-
cult.   In most J_n vitro studies,  RBCs were exposed by  bubbling high  0., concen-
trations  (>  1 ppm)  through  cell  suspensions.  Not only were the  0~  concentra-
                                                                 O
tions  unrealistic  and  the  method of exposure nonphysiological, but  the  toxic
species causing  RBC  injury may  be different  during i_n vitro and i_n  vivo  0.,
                                                                           O
exposures.   Because  of  its  reactivity,  it is  uncertain that 03 per  s_e 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.
10.4.3.3  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)
                                      o
investigated  the effects of 784  (jg/m  (0.4 ppm)  of 0., 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
10.3.1).
0190ZS/A                          10-155                               5/1/84

-------
     P'an and  Jegier  (1976) also reported changes  in  serum proteins after
                                3                         3
exposure of rabbits to  784  ug/m  (0.4 ppm) and 1960 |jg/m  (1.0 ppm) of Ov
                               3
Following exposure to 784 ug/m  (0.4 ppm) of 0» for 105  days,  the albumin
concentrations began to decrease,  and or and 6-globulin concentrations  began
to increase.  At the end of 210 days  of exposure, the mean albumin level fell
16 percent,  the a-globulin  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-
                                       3
tent  (8 hr/day)  exposure to  1568 ug/m  (0.8 ppm)  of  0~  for 7 days.   The
increased release of lysozyme  into the plasma was suggested  to  be a  result of
0- damage to alveolar macrophages.
                                         3
     After exposure of  rats  to 7840 ug/m  (4 ppm) of  0-  for 2, 4, and  8 hr,
Giri  et al.  (1980) observed increases in plasma prostaglandin (PG) F-  of 186,
109,  arid 25 percent, respectively.  Plasma concentrations of PGE» were approxi-
mately  twice  the  control  levels  after exposure  for  2  or 8  hr,  but  only a
slight  increase was observed after exposure for 4 hr.
     Veninga et al. (1981)  reported  that short-term exposures of mice to low
0~ concentrations induced changes in serum creatine phosphokinase (CPK)  activity.
 3
Ozone  doses  were expressed  as the product of concentration and time;  the
maximum 0-  concentration was  1600 ug/m   (0.8 ppm), and  the  maximum  exposure
          •3
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 0- doses employed.  The authors concluded that these
responses may  represent adaptation of the animals to 0_ toxicity by enhanced
metabolic processes.
10.4.3.4  Interspecies  Variations.   The  use  of animal  models  to investigate
the effects  of  0~ on the blood is  complicated,  because few species respond
                 O
like  humans.   The  rodent  model has  been  most commonly used to predict the
effects  of  0- on  human RBCs.   The  reliability of  this  model  was recently
challenged by Calabrese and Moore (1980) on the following grounds:   (1) ascorbic
acid  synthesis  was significantly increased   in  mice  following 0., exposure
(Veninga  and Lemstra, 1975), (2)  ascorbic acid protected  human  G-6-PD-deficient
RBCs  in vitro from the oxidant stress of acety1pheny1hydrazine  (Winterbourn,
0190ZS/A                          10-156                                5/1/84

-------
1979), and (3) humans lack the ability to synthesize ascorbic acid.  Although
Calabrese and Moore  (1980)  stressed that this hypothesis 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 model, because these mice
responded to 0, exposure (627 ng/m > 0.32 ppm for 6 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  Academy  of  Sciences, 1977).    However,  Dorset sheep were
found  to be  no  more sensitive than normal  humans  with  respect to 0.,-induced
changes in GSH and also differed from humans in the formation of methemoglobin.
Further studies (Calabrese  et al.,  1982, 1983b;  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 0, intermedi-
ates,  but  G-6-PD-deficient  human  erythrocytes were considerably more suscep-
tible.  Consequently,  the authors  also  questioned  the value  of the sheep
erythrocyte as a quantitatively accurate predictive model.

10.4.4  Reproductive and Teratogenic Effects
     Pregnant animals and developing fetuses may be at greater ris,k to effects
from photochemical  oxidants,  because  the volume of  air  inspired by females
generally increases from 15 to 50 percent during pregnancy (Allman 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 10.5), or they used oxidant concentra-
tions  greater  than  those typically found  in  ambient air.  Brinkman et al.
(1964) exposed pregnant mice to lower concentrations of 03, but the results of
their  experiments are difficult to  interpret, because the time of 03 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 pg/m  (0.1  ppm)  of 0~
(Table 10-21).   Unfortunately, without more details on the period of exposure,
0190ZS/A                          10-157                               5/1/84

-------
                                      PRELIMINARY DRAFT
                                                TABLE 10-21.   REPRODUCTIVE AND TERATOGENIC EFFECTS OF OZONE
o
en
Oo
Ozone
concentration
Mg/mJ
196
392
862
ppm
0.1
0.2
0.44
Measurement3
method
NO
ND
I
Exposure
duration and protocol
7 hr/day, 5 days/week
for 3 weeks
7 hr/day, 5 days/week
for 3 weeks
8 hr/day over entire
Observed effect(s)
Increased neonatal mortality £4.9 to 6.8%
vs. 1.6 to 1.9% for controls)0.
Unlimited growth of incisors (5.4% incidence
vs. 0.9% in controls) .
Decreased average maternal weight gain.
Species
Mouse
Mouse
Rat
Reference
Brinkman et al .
Veninga, 1967
Kavlock et al . ,

, 1964

1979
     2920
1.49
period of organogenesis
(days 6 to 15)

Continuous during mid-
gestation
                                                                   Increased fetal  resorption rate (50% vs.  9%
                                                                   for controls).
     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  =  iodometric  (Saltzman  and  Gilbert,  1959).
     No  statistical evaluation.

-------
it is  impossible  to  ascertain whether the decreased infant survival  rate was
due to development  interference  ij} 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,
                                                  3
Veninga (1967) found that mice exposed to 392 ug/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
    O
effects were  found  in rats  exposed 8 hr/day to concentrations  of 03 varying
from 863 to 3861  ug/m3  (0.44  to 1.97 ppm)  during  early  (days  6  to  9),  mid
(days 9 to  12),  of late  (days 17  to 20) gestation, or the entire period of
                                                                                3
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.
                                                                  3
A single dose of  150  mg/kg sodium  salicylate followed by 1960 ug/m   (1.0 ppm)
of OT during midterm produced a significant synergistic increase  in the resorp-
tion rate,  a  decrease in maternal weight  change, and  an average fetal weight.
                                               3
Exposure of pregnant  rats 8  hr/day to  862  ug/m  (0.44 ppm)  of  0.. 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 ug/m   (1.5 ppm)  of  0, for  periods  of 4 days
                                                                      3
during gestation.   Pups  from  litters of females exposed  to  1960  ug/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
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 ug/m  (1.5 ppm)  of 0~  (1.0 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.
0190ZS/A                          10-159                               5/1/84

-------
10.4.5  Chromosomal  and Mutational  Effects
10.4.5.1  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  |jg/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 conduc-
                          3
ted at or below 1,960  ug/m  (1 ppm) of 0~.   Data derived predominantly from j_n
                                                                         3
vitro experiments conducted  at  0,  concentrations in excess of 1,960 ug/m  (1
                                 o
ppm) of  0^ will be discussed  first  (Table 10-22), followed by  a  discussion of
the genotoxicity data  from both i_n  vitro and  in v"ivo research  conducted at. or
below I ppm of 03 (Table 10-23).
     The potential  for genotoxic effects relating to 0~ exposure was predicted
from the radiomimetic  properties  of GO-  The decomposition  of 0^ in water
produces OH and  H0« 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
human KB  cells  exposed to 15,680 ug/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, slightly more efficient  in the production  of
chromatid breaks than  50 rad of x-rays.  Significant mitotic delay was measured
in neuroblasts  from the  grasshopper Chortophaga  viridifaciata  exposed  to  3500
to 4500 ug/L of 0~ 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 when
cells were exposed to  1 ug/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 03-
     The molecular mechanism for the  clastogenic  and lethal  effects  resulting
from 0~  exposure  are  not precisely known.    Bubbling 8 percent 0., through a
phosphate buffer  solution (0.05M,  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 pyrim-
idine bases,  nucleosides,  and nucleotides.    In a more recent  report, Shiniki
et al. (1981)  examined the degradation of a mixture of 5' nucleotides, yeast
0190ZS/A                          10-160                               5/1/84

-------
                              TABLE 10-22.  CHROMOSOMAL EFFECTS FROM IN VITRO EXPOSURE TO HIGH OZONE CONCENTRATIONS
    03            Measurement
concentration         method
                                                   Exposure
                                              duration and protocol
                              Observed effect(s)
                                            Species
                                                                                   Reference
15,680 pg/m3 (8ppm)
98,000 |jg/m3 (50 ppm)
UKI 5-25 min Chromatid deletions.
MAST 30 min lex mutants deficient in
repair of x-ray- induced
Humans KB
cells
Escherichia
col i
Fetner, 1962
Hamelin and
Chung, 1974
                                                                      DNA strand breaks were more
                                                                      sensitive to lethal  effects
                                                                      of 03 than were the  wild-type
                                                                      repair-proficient parental  strains
      98,000  pg/m3  (50  ppm)     MASTQ
30 min
o
 i
CT>
DNA Polymerase I mutant strains
(KMBL 1787, 1789, 1791) were more
sensitive to the cytotoxic effects
of 03, and DNA 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  ug/m3  (50  ppm) !    MAST
30 min
Mucoid mutapt 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 03 and
degraded DNA to a greater extent
than the Ion + (MQ259) strain.
                                                                                                     E. coli
Hamelin et al. ,
1977b
      98,000  pg/m3  (50 ppm)     MAST0
up to 3 hrs         15 different DNA repair-deficient
                    strains were tested for sensitivity
                    to the cytotoxic effects of 03; DNA
                    Polymerase 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 10-22.   CHROMOSOMAL EFFECTS FROM IN VITRO EXPOSURE TO HIGH OZONE CONCENTRATIONS  (continued)
03 Measurement
concentration method
0.1 ug/ml [51 ppm] UV
0.5 ug/ml [255 ppm] UV
0.18 ug/ml [92 ppm] UKI
1.0 ug/ml [510 ppm]
0.5-6 ug/ml [255- NBKI
3061 ppm]
Si- 10 ug/ml [510- NBKI
^ 5100 ppm]
CTi
(V)
5% [50,000 ppm] NBKI
3.5-4.5 ug/ml [1786- UKI
2296 ppm]
2% [20,000 ppm] e
8% [80,000 ppm] GPT
Exposure
duration and protocol Observed effect(s) Species Reference
60 min (70 ml/min) Preferential degradation of yeast Yeast Shiniki et al.,
RNA at the N-glycosyl linkage; 1981
sugar-phosphate linkage was
03 stable. ;
30 min (330 ml/min) 5-ribonucleotide guanosine
monophosphate was degraded
most rapidly.
ND Release of nucleic acids; E. coli Scott and Lesher,
cell lethality. • 1963
0-5 min 03 reacts with pyrimidine E. coli Prat et al . , 1968
bases from nucleic acids
(thymidine > cytosine >
uracil ).
30 min Cell death and nonspecffic Chick , Sachsenmaier
chromosomal aberrations: embryo et al., 1965
shrunken and fragmented nuclei, fibroblasts
clumped metaphase chromosomes
and chromosome bridges.
0-40 min Rapid loss of glycolytic and Mouse !
respiratory capacity; loss of ascites
tumorigenicity after 20 min. cells
exposure.
ND Mitotic delay. Chortophaga Fetner, 1963
viridifaciata
3 min Abnormal nuclei; fragmentation. Am. oyster MacLean et al.,
1973
5, 15, 60 s Rapid degradation of nucleic ND Christensen and
acid bases, nucleosides or Giese, 1954
nuc"?eotides in 0.05M phosphate
buffer, pH 7.2.
aNot 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 flow rate given in (ml/min), when available.   ND  =  not described.
 Concentrations of 03 were not measured in the  cell suspensions.
e03 analyzer (Fisher and Porter, Warminster,  PA).

-------
3920     2.0
        TABLE 10-23.   CHROMOSOMAL EFFECTS FROM OZONE CONCENTRATIONS AT OR BELOW 1960 pg/m3 (1 ppm)
Concentration Measurement3 >b
Mg/mJ
294
412
1940
451



2548-
14,700


3234-
27,832
ppm method
0.15 NBKI
0.21
0.99
0.23 NBKI



1.3- NBKI
7.5


1.65- NBKI
14.2
Exposure0
duration and protocol
5 hr
5 hr
2 hr
5 hr
(in vitro)


ND
(in vitro)


NO
(in vitro)
Observed effect(s)
No effect induced by 03 treatment
on the frequency of chromosome or
chromatid aberrations in Chinese
hamster or mouse peripheral
blood lymphocytes stimulated with
PHA; no effect on spermatocytes
in mice 8 wk following exposure.
Peripheral blood lymphocytes exposed
to 03 in culture 12 hr after stim-
ulation with PKA showed na increase
in chromosome or chromatid aberrations.
Peripheral blood lymphocytes exposed
to 03 in culture, 36 hr after PHA
Species Reference
Mouse Gooch et al . ,
1976

Hamster


'
Human
lymphocytes


Human
lymphocytes
NBKI
                      5-90 min
                     (in vitro)
                                                                       stimulation,  showed no change  in the
                                                                       frequency of  chromosome or  chromatid
                                                                       aberrations at any concentration
                                                                       except 7.23 ppm of 03.

                                                                       No apparent increase in the fre-
                                                                       quency of chromosome or chromatid
                                                                       aberrations 12 or 36 hr after
                                                                       PHA stimulation.
                                                                                                                     Human
                                                                                                                     lymphocytes
392      0.2
MAST
UKI
                                                   5  hr
                                                 (in  vivo)
                                                                       Combined  exposure  to  03  and  radi-
                                                                       ation (227-233  rad) produced an
                                                                       additive  effect on the number
                                                                       of  chromosome breaks  measured in
                                                                       peripheral  blood lymphocytes.
                                                                                                                     Hamster    Zelac et al.
                                                                                                                                1971b
470-     0.24-
588      0.3
MAST
UKI
                                                   5  hr
                                                 (in  vivo)
A significant increase in chro-
mosome aberrations (deletions,
ring dicentrics) in the peri-
pheral blood lymphocytes; in-
creased break frequency was
still  apparent at 6 and 16 days
following exposure.
                                                                                                                     Hamster    Zelac et al .
                                                                                                                                1971a

-------
                           TABLE 10-23.   CHROMOSOMAL EFFECTS  FROM OZONE  CONCENTRATIONS AT OR BELOW 1960 pg/m3 (1 ppra)  (continued)
o
I—«
en
03
Concentration,
ug/m3 ppm
490- 0.25-
1960 1.0
588- 0.3-
1568 0.8
843 0.43
3920 2.0
1960- 1.0-
9800 5.0
Measurement * Fxpncnrp
_. p i —
method duration and protocol
UV 1 hr
(in vitro)

UV 8 days,
continuous
(in vitro)

UV, 5 hr
NBKI (j_n vivo)
6 hr
(in vivo)
CHEM 24 hr
NBKI (in vivo)

Observed effect(s)d Species Reference
Dose-related increase in SCE fre- Human Guerrero et
quency in WI-38 diploid fibroblasts fibroblasts al., 1979
exposed in culture.
Growth of cells from lung, breast, Human ! Sweet et al.,
and uterine tumors were inhibited tumor 1980
to a greater degree than IMR-90, cells
a nontumor diploid fibroblast.
Increase in chromatid-type Hamster Tice et al.,
aberrations in peripheral blood 1073
lymphocytes of 03-exposed hamsters;
Increase in deletions at 7 days
and increase in achromatic lesions
at 14 days after exposure; chromosome- !
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:  MAST =  Kl-coulometric  (Mast meter); NBKI = neutral buffered potassium  iodide; UV = UV photometry
     Calibration method:  UKI = unbuffered  potassium iodide; NBKI = neutral buffered potassium iodide.
     ND = not described.
     Abbreviations used:  PHA = phytohernagglutinin; SCE = sister chromatid exchange.

-------
t-RNA  or  tobacco mosaic virus  RNA with 0., (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 0_  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 extensive
ozonolysis of yeast t-RNA  (0.5 ug/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 0_-labile,
the sugar-phosphate backbone was intact and extremely stable against 0^.  Prat
et al. (1968)  investigated  the reactivity of  the  pyrimidines in £_._ coli DNA
with  0~  (0.5  to 6 mg/L, 0  to 5 min) and  radiation.   Ozone  preferentially
reacted with  thymidine,  then  with  cytosirie and uracil,  in decreasing order of
reactivity.  The results are slightly different from those reported by Shiniki
et al. (1981)  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_.   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  significa'ntly more sensitive to the cytotoxic effects of
OT 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  coworkers  investigated  the  survival  of bacterial strains  with
known  defects  in  DNA  repair.   Closely  related strains  of E^ coli  K-12 with
mutations in DNA polymerass I were shown to be more sensitive to the cytotoxic
effects of 03  than  the DNA polymerase  proficient  (pol  +) strain (Hamelin et
al. ,   1977a).   Polymerase I-deficient  strains also  exhibited  an extensive
degradation of DNA in  response to 0., 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
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

0190ZR/A                          10-165                                5/1/84

-------
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 thsy were exposed to 0,.
     Although DNA  polymerase  I was  shown  to  be involved in  the  repair  of
O^-induced DNA  damage  (Hamelin  et  al. ,  1977a),  £_._ coli  cell  strains with
mutations in  DNA  polymerase  II  or III were not found to be more sensitive to
CL 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 E^
coli with defects  in DNA synthesis (DNA A, B, C, D, and G) showed  no  enhanced
sensitivity to 0~.  Therefore, the DNA gene products are probably not involved
in the  repair of  0~ damage.   Recombinational  repair mutants,  rec A and rec B,
only showed a slightly increased sensitivity to CL  than the wild-type, suggest-
ing that the  rec  gene  products are  only partially involved in  the repair of
(k-induced DNA lesions (Hanelin and Chung,  1978).
     Other effects have been  observed than those described above on bacteria.
In the  commercial  American oyster exposed to  CL-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 CL (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  i_n  vi tro  effects of extremely high 0., concentrations on
biolog4cal systems or  biologically  important  cellular components.  Although
these investigations  may  be important for the  elucidation  of the types  of
damage or responses that might be expected to occur at lower 0_ concentrations,
the most relevant data or.  the  genotoxicity of  Oq should  be  obtained from
                                                                   3
investigations where the  0.,  concentration did not exceed 1960 ng/m   (1 ppm).
Research conducted at or below 1 ppm of 0,. will  be presented  below (See  Table
10-24).
     Several  investigators have examined the  i_n vivo cytogenetic effects of 0,,
in rodents and  human  subjects.   Until the reports of Zelac et al. (1971a,b),

0190ZR/A                          10-166                                5/1/84

-------
                                   PRELIMINARY  DRAFT
                                                        TABLE  10-24.   MUTATIONAL EFFECTS OF OZONE
°3 a
concentration Measurement Exposure
ug/m-* ppm method duration and protocol Observed effect(s)
196 0.1 MAST 60 min Various mutated, growth factor auto-
(2.1 ml/min) trophic states, were recovered; mutant
strains differed from parental strains
in sensitivity to UV light and excessive
production of capsular polysaccharide.
58,800 30 NBKI 3 hr Induction of a dominant lethal muta-
tion during stages of spermatogenesis;
sperm were found to be twice as sen-
sitive as earlier stages.


-------
the toxic effects  of  CL  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  0,,  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 (jg/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 0,-,-exposed animals
when compared with nonexposed control animals.  Zelac  et al.  (1971b) reported
that 0- was additive  with  radiation in the production of chromosome breaks.
Both 03 and radiation  produced chromosome breaks  independently  of each other.
Simultaneous exposure  to 3.2 ppm of CL  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 0~ 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 pg/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 (jg/m   (5 ppm)  of 0- 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 0--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 0^
treatment,  and the induction of alkaline-labile  lesions  in DNA  by  0,, which
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 0, effect on DNA at 9800 pg/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

0190ZR/A                          10-168                                5/1/84

-------
inhalation, Tice  et  al.  (1978)  tried to repeat the  experiments  of Zelac as
                                                               3
closely as possible.   Chinese hamsters were exposed to 843 jjg/m  (0.43 ppm) of
0~ by  inhalation  for 5  hr.   The  authors  investigated  chromatid and  chromosome
aberrations in  peripheral  blood  lymphocytes and bone marrow  of  control  and
(L-exposed animals immediately  after exposure and at 7  and 14 days after  0.,
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 0- for 6 hr were examined
in peripheral  blood  cells collected  from  the animals immediately  after 0~
exposure and at 7 and 14 days after 0- exposure.
     The authors  reported no significant increase  in  the SCE  frequency of  the
0_-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  0,-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 less 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 |jg/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
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 0~.
          Zelac et al. (1971a) administered 470 to 590 (jg/m3 (0.24 to  0.3
          ppm) of 03  to  Chinese hamsters for 5 hr, whereas Tice et al.
          (1976
          of 0,
(1978) exposed animals to an atmosphere of 840 pg/rn  (0.43 ppm)
              3'
0190ZR/A                          10-169                                5/1/84

-------
     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 i_n 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 0,
                                                                            O
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 v/ith the results with the use of polkweed mitogen.
     There  is evidence  thet  the B-lymphocyte may be  more  sensitive  to 0~ than
the T-lymphocyte.  Savino et al.  (1978)  measured the effects of CL on  human
cellular and humoral  immunity by measuring rosette formation with human lympho-
cytes (See  Chapter 11).  Rosette formation measures  the reaction of antigenic
red cells  with  surface membrane sites on  lymphocytes.   Different antigenic
RBCs are used to distinguish T-lymphocytes from B-lymphocytes.   Rosette forma-
tion with  B-lymphocytes  was  significantly depressed in eight human subjects
exposed to 784 ug/m  (0.4 ppm) of 0- by inhalation for 4 hr.  A similar inhibi-
tion of rosette  formation  was not observed with T-lymphocytes  from the same
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 0.,
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 &1.  (1978).  There are sufficient differences  in the
experimental protocols of the two reports so that the results need not be con-
sidered directly contradictory.

0190ZR/A                          10-170                                5/1/84

-------
     An assumption that is made in all of the reports in which lymphocytes are
stimulated with mitogens  is  that the lymphocytes from the O^-exposed animals
and the control animals are 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  11.7).  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 GQ  or G,  cells into the cycling state, cells  exposed  to sufficient concen-
trations of  0^ would  not be stimulated  to divide, and hence  no 0.,-induced cy-
togenic effects would  be  observed in activated cells.   In their report, Tice
et al.  (1978)  stated  that the  lymphocytes of the 03~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 0,-damaged lymphocytes were  selected
against in the cytogenetic investigations.
     In other  investigations  with  rodents,  Gooch et al.  (1976) analyzed bone
                                                        3
marrow samples from Chines? 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 C-H mice  were  exposed to 294 or 412  ug/m   (0.15
                                              o
or 0.21  ppm) of  03 for 5  hr,  or to  1940 ug/m   (0.99 ppm) of  03  for 2 hr.
Blood samples were drawn from these animals at various times for up to 2 weeks
following 03 exposure.   The  mice were  killed  8  weeks following 0,  exposure,
and spermatocyte preparations were made and analyzed for reciprocal transloca-
tions.   Data from the Chinese hamster bone marrow samples and the mouse leuko-
cytes  indicated that  there was no effect induced by 03 treatment on the fre-
quency of chromatid or chromosome aberrations, nor were there any recognizable
reciprocal  translocations in the primary spermatocytes.
     Several  investigators have examined  the effects of 0,, on htiman cells i_n
vitro.   Fetner (1962)  observed  the  induction of chromatid deletions in human
KB cells exposed to 15,68C ug/m (8 ppm) of  03 for  5 to  25 min.   Sweet  et al.
(1980) reported that the growth of human cells from breast,  lung,  and uterine
                                                     3
tumors was inhibited  by  exposure to 588 to 1568 ug/m  (0.3 to 0.8 ppm) of 03
for 8 days  in culture.

0190ZR/A                          10-171                                5/1/84

-------
     Guerrero et al.  (1979) performed SCE analysis  on diploid human fetal  lung
cells (WI-38) exposed to 0, 490,  980, 1470,  or 1960 ug/m3 (0, 0.25, 0.5,  0.75,
or 1.0 ppm)  0.,  for  1 hour in vitro.   A dose-related increase in the SCE  fre-
              ^
quency was observed  in the WI-38  human fibroblasts  exposed to 0^.   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  0_ by inhalation  (Chapter 11).  Unless the lymphocyte is  intrinsi-
cally less  sensitive  to the  induction of SCE  by 0,  than the WI-38  human fetal
lung fibroblast, the  results indicate that exposure of  human subjects to 980
ug/m  (0.5  ppm)  of  0_ for 2  hr  did  npt result in  a sufficiently  high con-
centration  of 03 or  03 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 ug/m  (0.25 ppm) for 1 hour.
     Gooch et al. (1976) also investigated the effects of 0^ exposure on  human
cells HI  vitro.  In  these experiments, lymphocytes  were  stimulated with  PHA
for  12  or  36 hr before the 0~ exposure to obviate the  potential  problems of
OT inactivation of  the PHA receptor.   Human leukocyte cultures were exposed to
          3
3920 M9/m   (2  ppm)  of 07  for  various  lengths  of time to  accumulate total 0.,
                        •3             o                                    ^
exposure  doses  of 3234 to 27,832 ug/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
aberrations at a total exposure of 14,170 ug/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
                         T   j
on cells  and cellular components including  the genetic  material.   Cytogenetic
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.
10.4.5.2  Mutational  Effects of  Ozone.   The  mutagenic effects of 0., have  been
investigated  in  surprisingly few instances  (Table  10-24).   No  publication to
date has  investigated the  mutagenic effects  of 0, in mammalian  cells.
0190ZR/A                          10-172                                5/1/84

-------
     Sparrow and Schairn  (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 CL concentration was reported in this publics-
                                 O
tion.
     E^ coli, strain MQ 259, were mutated to various growth factor auxotrophic
states,  including  requirenents for most  common  amino acids, vitamins, and
                                                                           3
purines and pyrimidines  (Hamelin  and  Chung, 1975a).   Ozonated air 196 ug/m ,
0.1 ppm) was  passed through the bacterial suspensions at a rate of 2.1 L/min
for 30 min.   Many  of the 0 - 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
                                                       3
mutations in  Drosophila virilis  exposed to 58,800 ug/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 Saccharomyces cerevisae.   Several  different strains
                   *j
were utilized  to investigate forward, reverse, and recombinational mutations.
ozone  (98,000  ng/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 (jQ/m » 50 ppm;  30 to 90 min) was compared with other known mutagens.
 13                                             2
The positive  controls  were  UV  light (1.54 J/m  per second, 1-tnin  exposure),
N-methyl-N1-nitrosoguanidine (MNNG) (50 ug/mL, 15 min), or x-rays  (2 kR/ min,
40  min).   By comparing the  induced mutation frequency at similar cellular
survival  levels for 03, MNNG, UV light,  and x-rays, it was shown that 03  was a
very weak  mutagen.   Induced mutation frequencies  were generally 20 to 200
times lower for 0  than for the other  three mutagens.
0190ZR/A                          10-173                                5/1/84

-------
     In summary,  the  mutagenic properties of 0.,  have  been demonstrated in
procaryotic and  eucaryotic  cells.  Only  one study, however,  (Hamelin  and
Chung, 1975a,  with E^ coli)  investigated the  mutagenic  effect of CL at concen-
trations of less  than  1  pom.   The results clearly indicate that if cells  in
culture are exposed to sufficiently high concentrations of (L 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  CL toward these  species  can be determined.

10.4.6    Other Extrapulmonary Effects
10.4.6.1  Liver.   A series of  studies reviewed  by Graham et al. (1983a) have
shown that 0.,  increases  drug-induced  sleeping time in  animals (Table 10-25).
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 sleeping time, because liver xenobiotic  metabolism transforms the drug
into an inactive form.   Therefore, this response is interpreted as an extrapul-
monary effect.
     Gardner et al.  (1974) were the first to observe that 0., increases pentobar-
bital-induced sleeping time.   Female  CD-I mice  were exposed for 3 hr/day for
                          2
up to  7  days  to 1960 ug/m  (1.0  ppm)  of CU and  the increase  was  found on
days 2 and 3 of exposure, with the greatest response occurring on day 2.  Com-
plete tolerance did not occur; when mice were pre-exposed  (1960 ug/m  , 1.0 ppm;
                                                                           2
3 hr/day for  7  days) and then challenged with  a  3-hr  exposure  to  9800  [jg/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  ex,  al.,  1981).
                 3
Ozone  (1960 ug/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  ug/m3 (1.0 ppm) of 03 (Graham et al.,  1981).    The females of all

0190ZR/A                          10-174                                5/1/84

-------
                                               TABLE 10-25.  EFFECTS OF OZONE ON THE LIVER
   Ozone
concentration
ug/m3
1960
ppm
          Measurement3'b
             method
     Exposure
duration and protocol
Observed effects(s)
Species     Reference
196- 0.1- CHEM 3 hr/day,
9800 5 GPT 1-17/days
588 0.3 UV 3 hr
1470 0.75 NBKI 3 hr
i-- 5880 3


-------
                                            TABLE 10-25.   EFFECTS OF OZONE ON THE LIVER  (continued)
          Ozone
       concentration  Measurement
       pgTm3"    ppm      method
       a,b
                      Exposure
                 duration and protocol
                                Observed effects(s)
                                            Species
                Reference
1960
1960
1 CHEM
GPT
1 CHEM
GPT
5 hr/day,
1,2,3 or
4 days
5 hr
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 hexobarital-and thiopen-
tal-induced sleeping time and zoxazo-
Mouse
(female)
Mouse
(female)
Graham et al . ,
1981
Graham et al . ,
1982a
CTl
                                             1 amine-induced paralysis time.  Pre-
                                             treatment with mixed function oxidase
                                             ip.d'jcers (phenobarbetal, pregr.eolor.e
                                             -16a-carbonitrile, and B-naphthofla-
                                             vone, but not pentobarbital) decreased
                                             phenobarbital-induced sleeping time in
                                             CD-I mice, and 03 increased the sleeping
                                             time in all groups.   Pretreatment 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
       9800
CHEM
GPT
5 hr
3 hr
No effect on hepatic cytochrome
P-450 concentration, aminopyrine
M-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.
  Mouse
(female)
Graham et al.
1982b

-------
                                   PRELIMINARY  DRAFT
                                                TABLE  10-25.   EFFECTS  OF  OZONE  ON  THE  LIVER   (continued)
   Ozone
concentration
ug/m3     ppm
Measurement
    method
           a,b
     Exposure
duration and protocol
        Observed effects(s)
Species  Reference
1960
9800
    CHEM
    GPT
    5 hr
At 1 ppm:   71% increase in plasma
half-life of pentobarbital,  decrease
(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
         brain
  Mouse  Graham 1979;
(female) Graham et al.
         1983a,b
1960
    NO
                                                   90  inin
                            No effect on hepatic cytochrome
                            P-450 concentration.
                                            Rabbit   Goldstein
                                                    ! et al., 1975
3920
    UV
    8 hr/day
Supplementing or depriving rats of
vitamin E or selenium altered the
03 effect.  03 caused changes in
several iji vitro enzyme activities
in the liver and kidney (see text).
                                                                                             Rat
         Reddy et al.
         1983
 Measurement method:   CHEM = gas-phase chemiluminescence;  NBKI  = neutral  buffered potassium iodide;  UV = UV photometry;  NO = not described
 Calibration method:   GPT = gas phase titration

-------
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.
                                                 3
Female CD-I mice were exposed to 196 to 9800 ug/m   (0.1 to 5.0 ppm) of 0,  for
                                                                           3
3 hr/day for a  varying  number of days (Graham et al. , 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 ug/m   (0.5 ppm),  the greatest
increase in pentobarbital-induced sleeping time was observed on day 3, but at
        3
490 ug/m  (0.25  ppm),  6 days  of exposure were required to cause an increase.
At the lowest  concentration evaluated (196 ug/m3, 0.1 ppm),  the increase was
only observed at days  15 and 16 of exposure.   Thus,  as  the concentration of 0~
                                                                             
-------
and metabolism.  Although there are different mechanisms involved in hexobarbi-
tal,   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 CL to differences in the actions pf 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 elucida-
ted.
     The effect  of  0., 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  0~  as  high as 9800  ug/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-
lase).  However,  this exposure  regimen  and another  (1960  ug/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
3-hr  exposure  to  9800 ug/rn  (5.0 ppm)  of  0-  did not affect brain concentra-
tions  of pentobarbital  at time  of awakening, even  though  sleeping time was
increased (Graham, 1979).   Therefore,  it appears that  0~  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, 1979).   Following this
exposure, first-order clearance kinetics of  pentobarbital were observed in
both  the air and  03  groups,  and 03 increased  t^e plasma half-life by 106  per-
cent  (Graham,  1979).  Mice exposed to 1960  ug/m3 (1.0 ppm) of  0.,  for 5  hr had
                                                                »3
a 71 percent increase in the plasma half-life of pentobarbital  (Graham et al. ,
1983b).  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.
0190ZR/A                          10-179                                5/1/84

-------
     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., 1983a) 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 10.2).  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  10.3.3).
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 0_ on liver xenobiotic metabolism by perform-
ing HI vitro  enzyme assays.   Although lower Og  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
cytochrome 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 0~  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  activity in 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
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

0190ZR/A                          10-180                                5/1/84

-------
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 0- 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  we"e 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  0,  and
examined at 5 postexposure periods up to 24 hr.   Rats had significantly increased
ascorbic  acid levels  in  both the Q~ and air groups, with the greater change
taking place in the. air group.  Thus, there were no changes due to 0.,.   Likewise,
there was no 0.. effect on reduced ascorbic acid content in the serum.
10.4.6.2  The Endocrine System.  A summary  of  the effects of 0^ on the endo-
crine system,  gastrointestinal  tract,  and  urine is given  in  Table 10-26.
Fairchild  and co-workers  were the  first to  observe  the  involvement of  the
                                                               3
endocrine system in 0~ toxicology.  Mice exposed 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  Q- 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  0.,.   Thyroid-blocking agents  and thyroidectomy  increased the
survival of mice and rats acutely exposed to 0~,  and injections of the thyroid
hormones,  thyroxine  (T.),  or  triodothyronine (T,) decreased their survival.
This response in animals  with altered thyroid function was not specific to a

0190ZR/A                          10-181                                5/1/84

-------
                                     TABLE 10-26.   EFFECTS OF  OZONE  ON  THE  ENDOCRINE  SYSTEM,  GASTROINTESTINAL  TRACT,  AND  URINE
Ozone
concentration Measurement
ug/m3
5.4
21
110
490
2940
980
1960
ppm method
0.003 c
0.01
0.056
0.25 d
1.5
0.5 I
1
Exposure
duration and
protocol
93 days,
continuous
2 hr
30 min
5 hr/day,
4 days
Observed effect(s) Species Reference
From 6th wk to end of exposure, 0.056 ppm Rat Eglite, 1968
increased the urine concentration
of 17-ketosteroids. After 93 days of
exposure to 0.056 ppm, the ascorbic and
level of the adrenal glands was decreased.
No data were presented for these effects.
1.5 ppm of 03 (30 min) inhibited gastric Rat Roth and Tansy,
mortality; recovery was rapid. The 1972.
lower level caused no effects.
No effects on thyroid release of 131I, Rat Fairchild et al.,
96-384 hr post lMI injection. 1964.
o
 I
00
ro
1470
0.75
ND
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
              ND         48 hr          Early postexposure  parathyroid glands
                         postexposure       enlarged and congested with focal
                         1-20 days          vasculitis.   After 7 days postexposure,
                                            parenchyma!  atrophy, leukocyte infiltra-
                                            tion and capillary proliferation.   Authors
                                            suggest lesions may be due to autoimmune
                                            reactions.
                                                                                                               Rabbit
                                                                                              Atwal  et al.,  1975
        1470
         0.75
            ND
           48 hr              Microvascular changes in the parathyroid
           postexposure       glands, including hemorrhage, endothelial
           12-18 days         proliferation, platelet aggregation, and
                              lymphocyte infiltration.
                                                                     Dog
                                                                Atwal and Pemsingh, 1981
        1568-    0.8-
        2940     1.5
        (range)  (range)
                     NBKI
                         6 hr/day,       Lower titratable acidity of urine,
                         4 days/wk       with no  changes  in  levels of creatine,
                         about 19 wk    uric acid/creatinine,  amino acid
                                        nitrogen/creatinine,  or excretion of
                                        12 ami no acids.
                                                                                Rat
                                                                                1962
                                                                                   Hathaway and Terrill,

-------
                          TABLE 10-26.   EFFECTS  OF  OZONE ON THE ENDOCRINE SYSTEM, GASTROINTESTINAL TRACT, AND URINE  (continued)
o
 i
00
GO
Ozone
concentration
ug/m3
1960
3920
7840
ppm
1
2
4
Exposure
Measurement duration and .
method protocol Observed effect(s) Species Reference
I 5 hr Decreased release of 131I from thyroid, Rat Fairchild et al.,
48-384 hr post 131Injection to all 03 1964.
levels above 1 ppm.
         1960
   ND
24 hr
Decreased serum level of thyroid-
stimulating hormone from anterior
pituitary, thyroid hormones (T3, 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 time up to a
potential maximum of 500 hr also caused
a decrease in TSH levels.
                                                                                                       Rat
demons and
Garcia, 1980a,b.
79,800 > 5.0 I
> 3 hr
< 8 hr
Anti-thyroid agents, thyroidectomy,
hypophysectomy, and adrenal ectomy
protected against 03- induced mor-
tality. Injection of thyroid hormones
decreased survival after 03 exposure.
Mouse,
rat
Fairchild et al. ,
1959; Fairchild and
Graham, 1963;
Fairchild, 1963.
         9800
CHEM
                                            3 hr
               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,  1956);  ND  = not  described.

          Abbreviations  used:   LM =  light microscopy; TEM = transmission electron microscopy; RER = rough endoplasmic reticulunr
          T3 =  trnodothyronine; T4  =  thyroxine.
          Spectrophotometric technique (dihydroacridine).
          Flow rates  from ozonator.

-------
concomitant altered  metabolic rate,  because  another drug  (dinitrophenol),
which increases metabolism, had no effect on the (k response.
     Hypophysectomy and adrenalectomy  also  protected against 0,-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 adrenocorti costeroi ds reduces 0.,
toxicity.  Rats  that  have  been adrenalectomized and treated with  adrenergic
blocking agents are more resistant to 0~ than rats that have only been adrena-
lectomized, indicating  that  decreases  in catecholamines reduce  0~ toxicity.
     Potential tolerance  to  the  effect of CL 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 ug/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  0,.  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  07  (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  |jg/m   (4.0  ppm) on the third day.  These animals exhibited a greater
                                                  3
effect than rats that  received only the 7840-|jg/m  (4.0  ppm) challenge.   This
difference persisted for 43, 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 0~ challenge.
     demons and Garcia (2.980a,b) extended  this area of research by investi-
gating the effects  of  0.,   on  the  hypothalamo-pitui tary-thyroi d axis  of rats.
Generally,  these three  endocrine  organs regulate the function of  each other
through complex  feedback  mechanisms.  Either stimulation or inhibition of the
hypothalamus  regulates  the release of  thyrotropin-rel easing hormone (TRH).
The thyroid hormones  (T~   and T.)  can  stimulate TRH.  Thyrotropin-releasing

0190ZR/A                          10-184                                5/1/84

-------
hormone and circulating thyroid  hormones  (T., 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 ng/m   (1.0
ppm) of 0~  caused  decreases  in the serum concentrations of TSH, T,, T.,  free
T,, and protein-bound  iodine.   There  was no change in  the uptake of T~, 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
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 exoosure 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 I960, 3920, 5880 or 7840 ug/m3  (1.0, 2.0, 3.0, or 4.0 ppm) of 03 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  of  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.
     The thyroid gland itself  was  altered by the 0.. exposure (apparently 1960
    3
Ug/m ,  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

0190ZR/A                          10-185                                5/1/84

-------
tissue.   There was  no  change  in T. release per  gland.   demons and Garcia
(I980a,b) interpreted these findings  as  an (L-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.
     The susceptibility of the  parathyroid gland to CL 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  0_ exposure.   Changes  included  hyperplasia  of chief cells;
hypertrophy and proliferation of the  rough endoplasmic reticulum,  free  ribo-
somes, mitochondria,  Golgl  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.
     Atwal  et al.  (1975) also investigated possible autoimmune involvement in
                                                                  3
parathyroiditis  of rabbits following a 48-hr exposure to  1470  M9/m   (0.75 ppm)
of 03-   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  e'ectron  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.   The microvascular  changes were further
studied  by Atwal  and Pernsingh (1981)  in dogs exposed to 1470 jjg/m3 (0.75 ppm)
of 0, for 48  hr.   They  reported focal hemorrhages, vascular endothelial  pro-
    «J
liferation,  intravascular platelet aggregation, and lymphocytic infiltration.
A potential autoimmunity after 0_ exposure was also observed by Scheel et al.
(1959), who showed the presence of circulating antibodies  against lung tissue.

0190ZR/A                          10-186                                 5/1/84

-------
     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
ug/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 live1" or kidney.
     Adrenal  cortex  function  after a 93-day (continuous) exposure to  0., was
                                                           3
investigated  in  rats  (Eglite,  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
 O
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.
10.4.6.3  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 relatively 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 0,,-exposed animals.   Titratable acidity was also lower on day 98.
              «J
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  differ-
ences were  observed.  The  authors interpreted the results  as a reflection of
respiratory alkalosis,  assuming no kidney toxicity.
     Gastric secreto-motor activities of the rat were investigated by Roth and
Tansy (1972).  A 2-hr  exposure to 490 pg/m   (0.25  ppm)  caused no effects.
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.
0190ZR/A                          10-187                                5/1/84

-------
10.5  EFFECTS OF OTHER PHOTOCHEMICAL OXIDANTS
10.5.1  Peroxyacetyl  Nitrate
     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  (LC5[)) (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. , 1977).  The concentration of
                                                3                      3
PAN used in  these  studies ranged from 22.3 mg/m  (4.5 ppm) to 750 mg/m  (150
ppm).   They  are  considerably  higher than the 0.174  mg/m   (0.037 ppm) daily
maximurr. concentration of  PAN reported for ambient air samples  in  areas  having
relatively high  oxidant  levels  (Chapter 6)  and are of questionable relevance
to the assessment of effects on human health.
     Campbell et al.  (1967) estimated that the LC,-n for  mice ranged from 500
           Q                                      OU
to 750 mg/m  (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 LC™ (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.   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 depres-
sed both the 6-hr  and 24-hr activity when  compared  to  similar pre-exposure
periods.  Depression  was  irore  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)
 O
(Section 10.4.1).
     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  pro!iferative
peri bronchiolitis.
     Thomas et  al.  (1977) 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
0190ZR/A                          10-188                                5/1/84

-------
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 et al. (1981a) published additional data that extend observations
of reduced  resistance  of  mice to streptococcal pneumonia over  a  range  of
                                                                            3
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
39 percent.  Mice  exposed to 0~ at  0.98 mg/m   (0.5 ppm) and challenged with
the  S. pyogenes  aerosol resulted  in  a mean excess mortality (38 percent) that
was  almost equivalent to  the excess mortality  for the group exposed to 28.4
mg/m  of  PAN.  The results  agreed with earlier  reports that PAN  is less toxic
than 0- to mice exposed under ambient conditions.   Exposure to 7.4 mg/m   (1.5
ppm)  of  PAN 3 hr/day,  5  Cays/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 Mycobacterium tubercu-
losis (RIRv) was  influenced by subsequent exposure to PAN.  The exposure  to
this oxidant (25  mg/m  for 6  days)  did  not  alter the  pattern of  bacterial
growth in the lungs of mice.

10.5.2  Hydrogen Peroxide
                                                                             3
     Although concentrations of hydrogen peroxide (HLO,,) as high as 0.14 mg/m
(0.10 ppm) have  been  reported  to  occur in  urban air samples (Chapter 6), very
little is  known  about the effects of H»0? from inhalation exposure.   Most of
the early work on H?0? toxicity involved exposure to very high concentrations.
Oberst et  al.  (1954)  investigated the inhalation toxicity of 90 percent H 09
                                                                        3
vapor in  rats,  dogs,  and rabbits at concentrations ranging from 10 mg/m   (7
ppm) daily for  six months to an 8-hr exposure to 338 mg/m  (243 ppm).   After
autopsy,  all  animals  showed abnormalities of the lung.   In a recent experiment,
Last et  al.  (1982) exposed rats for 7 days  to > 95-percent H?0« gas with a
                           3
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

0190ZR/A                          10-189                                5/1/84

-------
lobe weights, and  protein  or DNA content of  lung  homogenates.   The authors
suggested that because HLO,, 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 the H?CL explore possible mechanisms for the
effects of  H000.   These  include direct cellular effects  associated with in
             c- £-
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 Iff2 to IQ~1M of  H202
increased the surface tension of saline extracts of dog lung homogenates.  The
                                                 3
authors estimated  that  a  clog breathing 1.4 mg/m  (1  ppm) of H909 for 30 hr
                                              -2
would build  up a pollutant concentration of 10   M in the  surfactant, assuming
that all the H^O^  was retained by the lungs.   However, as stated above,  this
estimate of  tissue dose  is not realistic, because  most of  the H?02  would be
absorbed in  the upper airway.
     Another mechanism by which H«0« may affect ventilation is by changing the
                                                                      -4
tone of  airway  smooth muscle.  Stewart et al.  (1981)  reported that 10   M of
H?0? caused significant constriction  of strips of subpleural  canine lung
parenchyma  and of  bovine trachea!is  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 idomethacin
markedly reduced the  response to H^O,,.  This suggested that  the increase in
airway smooth muscle  tone  produced by H?02 involves prostaglandin-like  sub-
stances.
     The potential genotoxic effects from jm vitro H?0?  exposure have   been
evaluated in isolated cell  systems.   Bradley  et  al.  (1979)  reported that H202
produced both toxicity  and single-strand DNA breaks but was  not  mutagenic at
concentrations  up  to 530 |jM.   They  observed  a significant increase  in  the
frequency of reciprocal  sister chromatid exchanges in V-79 Chinese hamster
cells  at  a  concentration of  353 pM of H^CL.   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
H?0? for 3  hr at 37°C.  Increases  in the frequency  of  sister  chromatid exchanges
in  Chinese  hamster cells fiave  also  been  found  by  MacRae and Stich (1979),
Speit  and Vogel  (1982),  and Speit et al.  (1982).  Wilmer  and  Natarajan  (1981)
 0190ZR/A                          10-190                                 5/1/84

-------
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
H?CL.  In comparison, cells were killed with a concentration of 10  M of H^O,,.
Similarly, H?0? (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 al., 1978; Kawachi et al., 1980).
     Johnson et  al.  (1981) reported that the  intrapulmonary  instillation  of
glucose oxidase,  a generator  of  H»0?,  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-
HpOp-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
l-LOp  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
halide-dependent  pathway described  for  cell injury produced by H?0p and lac-
toperoxidase/myeloperoxidase (MPO) (Klebanoff and Clark,  1975).   Any source of
HLOp 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  H^O,,  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 j_n vitro experiments, Suttorp and Simon (1982) demonstrated that
HpOp 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 jjn situ type 2 pneumo-
cytes from which  they were derived.
0190ZR/A                          10-191                                5/1/84

-------
10.5.3  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
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 03  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 0,,
but most of them could  have  been produced by CL.  In most cases,  however, the
biological  effects  presented would be difficult  to  associate with any one
pol1utant.
     Research  on ambient  air and  UV-irradiated or nonirradiated  exhaust mix-
tures  is summarized  in  Table 10-27.  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 infarr:  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

0190ZR/A                          10-192                                5/1/84

-------
                                                    TABLE  10-27.  EFFECTS OF COMPLEX  POLLUTANT MIXTURES
o
 i
GO

Concentration,3
(ppro)


Pollutantb
Exposure
duration and
protocol


Observed effect(s)


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)

0.
c5
NOj,
NO



ox
COX
HC
NO;,
NO
PAN


ox
NO*
HC
ox
COX
HC
NO
X

0
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 ppm or at com-
bined concentrations of > 40 ppm of CO,
16 ppm of HC, and 1.2 ppm of N0x- Indi-
dividual sensitivity demonstrated. Tem-
perature was an important variable.
No consistent effects on conception rate,
litter rate, or newborn survival.
Mouse, Wayne and Chambers,
rat, 1968
hamster,
guinea pig,
rabbit


Mouse, Esnik ct al. , 1971
rat,
rabbit





Mouse Emik and Plata, 1969.


Guinea pig Swann and Balchum, 1966





Mouse Kotin and Thomas, 1957


-------
                                              TABLE 10-27.   EFFECTS  OF  COMPLEX  POLLUTANT  MIXTURES   (continued)
O
i
Concentration,3 .
(ppm) Pollutant
Exposure
duration and
protocol Observed effect(s)
Species
Reference
B, Automobile exhaust
0.012-
3.0
0.04 -
0.06 -
0.06 -

0.04 -
0.15 -
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
1.8
36
100
0.5
0.6
B
60
0.4

30
1.0
2.0
0.49
0.03
Os (max)
HC (propylene)
N02
NO
SO,

03
N02
NO
HC (CH4)

ox
N0x
HC*
CO
03
CO
HC (CH4)
NOj.
NO
SOj,
H2S04
0.5-6 hr
(diesel )




1.5-23 mo




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- irradiation of propylene, SOj, , NO, and
NOj, produced Oa and a routagenic moiety
when collected particles were tested by
the plate-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.
Salmonel la
typhimurium




Mouse,
rat,
hamster,
guinea pig

Mouse



Dog


Dog



Claxton and Barnes, 1981





Hueter et al. , 1966
Lewis et al. , 19t>/



Nakajima et al. , 1972



Bloch et al. , 1972, 1973
Gillespie, 1980

Orthoefer et al. , 1976



                                                           Pulmonary function  for  groups  receiving
                                                           oxidants [irradiated  exhaust  (I) +  SO  ]
                                                           18 months:  no effects
                                                           36 months:  no effects
Dog
Vaughan et al.,  1969
Lewis et al.,  1974

-------
                                              TABLE 10-27.   EFFECTS OF COMPLEX POLLUTANT MIXTURES  (continued)
          Concentration/
             (ppm)
Pollutant
  Exposure
duration and
  protocol
Observed effect(s)c
Species
Reference
o
 i
en
                             61 months:
                                                           RL increased (I, I+SOx).
                                                                          washout increased (I);
                                 P CO,
                                  3  £.
                                                                                    increased (I+SO );
                                                                                                   X
                             2 years  recovery:
                             VD  increased  (I,  I+SOx); DLcQ/TLC  decreased
                             and V  increased  in  all groups;  lung  com-

                             partment volumes  increased  (I+SO ).

                             Morphology  (32-36 months recovery):   air
                             space enlargement; nonci listed bronchi olar
                             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


ox
NOz
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
0.76
2.05
1.71
0.3
1.0
2.0
03
SO*
T-2 Butene
acetaldehyde
o,
NO.,
SO.
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
(AN;;) and increased diffusing capacity.


Voluntary activity (wheel running) Mouse Stinson and Loosli, 1979
decreased 75% after 1-3 days, returning
to 85% of pre-exposure levels by the end
                                                           of 14 days.

-------
O
 I
en
                                              TABLE 10-27.   EFFECTS OF COMPLEX POLLUTANT MIXTURES  (continued)
Exposure
Concentration,3 . duration and
(ppm) Pollutant protocol Observed effect(s) Species Reference
0.40 - 0.52
1.0 - 2.15
1.25

03 24 hr Decreased spontaneous wheel running Mouse Boche and Quilligan,
0^ (gas vapor) activity. I960
Ox (gas vapor) 19 weeks, Decreased conception rate, litter rate, Mouse Kotin and Thomas, 1957
continuous and newborn survival.
          .Ranked by nonspecific oxidant concentration (03 or 0 ).
           Abbreviations used:  03 = ozone; 0  = oxidant; CO = carbon monoxide; NO = nitrogen oxide, NO^ = nitrogen dioxide;
           NO  = nitrogen oxides; SO;, = sulfur* dioxide; HC = hydrocarbon; CH4 = methane; H2S04 = sulfuric acid; PAN = peroxyacetyl  nitrate.
          c  x
           See Glossary for the identification of pulmonary symbols.

-------
products that were  biologically more active than  nonirradiated mixtures.  The
concentration of  total  oxidant as expressed by (L  ranged from 588 to  1568
    3
ug/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
(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 10-27  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
alone or in combination with SO .
                               /c
     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
                      J\
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
                      X                                                     X

0190ZR/A                          10-197                                5/1/84

-------
groups.  The  diffusing capacity for carbon mom'xide (D.   ) was similar  in  all
exposure groups, but when normalized for total lung capacity (TLC), the D.   /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
                                   /\
dogs  exposed  to  I was too small (n = 5) to permit meaningful comparisons.   In
general, pulmonary function changes  were  found to be  similar  in all  groups
exposed to  automobile exhaust  alone  or supplemented with SO  .   Exposure to
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 rconths 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+SO  groups  showed  ramdom enlargement of alveolar airspaces centered in
        A
respiratory bronchioles  and alveolar  ducts.   Small  hyperplastic lesions were
observed at the junction o~ the terminal  bronchiole and the first-order respi-
ratory bronchiole.   Foci  cf ciliary  loss  associated with squamous metaplasia
were also observed in the intrapulmonary bronchi  of the I+SO  group.   However,
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
                     )\
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  plate-incorporation  test (Ames  et al. , 1975).
The  authors demonstrated  that  the irradiation of propylene, SO,,, NO,  and N02
produced 0- and a mutagenic moiety.   In baseline  studies on diesel  exhaust, in
which  0- 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.
0190ZR/A                          10-198                                5/1/84

-------
     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 |jg/m
(0.3 ppm) of GO, 1 ppm of NCL, and 2 ppm of SO,,, 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  NO^  and  SOp  alone had only moderate  effects,  the authors
concluded that 0- had the major influence on depression of activity.
     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,  0^,  and $$2-   Decreases in the  nitrogen washout slope and
increases  in  the diffusing  capacity  indicated a  significant compensatory
change in distrubution of ventilation  in the lungs of exposed animals.   Animals
compromised by the presence of emphysema were unable to respond to this pulmo-
nary insult  in the same manner as  animals without impaired  lung  function.
10.6  SUMMARY
10.6.1  Introduction
     The biological effects of 0, have been studied extensively in animals and
a wide  array  of toxic effects have been ascribed to 03 inhalation.  Although
much has been  accomplished  to  improve  the  existing  data base,  refine the  con-
centration-response relationships and  interpret better the mechanisms  of On
effects, many  of  the  present data were  not  accumulated  with the  idea  that
quantitative comparisons  would be drawn.   In many cases, only qualitative
comparisons can be  made.   To maximize the extent  that animal  toxicological
data can be  used  to estimate the human  health risk of exposure  to 0.,, the
qualitative as  well as  quantitative  similarities between  the  toxicity  of 0.,
to animals and man  must be considered 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 scien-
tifically sound manner, the possible  human responses to 0.
0190ZR/A                          10-199                                5/1/84

-------
10.6.2  Regional  Dpsimetry in the Respiratory Tract
     Experiments  on  the  nasopharyngeal  removal  of 0  in animals demonstrate
that the fraction  of 0~  uptake depends inversely on flow rate and 0~ concen-
tration, that uptake  is  greater for nose than for mouth breathing,  and that
tracheal 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.
     The model of Aharonson et al.  (1974) is useful in analyzing nasopharyngeal
uptake data.  Applied  to 0~  data,  the model  indicates that  the average mass
transfer coefficient  in  the  nasopharyngeal  region increases with  increasing
air flow.   Application of the LaBell et al.  (1955)  model  to nasopharyngeal
uptake is  inappropriate,  because it does not include the effects of chemical
reactions.
     The McJilton  model  (McJilton  et al.,  1972) and the Miller model (Miller
et al. , 1978) for  lower  respiratory  tract 0.,  uptake  are very similar in their
treatment  of  0.,  in the airways  (taking into  account convection,  diffusion,
wall  losses,  and ventilatcry  patterns) and  in their  use of morphological  data
to  define  the dimensions  of  the airways and liquid lining.  However, the
Miller model  is more realistic,  because it  accounts  for chemical reactions  of
0- with  constituents of  the mucous-serous layer.   Tissue dose  is predicted  by
the Miller model to be relatively low in the trachea and to  increase  to a
maximum around the junction  of the  conducting airways  and  the gas  exchange
region and then to decrease distally.  Comparison  of the Miller model  results
to  the  morphological effects of 0-  indicates qualitative agreement  in the
pulmonary  region.   However, comparison in the tracheobronchial region indicates
further research is  needed to define the relevant toxic, chemical, and physical
mechanisms.
     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 0~ and its role in toxicity, but also  to
support  and to  lend confidence to the modeling  efforts.   With experimental
confirmation, models will become practical tools and further our understanding
of the role of 0~  in the  respiratory tract.
     Ozone  dosimetry modeling is in  its initial stages; refinements  and experi-
mental  information are needed.   The  models  (nasopharyngeal and lower respira-
tory  tract) raise  many questions,  not only  about  the model  formulations  and

0190JE/A                          10-200                               5/1/84

-------
assumptions, but more importantly, they bring to the forefront questions about
what is needed  to  understand the absorption of 0_  in the respiratory tract.

10.6.3  Effects of Ozone on the Respiratory Tract
10.6.3.1  Morphological Effects.  Morphological  studies  of  the effects of 0
have indicated  that  the  pattern and distribution of the  tissue lesions are
similar in  the  species  studied and depend on the locations of  the sensitive
cells and the  junction  between the conducting airways and  the gas exchange
regions of  the  lung.   Damage to all parts of the respiratory tract can occur
in animals, depending  on  the 0_ concentration.   At low concentrations  of 0_
            3
(< 1960 ug/m ,  1 ppm)  damage is principally confined to  the junction between
the alveoli and the  conducting  airways.   Dogs, monkeys, and man have respira-
tory and  nonrespiratory bronchioles,  but the common  experimental  animals
(mice, rats, rabbits,  and  guinea pigs) have only nonrespiratory bronchioles.
The location of the  0~ lesion thus differs according to  the species examined
                      O
(Plopper et al., 1979; Castleman  et al.,  1977, 1980; Dungworth  et al.,  1975a;
Eustis et al. ,  1981).   In  both types  of  lungs,  the effects of 0_ have been
                                           3
found at concentrations  as  low as 392 ug/m  (0.2 ppm)  and for exposure times
as short as 2 hr (Stephens et al., 1974a).
     In the upper  and  lower conducting airways, ciliated cells appear  to be
the most sensitive cell  type;  they are damaged by exposures as low as 392 to
1568 ug/m   (0.2 to 0.8 ppm) 8 or 24 hr for  7 days  in rats  (Schwartz et al.,
1976), bonnet monkeys  (Castleman et al.,  1977), Rhesus monkeys (Dungworth et
al., 1975b; Mellick  et al. , 1977), and mice (Ibrahim  et al. ,  1980).   When
Moore and Schwartz (1981)  and Boorman et  al. (1980) extended  these exposure
levels to 180  days,  similar changes in these ciliated cells  were observed.
Uniform damage  is  not  always noted; most  often  reports are of  shortened and
less dense  cilia  occurring  in random  patches.   Electron microscopy revealed
severe cytoplasmic changes  and condensed  nuclei.  Damage is present in both
trachae and bronchi  (Eustis et al. , 1981; Mellick et al. , 1977; Castleman et
al., 1977).  The  damaged  cilia are replaced by  nonciliated clara cells that
become hyperplasic (Evans et al., 1976a; Lum et al. , 1978).
                                             3
     In mice, 0  levels  of 980 and 1568  ug/m   (0.5 and  0.8 ppm) elicited a
pronounced  hyperplasia of  these nonciliated cells that persisted for 10 days
after cessation of the  exposure (Zitnik et al., 1978;  Ibrahim et al.,   1980).
In a  number of species, 0» damage is clearly evident at  the centriacinar

0190JE/A                          10-201                               5/1/84

-------
region, which includes the terminal  bronchiole region,  portions of the respira-
tory bronchioles, and  possibly  the  alveolar ducts, depending on the  species
(Stephens et al., 1973, 1974a,b; Schwartz et al.,  1976; Mellick et al. ,  1977).
     The type 1  epithelial cells are significantly affected (Stephens et al. ,
1974a;  Evans et al.,  1976a; Castleman et al., 1980; Eustis et al. ,  1981; Barry
et al.,  1983; Boorman  et al. ,  1980).   In monkeys, for example, at 1764 pg/m
(0.9 ppm) the type 1 cell death reaches a maximum at  12  hr after continuous
exposure (Castleman et al., 1980).   With the destruction of these  cells, there
is a  hyperplasia of  type 2  alveolar epithelial  cells, which  recovers  the
denuded basal lamina  (Stephens  et  al.,  1974a,b; Sherwin et al., 1983; Eustis
et al.  ,  1981).   The  type 2 cells are relatively  resistant to 0,. exposure.
                                                                O
     Following and during continued  exposures,  these type 2 cells  begin to
proliferate, which is a hallmark of Q» injury regardless of species.  In rats,
DNA synthesis in type  2  cells was reported as early as 4  hr after exposure  to
        3
392 ng/m  (0.2 ppm)  (Stephens  et al. ,  1974a) and reached a maximum at 2 days
                                                3
following continuous exposure to 686 or 980  ug/m   (0.35 or 0.5  ppm) (Evans  et
                           3
al., 1976a,b), or 1568 (jg/m  (0.8 ppm) (Boorman et al. , 1980).  Although type
2 cells  proliferated  in  the  0_-exposed  lung, complete  maturation  of  type 2
cells to type 1  cells did not  occur, even  as late as 180 days of  exposure
(Moore and  Schwartz,  1981).   In the normal  progression,  type  2 cells would
have matured  into type 1 cells.  Continued  0_ exposure inhibits both cilia-
                                              •5
genesis and type 2 cell maturation.
     Inflammation occurs  in  all species so  far examined.   The inflammatory
response is  seen as  early as 4  hr after exposure  to 1568 M9/m  (0.8  ppm)  in
monkeys  (Castleman et  al. , 1980).  In rats and monkeys, inflammation  persists
with continued exposure,  although at reduced levels.   The inflammatory exudate
includes both  fibrin and various leukocytes  in the  initial  phase.    In  the
later phase the  inflammatory cells are predominantly macrophages (Castleman et
al., 1980;  Brummer et al. , 1977; Boorman et al. ,  1980;  Moore and Schwartz,
1981).    Quantitative  estimates  of the degree of  inflammation  are,  however,
lacking  at  present.   The contribution of the inflammatory  response to  sub-
sequent long-term features of 0_ toxicity has not been studied in detail, even
though  techniques are available.  A  number of studies  have  reported that
                     3
exposure to  980  (jg/m   (0.5 ppm)  and above thickens the  interalveolar  septa  of
centriacinar alveoli   (Schwartz  et  al. ,  1976; Castleman et al., 1980;  Boorman
et al., 1980).  Thickness could be due to interstitial  fibrosis.
0190JE/A                          10-202                               5/1/84

-------
     Although some delay has been observed between the exposure to 0_ and the
                                                                    O
maximum manifestation of morphological changes, the shortest time required for
0_ exposure to result in significant morphological changes is 2 hr of exposure
to 392 or 980 ng/m3  (0.2 or 0.5 ppm)  in  rats  (Stephens et al., 1973, 1974a,b)
or 4.7 or  6.6  hr of exposure by  endotracheal  tube in cats  to 510, 980, or
         3
1960 ug/m  (0.26, 0.5,  or  1 ppm) (Boatman et  al.,  1974).   However,  most of
these morphological  studies were  specifically designed to  look at long-term
(i.e., days, months, years)  exposures rather than acute effects.   These data
based on acute  studies  are in agreement with  molecular  theories  that ozone
acts rapidly to  oxidize some critical tissue component.   Little difference in
severity was noted  by  Schwartz  et al. (1976) between exposure to 8 hr/day or
24 hr/c!ay when the exposure was continued for 7 days.  Quantitative studies of
different exposure regimens are lacking, but technically feasible.  The sequence
in which the anatomic sites are affected appears to be a function of concentra-
tion  rather  than of exposure duration.   Increasing  concentrations  not  only
results in more  severe  lesions, but also extends  the  lesion  to higher genera-
tions cf the respiratory structure.
     Morphological studies of vitamin E-deficient or supplemented rats have
been  undertaken  to  correlate the biochemical  findings  with morphological
alterations  (Plopper et al. ,  1979;  Stephens et al., 1983; Chow et al.,  1981;
Schwartz et  al. ,  1976;  Sato et al. ,  1976a,b,  1978,  1980).   Despite the pre-
sence of vitamin E  in the  diets of these animals,  the morphological lesion of
0,. exposure was unchanged.
     When comparisons are  made at the analogous anatomical  site,  the morphol-
ogical effects of 0,. on the lungs of a  number of species of animals are re-
markably similar.  Despite  the  inherent differences in anatomy between most
experimental  animals and man, the junction between the conducting airways and
the gas exchange  region is most affected by 0_.
10.6.3.2  Lung Function.   Changes in  lung function following 0~ exposure have
been  studied in  mice,  rats,  guinea  pigs,  rabbits,  cats, dogs,  sheep,  and
                                                                            3
monkeys.   Short-term exposure for 2  hr  to concentrations of 431 to 980 yq/m
(0.22 to 0.5 ppm) produces rapid, shallow breathing and  increased pulmonary
resistance measured  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  fol-
lowing short-term 0- exposures  lasting  3 hr to 14 days are  usually greatest
0190JE/A                          10-203                               5/1/84

-------
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 a!., 1979).
     Long-term exposure  of 4  to  6 weeks to  0  concentrations  of  392 to
        3
490 ug/m  (0-2 to 0.25 ppm) increases lung distensibility at high lung volumes
in young rats (Bartlett et al., 1974;  Raub et  al. ,  1983a).  Similar  increases
                                                                            3
in lung distensibility were  found in older rats exposed to 784 to 1568 ug/m
(0.4 to 0.8  ppm)  for up to 180 days (Moore and Schwartz, 1981; Costa et al.,
1983; Martin et al.,  1983).   Three to  twelve months of exposure  to 0»  concen-
                             3
trations of 1176 to 1568 ug/m  (0.6 to 0.8 ppm) increased pulmonary resistance
and caused impaired  stability of  the  small, peripheral  airways  in both  rats
and monkeys  (Costa et al. ,  1983;  Wegner, 1982).  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.
     Studies of airway reactivity following 0,. exposure in experimental animals
show that 0~  increases  the reactivity of the lung to mechanical and chemical
stimulation  partly by increased  sensory  neural  activity travelling in  the
vagus nerve  and partly  by a  local action  of  03 on the tissues. Aerosolized
ovalbumin can reach  immunologic receptors  in the lungs of mice  exposed to  980
            3
or 1568 ug/m  (0.5 or 0.8 ppm) continuously for 3 to 5 days (Osebold et al. ,
1980), resulting in an increased incidence of anaphylaxis.  Increased sensiti-
vity to histamine  or cholinomimetic drugs by  aerosol  or injection has been
noted in several  species.   Easton and Murphy  (1967)  showed that the lethal
dose of  histamine in guinea  pigs was reduced  in those  animals exposed to
concentrations as  low as 980 or 1960 ug/m  (0.5 or 1 ppm) of 0. for 2 hr.  The
pulmonary resistance  due  to  subcutaneous injection of histamine increased in
                                3
guinea pigs  exposed  to 1568 ug/m   (0.8 ppm) of 0« for  1  hr (Gordon  and Amdur,
                                                     3
1980).   Similarly, dogs exposed to 1372 and 1960 ug/m  (0.7 and 1.0 ppm) of 03
for  2 hr  had greater  changes in  pulmonary  resistance following histamine
aerosol inhalation (Lee  et al., 1977;  Holtzman et al., 1983a).   Sheep  exposed
to 980 ug/m   (0.5  ppm)  of 0_ for 2 hr experienced increased pulmonary resis-
                            •J
tance for the cholinomimetic drug, carbachol (Abraham et al., 1980).  The time
course of 0_-induced airway  hyperreactivity suggests a  possible association
with  inflammation  (Holtzman  et al. , 1983a,b;  Sielczak et al. ,  1983; Fabbri
et al., 1984), but responses are variable and not very well understood.
0190JE/A                          10-204                               5/1/84

-------
Additional studies that demonstrate  increased collateral  resistance following
30-min local exposure of Cu or  histamine  in  sublobar bronchi of dogs  (Gertner
et a!., 1983a,b,c) suggest that other mechanisms, along with amplification of
reflex pathways, may contribute to changes in airway reactivity.
10.6.3.3  Biochemically Detected Effects of Ozone.  Increased activity of the
glutathione peroxidase system is one of the most sensitive indices of exposure
to <_  1960 ug/m   (1 ppm)  of 0,.  measured biochemically.  The increases in the
glutathione peroxidase  system  are  thought to  be protective,  because this
system  is  involved in antioxidant  metabolism.   Concurrent increases in 0?
consumption of  mitochondria  are also observed.   Mustafa  (1975)  and Mustafa
et al. (1976, 1983) summarized  a number of experiments with rats; they showed
that  continuous  exposure between  392 and  1568 ug/m  (0.2 and 0.8 ppm) for 7
days  caused a concentration-related increase in the activities of  succinate
oxidase,  succinate-cytochrome  c  reductase,  NADPH-cytochrome c  reductase,
lactate dehydrogenase, glucose-6-phosphate dehydrogenase,  glutathione reductase,
glutathione peroxidase,  and   the  levels  of  reduced  glutathione.   Numerous
investigators have also found similar increases when studying rats, mice, and
monkeys (Chow,  1983;  Plopper  et al. , 1979;  Schwartz et  al. ,  1976;  Moore et
al.,  1980;  Fukase et al. , 1975;  Chow  et al. , 1974, 1975; Mustafa  et al. ,
1982).  Increases  in  the glutathione peroxidase  system   have been  reported
after  exposure  of rats on a  vitamin E-deficient diet to  levels  as low  as
196 (jg/m  (0.1  ppm) for 7 days  (Chow et al., 1981; Mustafa, 1975; Mustafa and
Lee,  1976).  The dietary vitamin E  fed to  the rats influenced the increase of
this  system.  For  example, when the diet  of  rats  had  66  ppm of vitamin E,
increased activities were  observed  at  392 ug/m  (0.2 ppm) of 0 •  with 11 ppm
                                             3
of vitamin  E,  increases  occurred at 196  pg/m   (0.1  ppm)   (Mustafa and Lee,
1976).  Several   other  investigators have shown that vitamin E deficiency  in
rats  makes  them more  susceptible to these enzymatic  changes  (Chow et al. ,
1981; Plopper et al. ,  1979; Chow and Tappel,  1972).
     The influence of the pattern of exposure on these effects of 0   on  anti-
                                                                    «3
oxidant metabolism and 0- consumption have been investigated (DeLucia et al. ,
1975a,b; Fukase  et al. ,  1975; Chow and Tappel,  1973;  Mustafa and Lee, 1976).
Generally, over  a  7-day  continuous  exposure  to about 1568 pg/m  (0.8 ppm)  of
03, there  is a  slight (usually nonsignificant)  decrease   in enzyme activities
on day 1 of exposure.   The activities then increase by day 2,  plateau by about
day 4, and are still  elevated at day 7.   When both continuous and intermittent
(8 hr/day) exposure regimens  are compared  for several 0   concentrations  (392,
                                                       «3
0190JE/A                          10-205                               5/1/84

-------
                 3
980, or 1568 (jg/m ; 0.2, 0.5 or 0.8 ppm) , no differences between the regimens
are observed in  rats  (Chow et al . ,  1974; Schwartz et al., 1976; Mustafa and
Lee, 1976; DeLucia et al . ,  1975a).
     Antioxidant metabolism and  0^  consumption have been examined  in  mice,
rats, and monkeys after 0,, exposure.   Generally,  similar effects were  noted,
although different concentrations were  required sometimes to elicit effects.
                                                                            Q
For example, rats were  affected  by  an 8-hr/day, 7-day exposure to 980
(0.5 ppm) of  0-,  but rhesus monkeys were  not  affected (Chow et al., 1975).
Rhesus  monkeys  were  affected  after 7  days  of  intermittent  exposure to
         3
1568 (jg/m  (0.8 ppm)  (Mustafa  and  Lee,  1976).   Rats and bonnet monkeys have
been compared by  Mustafa  and Lee (1976).   After 7 days of intermittent expo-
sure to 980 |jg/m   (0.5 ppm), the rats and  monkeys had  equivalent increases in
succinate oxidase activity,   Direct comparison of mice and rats shows that for
several of the biochemical endpoints, the increases in mice were significantly
greater than  the  increases  in rats (Mustafa et  al . ,  1982).   Quantitative
evaluation of species  sensitivity  from  these studies  is  not  possible.   For
example, in the rat and monkey studies,  some 03 concentrations were not equal,
and statistical  design and analyses were inadequate.   In addition,  differences
in the  deposition  of 0.,  between rats and monkeys is not known.  Thus, it may
                       O
only be generally stated that apparently rats were more responsive  than monkeys.
     Age-dependent responsiveness to 0, has been  described  in  rats exposed to
                 3
1568 or 1764 ng/m  (0.8 or 0.9 ppm) of 0~.   Generally, the youngest rats (7 to
18 days old)  exhibited decreases in  enzyme activities;  20-  to  24-day-old  rats
had marginal  increases (if  at all);  adult,  35- to 90-day-old  rats  had an
increase in enzyme activities (Tyson et al.,  1982; Lunan et al . , 1977; Elsayed
et al .  , 1982a).   The reasons for this age dependence are unknown.   It might be
due to  a  complex interrelationship between sensitivity (because basal levels
of enzymes studied differed by age) and dosimetry of 03_
     The numerous studies of antioxidant metabolism and oxygen consumption are
summarized in Table 10-5.   A few mechanisms for the increases in these metabo-
lic systems are  possible:   increased enzyme content within  a  given cell;  same
enzyme content,  but an increase in the number of cells with such enzymes; or a
combination of the two.   The former  potential mechanism has not been  studied.
Mustafa et al.  (1973),  Mustafa (1975),  and DeLucia et al . (1975a)  found that
Op consumption of  the whole  lung increased, the  number of mitochondria  in the
lung increased,  and the specific enzyme activity of mitochondria (activity per
0190JE/A                          10-206                               5/1/84

-------
milligram of mitochondria! protein, e.g., per mitochondrion) did not increase.
Several  investigators  have  found close correlations between  the  pattern of
changes  in  these  enzymes  and increases in the number of type 2 cells, which
have more mitochondria  than  type 1 cells, in the lung (Plopper et al., 1979;
Chow et  al.,  1981;  Schwartz  et al., 1976; DeLucia et al. ,  1975a).  Thus, the
available experimental  evidence  suggests  that the mechanism  responsible for
these enzymatic changes is the hyperplasia of type 2 cells  after  0_ exposure.
                                                                  O
However, the  presence  of  a  concurrent increase  in activities per cell that
makes a small contribution to the overall response cannot be ruled out.
     The activities of  other enzymes increase in the lung following 0  expo-
sure.  The total lactate dehydrogenase activity was increased in lungs of rats
after exposure  to 980 ug/irr   (0.5 ppm)  (Chow  et al., 1977).  Lysosomal  enzymes
were increased  in lungs  of  rats  (Oil lard et al. , 1972; Castleman  et al. ,
1973a)  and  may  reflect  part  of the inflammatory process (e.g., the influx of
alveolar macrophages,  which have high levels of these enzymes) seen throughout
periods  of  0,  exposure,  as  well  as the breakdown of the  exposed alveolar
macrophages.
     Decreases in the activities of several  enzymes in the lung have also been
reported.  The cytochrome P-450 system has been studied because of its function
in drug  and carcinogen  metabolism.   Montgomery and Niewoehner  (1979)  reported
a 50 percent  decrease  of benzphetamine N-demethylase  activity,  which is  a
cytochrome  P-450-dependent enzyme,  in  the rat following a  24-hr  exposure to
         3
1960 ug/m  (1 ppm)-   In other studies, a decrease in the activity of benzpyrene
hydroxylase  in  lung parenchyma of  hamsters and in tracheobronchial mucosae  of
                                     3
the  rabbit  was  observed at 1470 ug/m  (0.75  ppm)  (Palmer et al.,  1971, 1972).
                                       3
Acute exposure  of rabbits to 1960 ug/m  (1.0 ppm) also decreased levels of
lung micrososomal  cytochrome P-450 (Goldstein, 1975).
     Collagen,  a major  structural  protein of the  lungs, is  also altered  by  0-
exposure.  Typically, the activities of prolyl hydroxylase  (the rate-limiting
enzyme  for the formation of hydroxyproline)  and hydroxyproline (a major consti-
tuent of collagen)  are  measured.   Last et al.  (1979) observed a linear rela-
tionship between  the  increase in  collagen  synthesis  and 0»  concentrations
                         3
between 980 and 3920 ug/m  (0.5 and 2 ppm).   If exposure was prolonged, hydroxy-
                                                      3
proline decreased by  180  days of  exposure to 980 ug/m   (0.5  ppm) (Last and
Greenberg,   1980).   Recovery  from  increases  in  collagen metabolism are  not
                                                       3
immediate.   After a 7-day exposure of rats to 1568 ug/m  (0.8 ppm) of 0,. ceased,

0190JE/A                          10-207                               5/1/84

-------
about 10  days were  required  for recovery of prolyl  hydroxylase activity;
hydroxyproline levels remained increased 28 days postexposure (Hussain, 1976a).
The minimal exposure regimen causing an increase in prolyl hydroxylase activity
                                                    3                     3
in  rats  is a  7-day continuous exposure to 980 ug/m   (0.5  ppm); 392
(0.2 ppm) caused no effects (Hussain et a!., 1976a,b).   The increased collagen
synthesis has been correlated with fibrosis detected histologically in the same
experiment (Last et  al . ,  1979;  Moore and Schwartz, 1981).  Bhatnagar et  al .
(1983) found  that collagen  synthesis  also  increases  in mice  exposed  to
882 ug/m  (0.45  ppm) of  0_.   Thus, the response  is  not species-specific.
                           O
     Although several  investigators  have  found that in  vitro exposure to 0_
                                                       "                     O
affects  lipids,  very few jjn vivo  studies  have been conducted.   Continuous
exposure to 980 ug/m  (0.5 ppm) for  2, 4, or 6 weeks alters fatty  acid compo-
sition of the lungs  in rats, and  acute  (4 hr) exposure  to a  higher level
          3
(1960 pg/m ;  1 ppm)  decreases  incorporation of fatty  acids into lecithin in
rabbits, (Roehm et al  . ,  1972;  Kyei-Aboagye et al . ,  1973).
                                       3
     Ozone exposure  (1176  to 4116 ug/m ;  0.6 to 2.1 ppm) inhibits the phenol
red activity transport mecianism from the lung to the circulation in a concen-
tration-related  manner (Williams  et al.,   1980).    The noncarrier-mediated
diffusion of phenol  red from the lung increased simutaneously with the inhibi-
tion of  the carrier-mediated transport.  An increase in  lung permeability,  as
evident,  by  the  diffusion of plasma  proteins into  the  airways  following  0.,
exposure, has been reported  in dogs (Reasor et al . ,  1979),  rats (Alpert et
al . , 1971a),  and  guinea pigs (Hu et al . , 1982).   The protein content of  lung
lavage fluid from vitamin C-deficient guinea pigs was not different from  those
receiving adequate vitamin C.   The  influx of protein was proportional to the
0_  concentration  to  which the guinea pigs had been exposed in the concentra-
tion range of 510 to 1470 jjg/m3 (0.26 to 0.75 ppm) for 3 hr (Hu et al . , 1982).
     It  is  generally agreed  that the toxic  effects of 0_  can be ascribed to
                                                         O
its oxidative capacity.   As  a powerful oxidant, 0_ can  react with virtually
                                                   »J
every class of biological substance; this makes identification of the critical
or  essential toxic lesion almost impossible.
     Two general  hypotheses  have  been put forth to explain the toxic effects
of  03  on a  molecular basis.   The  first  hypothesis is  that the oxidation of
thiols or amino acids in tissue proteins or  small  molecular weight peptides is
responsible (Mudd  and  Freeman,  1977).   The  second  is  that the  oxidation of
polyunsaturated fatty  acids  (PUFA) to fatty acid  peroxides (FAR)  results in
0190JE/A                          10-208                               5/1/84

-------
toxicity (Menzel, 1970).   A  careful  study of the biochemical, physiological,
and morphological effects  of 0, exposure points to the cellular membranes as
the site of  toxicity.   Both  mechanisms could occur at the membrane.   Because
membranes are composed of both proteins and lipids, damage to one is difficult
to separate  from  damage  tc the other  or  to both.  Because the metabolism of
FAR is  linked  to  the glutathione peroxidase system  (Chow and Tappel, 1972),
depletion of thiols such as glutathione (GSH) by 03 could enhance the toxicity
of FAPs.  The  two theories are  thus not separate,  but  complementary, and most
likely oxidation of both PUFAs and proteins occurs simultaneously.
     Although  CL  is  a general reactant,  the relative  rates  of reaction with
different biological  substances  is  remarkably different.   Ozone  oxidizes
ethylene groups 1000  times faster than other compounds.  Consequently, atten-
tion has been  focused on PUFAs as targets for 0- reaction.   The Criegee mecha-
nism for addition of 0^ across unsaturations is still a popular hypothesis.  In
the presence of protom'c solvents such as water, peroxides can result directly
from the decomposition  of  the initial 03  adduct.   Pryor et  al.  (1983) have
also suggested that  free radicals are generated directly by other mechanisms
during  the ozonation  of  PUFAs.   Regardless of  the chemical  mechanism,  once
free radicals  are generated  in membranes, the toxic  effects  are  similar.
These effects  are  a  loss of regulation of electrolytes and water (changes in
permeability),  leakage of  essential  intracellular enzymes from the cell,  and
inhibition of  metabolic  chains  imbedded  in mitochondria! or  microsomal matri-
cies (Menzel,  1976).   If damage is sufficiently severe, the cell  cannot recover
and dies.  Necrosis  is  commonly reported in the  lungs  of animals exposed to
high concentrations of 0^.
     Perhaps the most compelling evidence for PUFA peroxidation jji vivo is the
protective effect of  dietary vitamin  E on 0- toxicity (Roehm et al., 1971a,b
and 1972; Chow and  Tappel.  1972; Fletcher and  Tappel,  1973; Donovan et al.,
1977; Sato et  al., 1976a,b;  Chow  and  Kaneko, 1979; Plopper et al., 1979; Chow
et al.,  1981;  Chow,  1983;  Mustafa et  al., 1982).   Vitamin  E as a phenolic
antioxidant easily inhibits the peroxidation of PUFAs initiated in vitro by 00
                                                                   , . ,        j
(Roehm  et  al., 1971a).   Indirect evidence that vitamin E protects against
03-initiated peroxidation  of PUFAs  u> vivo is  presented by  the studies of
Dumelir et al.  (1978b),  who  found that the exhaled ethane and pentane in  the
breath  of  rats was decreased when  vitamin E was  provided in  their diets.
0190JE/A                          10-209                               5/1/84

-------
Ethane and pentane are thought to be produced by peroxidation of PUFAs initi-
ated by 0  inhalation.  Changes in lung fatty acid composition due to peroxi-
dation are  not easily interpreted because  of the complexity of  lung PUFA
composition and  difficulties  in quantitative analysis.  Direct  evidence of
peroxidation of PUFAs in  human subjects has, however, been  lacking.
     Amino acids are  oxidized  by  0-3  (Mudd  et al. ,  1969;  Mudd and Freeman,
1977; Previero et  al. , 1964).   Cysteine,  methionine and tryptophan  are  the
three most easily  oxidized  amino  acids.   Human  alpha-1-protease inhibitor is
destroyed by  ozonolysis  i_n vitro,  presumably by oxidation  of  methionine,
tyrosine and  tryptophan amino  acid  residues in  the  molecule (Johnson, 1980).
The activity  of  a  large  number of enzymes  is decreased immediately  after  0
                             3
exposure to 1960 to 7840 ug/m   (1 to 4 ppm) for 1 hr or more (DeLucia et al.,
1972; Mountain, 1963; Kyei-Aboagye et al., 1973;  Holzman et al., 1968; Goldstein
et al., 1975;  Menzel  et al., 1976).  Lung mitochondria  lose their respiratory
control following  0_  exposure  (Mustafa et al., 1973).   DeLucia et al. (1975b)
also reported  the  depletion of GSH and the formation of mixed disulfides  in
the lungs of 0 -exposed rats.
              O
     In contrast to  the  decrease  in enzyme activity by acute, high-level  0_
                                                                            O
exposure, the  glutathione peroxidase system consisting  of  glucose-6-phosphate
dehydrogenase, glutathione  reductase and  glutathione peroxidase is increased
                                                                          3
in the lungs of rats exposed repeatedly to lower levels of 0_ (< 1960 yg/m , I
                                                            o
ppm) (Chow et al., 1981,  1974; Chow and Tappel,  1972,  1973; Mustafa and Lee,
1976; Moore et al. ,  1980).   These effects  would  favor  the concept that 0
                                                                           o
increases PUFA peroxidation,  which  induces  the  glutathione peroxidase system
as a protective  mechanism.  Thus,  with the available  data,  it  would appear
that both direct oxidation  of  low-molecular-weight compounds and proteins,  as
well as lipid peroxidation, can occur with 0~ exposure.
10.6.3.4  Host Defense Mechanisms.   To protect  man  and other mammals from
inhaled infectious  agents,  the mammalian  lung possesses an array of defense
mechanisms.    Considerable effort  has  been expended in studies of the effects
of  0-  on  such defense mechanisms in hopes  of understanding the underlying
     O
mechanisms and predicting concentration-response relationships.
     Inhaled  viable  microorganisms as  well  as airborne particles are removed
from  the  lung by  two mechanisms.  Ciliated cells of the  respiratory tract
propel mucus  carrying entrapped particles and infectious agents upwards from
0190JE/A                          10-210                               5/1/84

-------
the distal  regions of  the  lung  to  be  swallowed  and  excreted.   Alveolar macro-
phages engulf, kill,  and remove infectious agents from the respiratory regions
of the  lung either  by migrating with  the  microbe to the mucus lining to be
propelled upwards or  to the lymphatic system to  be drained from the lung.
     Ciliated cells are dcimaged 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.  Ciliated cell
damage should result in decreased transport of viable and non-viable particles
from the  lung.   Rats  exposed to 784,  1568,  or  2352 ng/m  (0.4, 0.8, or 1.2
ppm) for  times as short as  4 hr have  decreased  short-term clearance  of parti-
cles from the lung (Phalen  et al.,  1980; Frager et  al. ,  1979;  Kenoyer et al.,
1981).   Short-term clearance is mostly due to mucus removal of particles, and
the decreased short-term  clearance is an  anticipated  functional  result pre-
dicted from morphological observations.  The mucous glycoprotein production of
the trachea is  also  altered  by 0~ exposure, but not in a simple manner.
Mucous  glycoprotein biosynthesis,  as measured ex vivo  in cultured  tracheal
explants from exposed rats, was inhibited by short-term continuous exposure to
         3
1568 ug/m   (0.8 ppm)  of 0  for 3  to  5 days  (Last  and Cross,  1978;  Last and
                          *.!
Kaizu, 1980;  Last et  al.,  1977).  Glycoprotein synthesis and secretion  re-
covered to  control values  after 5 to  10 days  of exposure and  increased to
greater than control  values after 10 days of exposure.
     The  long-term clearance of particles  from the lung  is a function of the
alveolar macrophages  (AM),  which are damaged by 0~ inhalation.  Most studies
                                                            3
have been with rabbits, in which a 3-hr exposure to 980 ug/m  (0.5 ppm)  reduces
phagocytosis (Coffin  and Gardner,  1972b;  Coffin  et al. ,  1968b),  disrupts AM
membranes (Dowel! et  al. ,  1970; Goldstein  et al., 1977;  Hadley et al., 1977),
reduces AM  lysosomal  hydrolyase activity (Hurst et  al.,  1970,  1971), prevents
AM migration (Schwartz and Christman, 1979; McAllen et al.,  1981), and reduces
AM numbers  (Coffin et al.,, 1968b; Alpert  et al. , 1971b).  The bactericidal
capacity of  AMs  is  also impaired; this has been attributed to the inhibition
of superoxide anion radical  production capacity by 0~ exposure (Witz et al.,
                                                     «5
1982;  Amoruso et al.,  1981; Goldstein et al., 1971a,b, 1972; Warshauer et al.,
1974;  Bergers et al.,  1982).
     An integrated measure of  both of these mechanisms of viable organism
removal and  the more  general measures  of host defense against  airborne infec-
tions   can be determined by the  airborne infection  of animals following 0
0190JE/A                          10-211                               5/1/84

-------
exposure.   This method, known as the infectivity model, has recently been de-
scribed in  detail  by  Gardner  (1982a).   A number of different  animals  and
bacterial  infectious agents have been used to illustrate the increased suscep-
tibility of animals to airborne infections following very modest concentrations
and times of  exposure.  To date, it is one of the most sensitive measures of
ozone toxicity.  Mice  exposed  to 157 to 196 ug/m  (0.08 to 0.1 ppm) for 3 hr
are significantly more susceptible to bacterial infection than mice exposed to
clean air (Coffin  et  al. ,  1968a; Ehrlich et al., 1977; Miller et al. , 1978).
This increase in susceptibility was also evident after an intermittent exposure
           3
to 196 ug/m   (0.1  ppm) fo<~ 103 days (Aranyi  et al. ,  1983).  Adding SO  and
(NH.)? SO. to 0- did not further enhance the mortality rate.  However, combining
0- with either  NO-  or  H?SC. had  an  additive  effect  in  this  infectivity  system
(Gardner et al. ,  1977; Grose  et al. , 1982;  Ehrlich et al., 1977;  Ehrlich et
al., 1979;  Ehrlich, 1980).  The effects of  0»  were more pronounced when the
test animals exercised during  0« exposure (Illing et al., 1980.)  Such a study
                               o
demonstrates  that  the activity  level  of the test  subject  is  an important
variable that can influence the host response.
     Immunological  effects of  0_ have  only been scantily studied in mice and
guinea pigs.  Depression of the cell-mediated immune response has been measured
                                     3
at high 0.,  concentrations  (2901 ug/m ,  1.48 ppm)  in  the guinea pig (Thomas
et al., 1981b).  The syste;nic  immune response is depressed  in mice, as measured
by the  blastogenic  response of  splenic  lymphocytes  to  the T-cell  mitogens  PHA
                                                                3
and Con-A  (Aranyi  et  al. ,  1983).  Mice were exposed to 196 ug/m  (0.1 ppm)  5
hr/day, 5 days/week for 90 days.
10.6.3.5  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 low  levels of exposure to  0_ will certainly protect
against subsequent  lethal  doses and the development of edema (Stokinger et al.,
1956; Fairchild,  1967; Coffin  and Gardner, 1972a).  The delay in mucociliary
activity  (i.e., clearance  reported for 0_)  can also be eliminated by  pre-
exposure  to  a lower concentration. This effect is only 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 03,  such as reduced  functioning
activity  of  the pulmonary  defense  system (Gardner et  al.,  1972);  hyperplasia
0190JE/A                          10-212                               5/1/84

-------
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 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  cells  attempt to initiate early  repair  of the
              O
specific lesion.    This  repair thus results in a reduction  of the effect first
observed.   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).  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 routes, 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 nonreversible pulmonary changes.

10.6.4  Extrapulmonary Effects
     It was  formerly believed that 0-, on  contact with respiratory system
tissue,  immediately reacted and thus was not absorbed or transported to extra-
pulmonary sites.    However, several studies suggest that either 0., or products
                                                                O
formed by the  interaction of 0~ and respiratory system tissue produce  effects
in lymphocytes, erythrocyt.es, and serum, as well  as  in the parathyroid gland,
the myocardium, the liver, and the CNS.  Ozone  exposure also produces  effects
on animal behavior that may be caused by pulmonary consequences of 0  , or by
nonpulmonary (CNS) mechanisms.   The mechanism by which 0_ causes extrapulmonary
                                                        O
changes is unknown.  Mathematical  models of 03 dosimetry predict that very little
0_ 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.
10.6.4.1  Central   Nervous System and Behavioral Effects.   Ozone significantly
affects the  behavior  of  rats  during  exposure  to  concentrations as  low as
        3
235 ug/m  (0.12 ppm)  for 6 hr.  With  increasing concentrations of 0~,  further
                                                                   O
decreases in unspecified motor activity and in  operant learned behaviors have
0190JE/A                          10-213                               5/1/84

-------
been observed  (Konigsberg and Bachman,  1970; Tepper  et  al. ,  1982;  Murphy
et al. ,  1964; and Weiss et al., 1981).  Tolerance to the  observed decrease in
motor activity may occur  on repeated exposure.  At low CL exposure concentra-
               3
tions (490 ug/m ,  0.25 ppm) an increase in activity is observed after exposure
                                          3
ends.  Higher CL  concentrations  (980 ug/m ,  0.5 ppm) produce  a  decrease in
                •3
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 0~,  although this
                                                              •j
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 CL impairs the inclination to respond.
                                         O
Two  studies  indicate  that mice  will  respond to remove  themselves  from an
atmosphere containing greater  than 980 ug/m   (0.5 ppm)  (Peterson and Andrews,
1963, Tepper et al. ,  1983).   These  studies suggest that the aversive effects
of 0. may  be due  to lung "irritation.   It is unknown whether lung irritation,
    O
odor, or a  direct  effect  on the CNS causes change in rodent behavior at lower
0- concentrations.
 %J
10.6.4.2  Cardiovascular  Effects.   Studies on the effects of CL on the cardio-
vascular system are few,  and to date there are  no reports of attempts to con-
firm these  studies.   Structural  changes   in the cell membranes and nuclei  of
the myocardium muscle fibers  in  mice were found after 3 weeks of exposure to
        3
392 ug/m  (0.2 ppm) (Brinkman et al., 1964), and these effects were reversible
in clean air.  The exposure of rats to 0,, alone or in combination with cadmium
          3
(1176 |-g/m  , 0.6 ppm 0~)  resulted in measurable increases in systolic pressure
                      •J
and heart rate (Revis et  al., 1981).   No  additive or antagonistic response was
observed with the  combined exposure.   Pulmonary capillary blood  flow and PaO_
                                                          3
decreased  30 min   following exposure  of  dogs to 588 ug/m  (0.3  ppm)  of 0-
(Friedman et al. ,  1983).   The  decrease in pulmonary  capillary  blood  flow per-
sisted for as long as 24  hr following exposure.
10.6.4.3   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.
0190JE/A                          10-214                               5/1/84

-------
     Effects of 0-  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
RBCs isolated from monkeys exposed to 1470 ug/m  (0.75 ppm) of 0_ 4 hr/day for
                                                                  3
4 days (Clark  et  al. ,  1978).   A single 4-hr exposure  to 392  ug/m (0.2  ppm)
also caused increased fragility as well as sphering of RBCs of rabbits (Brinkman
et al., 1964).   An increase in the number of RBCs with Heinz bodies was detected
                                        3
following a 4-hr exposure to 1666  ug/m  (0.85 ppm).   The presence of  such
inclusion bodies  in  RBCs  is  an indication of  oxidant  stress  (Menzel  et  al.,
1975a).
     These morphological changes are frequently accompanied by a wide range of
biochemical effects.   RBCs of  monkeys exposed to 1470 ug/m  (0.75 ppm) of 0_
for 4  days  also  had a decreased level of glutathione (GSH)  and decreased
acety1cholinesterase (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 uniquely sensitive to 0~.  The  RBCs
from these  animals,  after being exposed to 03, had a  significant increase in
the activity of GSH peroxidase,  pyruvate kinase, and lactic dehydrogenase, but
had a  decrease  in RBC  GSH after exposure to 1568 ug/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,  or  catalase activities.   At a lower  level  (980 ug/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 980 ug/m  (0.5 ppm).   These effects were not observed
in vitami;i  E-supplemented  rats.   Mice on a vitamin  E-supplemented  diet and
those  on  a deficient  diet showed an  increase in G-6-PD activity after an
exposure of 627 ug/m   (0.32  ppm)  of 0  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
(Veninga,  1967).   Six-  and 10-month  exposures  of rabbits to 784 ug/m3 (0.4 ppm)
0190JE/A                          10-215                               5/1/84

-------
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;  and  Jegier, 1973).   Chow et al.  (1974)  reported that
the  se'^um  lysozyme level  of  rats increased  significantly after 3 days  of
continuous exposure to 0, but was not affected when the exposure was interim" t-
                        6                                                   3
tent  (8 hr/day,  7 days).   The CL  concentration in both  studies was  1568 ug/m
(0.8 ppm) of 0-.
     Short-term  exposure  to  low  concentrations  of  CL  induced an  immediate
change in the serum creatine  phosphokinase level in mice.  In this  study,  the
CL 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
0. have also been seen following in vitro exposure of RBCs from humans (Freeman
 3                                  ______
and  Mudd, 1981;  Menzel  et al., 1975b; Verweij and Van  Steveninck,  1981).   A
common effect observed  by a  number of  investigators is  that 0~  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 sphercytosis  and increased fragility of RBCs seen in  animal and human
cells.
      Although these ij} 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 camot be determined with any accuracy.
10.6.4.4  Reproductive  anc Teratogenic Effects.   Pregnant animals may be  at
greater risk to  the effects of 0~ because the volume of air  inspired generally
                                O
increases up  to 50 percent during pregnancy.   Experiments with  mice  suggest
                                   3
that  perinatal exposure to 392 ug/m  (0.2 ppm) of 0_ can  reduce  infant survival
(Brinkman et al., 1964) and increase the incidence of unlimited  incisor growth
                                                                  3
in neonates  (Veninga,  1967).   Pregnant rats  exposed to 2920  ug/m   (1.49 ppm)
of 0_ during  mid-gestation had a significantly  higher  embryo resorption rate
     O
than  control   rats  (Kavlock  et al. ,  1979).   Progeny  of  dams exposed  to
         3
1960  ug/m   (1.0  ppm)  of 0., during late gestation showed  significantly slower
 0190JE/A                          10-216                                5/1/84

-------
neonatal growth rates, retarded early reflex development, and delayed onset of
grooming and rearing behavior (Kavlock et al.,  1980).
10.6.4.5  Chromosomal and Mutational Effects.   That 0_ may  have the potential
                                                     O
for producing genotoxic effects has been postulated because of its radiomimetic
properties.   The decomposition  of 0- in water produces  OH  and HO  radicals,
and it is these species that are considered to be the most biologically active
products of  ionizing  radiation.   Cytogenetic  toxicity has been reported both
in celTs in  culture and in cells  isolated  from  animals  exposure to 0_.  A
                                                                      »J
large portion of this data base, however, has  been derived from studies conduc-
ted at  extremely high levels of 0 , and because of this, these data will not
                                  O
be discussed further in this summary.  The most relevant data concerning geno-
toxicity of  0   are  obtained  from  investigations  in which the 0  concentration
                         3
did not exceed 1,960 ug/m  (1.0 ppm).
     Zelac et  al.  (1971a,b) reported  a significant  increase in chromosome
aberrations  in  peripheral  blood lymphocytes in  Chinese  hamsters  exposed to
        3
392 ug/m  (0.2 ppm) for  5 hr.   It was  also  reported that 0  was additive with
radiation.  Tice et  al.  (1978) attempted to  repeat the Zelac studies; however,
a number of  significant differences existed between these  two  studies  that
make direct  comparison  of  the  data difficult.    Tice et  al. did not find any
increase in  sister  chromatid exchange  or any increase in number of chromosome
aberrations  of  phytohemagglutinin-stimulated peripheral  blood,  but they did
find an  increase in the  number  of chromatid deletions and  achromatic lesions.
Although both studies reported significant increases  in chromosome aberrations
in peripheral  blood lymphocytes,  the specific type of lesions reported were
clearly different.   Section 10.4.5 gives a detailed discussion of the importance
of these differences.   Gooch et al.  (1976) analyzed the bone marrow samples
                                          3
from Chinese hamsters  exposed  to 451 \ig/m  (0.23 ppm) of 0  for 5 hr and the
                                                                            3
leukocytes and  spermatocytes from mice  exposed for up to 2  weeks to 412 |jg/m
(0.21 ppm) of 0_.   No effect was found on either the  frequency of chromatid or
chromosome aberrations,  nor  were  there any reciprocal translocations in the
primary spermatocytes.  Guerrero et al. (1979)  reported a concentration-related
increase in  the sister  chromatid exchange  in human fetal  lung cells (WI-38)
exposed i_n vitro to concentrations ranging from 490 to  1960  ug/m  (0.25 to
1.0 ppm) for 1 hr.
     In summary, 0^ has been shown to produce  a wide  range of toxic effects on
cells and  on cellular components,  including genetic  material.   Cytogenetic
toxicity has been  reported  in  cells in  culture  and  in  cells isolated from
0190JE/A                          10-217                               5/1/84

-------
animals if 03 exposure has occurred at sufficiently high levels  and for suffi-
cient long periods.  The irutagenic effects of 0- have been examined in only a
very few  instances,  and  unfortunately none of these investigations have used
mammalian cells.  The  mutagenic  properties of 0- have  been  demonstrated  in
procaryotic and eucaryotic  cells;  however, the concentrations used were many
orders of  magnitude  highe1" than what is  found  in  the ambient environment.
10.6.4.6  Other Extrapulmonary Effects.   A series of studies was conducted to
show that 0-  increases drug-induced  sleeping time  in a number of  species of
animals  (Gardner  et  al.,  1974;  Graham,  1979; Graham et al.,  1981, 1982a,b,
1983a,b).  At 1960 ug/m3  (1.0 ppm),  effects  were observed after 1, 2, and 3
days of  exposure.  As  the  concentration of 03 was reduced, increasing  numbers
of dai'y 3-hr exposures were required to produce a significant effect.   At the
                                      3
lowest concentration studied (196  ug/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., 1983a).
     Several   investigators  have  attempted  to elucidate  the involvement of the
endocrine  system  in 03  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)  investigated the  effects of a  24-hr exposure to  1960
(1.0 ppm) of  03  on the hypothal amo-pi tui tary-thyroid system of rats.  These
three  organs  regulate the function of  each  other through various hormonal
feedback mechanisms.  Ozone caused decreases in serum concentration 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 03 on TSH.   The thyroid
gland  itself  was  altered  (e.g., edema) by 0-.  The authors interpreted these
findings as  an 03-induced lowering  of the hypothalamic  set point  for the
pituitary-thyroid  axis  and a  simultaneous  reduction of the activity of pro-
lactin inhibiting  factor  in  the hypothalamus.  The anterior pituitaries had

0190JE/A                          10-218                               5/1/84

-------
fewer cells but more TSH per cell.   These cells also released a greater amount
of TSH into the culture medium.
     By using  LM  and  TEM,  Atwal et al.  (1975), Atwal and Wilson (1974), and
Atwal and  Pernsingh  (1981) described changes  in parathyroid glands  in dogs and
rabbits following short-term (up to 48 hr) exposure to 0.75 ppm of 0 .   Changes
                                                                    •5
identified include parenchyma! atrophy, congestion, abundant secretory granules,
leukocyte infiltration, and focal hemorrhages.

10.6.5  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.  The  few toxicological
studies on PAN  indicate that  it is much  less acutely  toxic than 0  .  When the
                                                                 O
effects seen  after  exposure to  0, 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  0« to  produce a similar effect on  lethality,
                                 «J
behavior modification,  morphology,  or significant  alterations  in  host pul-
monary defense  system (Campbell et al.,  1967;  Dungworth et al. , 1969;  Thomas
et al, , 1977,  1981a).   The  concentrations of PAN  required  to produce  these
effects are many  times  greater  than what has been measured in the atmosphere
(0.037 ppm).
     Similarly, most of the investigations reporting H?CL toxicity have involved
concentrations much higher  than those found in the ambient air (0.1 ppm),  or
the  investigations  were conducted  by using various j_n  vitro techniques for
exposure.   Very limited information is available on the health significance of
inhalation exposure to  gaseous  H_0_.   Because H_0? is  highly soluble,  it is
generally assumed that  it does  not penetrate into  the alveolar  regions  of the
lung but is instead deposited on the surface of the upper airways  (Last et al.,
1982).  Unfortunately,  there  have  not been studies designed to look for pos-
sible effects in this region of the respiratory tract.
     A few j_n vitro studies have reported cytotoxic, genotoxic,  and biochemical
effects of H?0? 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.

0190JE/A                          10-219                               5/1/84

-------
     Field and epidemiological  studies  have  shown that human health effects
from exposure to  ambient  nixtures  of oxidants and other airborne pollutants
can produce adverse human health effects (Chapter 12).  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 tc  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.,
1963; Murphy,  1964;  Nakajima et al.  , 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 end 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
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  S0v did not  interact  in any  synergistic or additive
                          s\
manner.
0190JE/A                          10-220                               5/1/84

-------
10.7  REFERENCES


Abraham, W.  M. ;  Januszkiev/icz, A.  J.;  Mingle, M.; Welker,  M.;  Wanner,  A.;
     Sackner, M.  A.  (1980)  Sensitivity of bronchoprovocation  and trachea!
     mucous  velocity  in detecting airway  responses to  0  .   J.  Appl.  Physio!.
     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.

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. 27: 654-657.

Allman, P.; Dittmer,  D. S. (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.;  Schultz,  R. Z. ; Drinker,  P. (1952) Toxicity  of  sulfuric  acid
     mist to guinea pigs.  Arch.  Ind. Hyg. Occup. Med.  5: 318-329.

Amdur, M. 0.; Ugro, V.; Underhill, 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.

Aranyi, C.  ;  Fenters,  J.;  Ehrlich,  R.; Gardner, D. E.  (1976) Scanning  electron
     microscopy  of  alveolar macrophages after exposure to oxygen, nitrogen
     dioxide and ozone.   EHP Environ. Health  Perspect.  16:  180.
0190JE/A                          10-221                               5/1/84

-------
Aranyi,  C., Vana, S. C.;  Thomas, P.  T. ;  Bradof, J. N.; Renters, J. D.; Graham,
     J.  A.; Miller,  F. J.  (1983) Effects of subchronic exposure to a mixture
     of CL, S09  and (NHA)5SO.  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.
     Am.  J.  Pathol. 102:  297-307.

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,
     ultrastructural,  and  immunofluorescent  study.   Am. J.  Pathol.  80:  53-68.

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. ;  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.; Dura,  E. A.  (1983) Influ-
     ence of air polluting components viz.  ozone and  the open air  factor on
     host-resistance towards respiratory infections.  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.  459-467.
     (Advances in modern environmental toxicology:  vol.  5).

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

Bhatnacar,  R.  S.; Hussain,  M. Z.; Sorensen,  K.   R. ; Mustafa,  M. G. ; von Dohlen,
     FT 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,   N.C.   Princeton,  NJ:    Princeton
     Scientific  Publishers,  Inc.; pp. 311-321.   (Advances  in modern  environ-
     mental toxicology:  vol.  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.

0190JE/A                          10-222                               5/1/84

-------
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.

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.

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.; Elartin,  C. J.  (1983)   Morphometric changes in  rabbit
     lungs before and after pneumonectomy  and  exposure to ozone.   J.  Appl.
     Physiol.  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.; Dungworth, D.  L. (1980) Pulmonary effects of
     prolonged ozone  insult  in rats.  Morphometric evaluation of the central
     acinus.   Lab.  Invest. 43: 108-115.

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

Boushey, H.  A.;  Holtzman, M.  J.; Sheller, J.  R.; Nadel, J. A.  (1980) State of
     the art:   bronchial nyperreactivity. Am.  Rev. Respir. Dis. 121: 389-413.

Bowes, C. ;  Gumming, G. ; Horsfield,  K.;  Loughhead, J.;  Preston,  S.  (1982)  Gas
     mixing  in  an asymmetrical model of the pulmonary acinus and asymmetrical
     alveolar  ducts.   J. Appl.  Physiol.  Respir.  Environ. Exercise  Physiol.
     52':  221-225.

Bradley, M.  0. ;  Erickson, L.   C. (1981) Comparison of the  effects of hydrogen
     peroxide  and x-ray irradiation on  toxicity, mutation,  and DMA 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. ; jap.  849-891.

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.
0190JE/A                          10-223                                5/1/84

-------
Brinkman, R.;  Lamberts, H.;  Veninga, T.  (1964) Radiomimetic toxicity of ozonized
     air.  Lancet I:  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  jarocedures.   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.  S.;  Greenwald, 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 Delated photochemical oxidants; March 1982; Pinehurst,
     NC. Princeton, NJ: Princeton  Scientific  Publishers,  Inc.;  pp.  103-117.
     (Advances in modern environmental toxicology:   vol. 5).

Calabrese,  E.  J.; Moore, G.  S.;  Williams, P. S. (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. ; Williams, P.  S.  (1983b) 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.

Campbell,  K.  I.;  Hilsenroth,  R. H.  (1976)  Impaired resistance to  toxin  in
     toxoid-immunized mice exposed  to ozone on 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 29: 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.
0190JE/A                          10-224                                5/1/84

-------
Case, G. D. ; Dixon, 0. 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.
     Psthol. 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.

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  combi-
     nation.  J. Toxicol.  Environ.  Health  4: 845-852.

Cavender, F.  L. ; Steinhagen,  W.  H.; Ulrich, C.  E.; Busey, W.  M.; Cockrell, B.
     Y. ; Baseman,  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, Tpxicol. Environ. Health 3: 521-533.

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

Chaney, S. ;  DeWitt, P.;  Blomquist, W. ; Muller,  K. ;  Bruce,  R.;  Goldstein, G.
     (1979) Biochemical changes in  humans  upon exposure to ozone  and exercise.
     Research Triangle Par«, NC:   U.S.  Environmental Protection Agency,  Health
     Effects Research  Laboratory;  EPA  report  no.  EPA-600/1-79-026.  Available
     from:  NTIS, Springfield, VA;  PB80-105554.

Chaney, S. G.  (1981)  Effects  of ozone on leukocyte DMA.   Research  Triangle
     Park, NC:   U.S.  Environmental  Protection Agency,  Health  Effects Research
     Laboratory;  EPA  report no.  EPA-600/1-81-031.  Available  from:   NTIS,
     Springfield, VA; PB81-179277.

Chang,  H-K;  Farhi,  L.  (1973)  On mathematical  analysis  of  gas  transport in the
     lung.  Respir. Physiol. 18:  370-385.

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.

0190JE/A                          10-225                               5/1/84

-------
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.;  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.; 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. ; 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.

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. ; 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.; 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.

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. I. ;  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.

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
     macrophages induced by  short-term  ozone insult.   Environ. Res. 28: 241-250.


0190JE/A                          10-226                                5/1/84

-------
Clark, R.  A.;  Klebanoff,  S. J.  (1975)  Neutrophil-mediated tumor cell  cyto-
     toxicity:  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 typhimuriun test system.  Mutat. Res.  88: 255-272.

Clement,  J. ;  Afschrift, M.; Pardens, J. ;  Van de Woestline,  K.  (1973)  Peak
     expiratory flow  rate  and rate of.  change of pleura! pressure.   Respir.
     Physiol.  18:  222.

demons,  G.  K. ; Garcia, J.  F. (1980a)  Endocrine aspects of ozone exposure in
     rats.  Arch.  Toxicol.  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.

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, England.   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 infections.
     In:   Proceedings of the third annual conference on atmospheric contamina-
     tion  in  confined spaces; May; Dayton,  OH.   Wright-Patterson  Air  Force
     Base,  OH:   Aerospace   Medical  Research Laboratories;  pp.  71-80;
     AMRL-TR-67-200.

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.

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:  265-71.

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).

0190JE/A                           10-227                                5/1/84

-------
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.

DeLucia,  A. J.; Hogue, 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:  no. 147,  v. 71).

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.

Donovan,  D. H.; Menzel, D.  B. (1975) Dietary  and environmental  modification  of
     mouse lung lipids.   American Oil Chemists'  Society  (Abstract).

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

Donovan,  D.  H.;  Williams,  S. J.;  Charles, J. M.;  Menzel, 0.  B. (1977) Ozone
     toxicity:    Effect  of  dietary vitamin E  and polyunsaturated fatty acids.
     Tcxicol. 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
     congress  on  toxicology;  September;  San Diego,  CA.   Toxicol. Lett.
     18(suppl.  1): 146.

Doty,  R.  L.  (1975) Determination of  odor preferences  in  rodents;  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.

Dowell,  A.  R.;  Lohrbauer,  L. A.; Hurst,  D.;  Lee, S.  D. (1970) Rabbit alveolar
     macrophage  damage  caused by jj\  vivo ozone inhalation.   Arch.  Environ.
     Health  21: 121-127.
0190JE/A                           10-228                                5/1/84

-------
Downey, J. E.; Irving, D. H.; Tappel, A. L. (1978) Effects  of  dietary  antioxi-
     dants on jji  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
     mutations  and  genetic  recombination in Saccharomyces  cerevisiae.   Mutat.
     Res.  102: 249-259.

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:
     129-134.

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 exposure to peroxyacetylnitrates.  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) Effects  of ambient levels  of
     ozone on monkeys.   Fed.  Proc.   Fed. Am.  Soc.  Exp.  Biol.  34:  1670-1674.

Easton, R. E.;  Murphy,  S.  0.  (1967)  Experimental  ozone  preexposure and hista-
     mine. Arch. Env. 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  6: 638-642.

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

Ehrlich, R.  (1983)  Changes  in susceptibility  to respiratory infection caused
     by exposure 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).


0190JE/A                          10-229                               5/1/84

-------
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.  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.

Elsayed, N.;  Hacker, A.;  Mustafa,  M.; Kuehn, K.; Schrauzer, G.  (1982b) Effects
     of decreased  glutathlone  peroxidase activity on the pentose phosphate
     cycle in  mouse lung.  Biochem.  Biophys.  Res.  Commun.  104:  564-569.

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.   International Review of Physiology.   Vol.
     14:   Respirat.  Physiol.  II,  Widdicombe, J. G. ,  (ed.); University  Park
     Press, p.  321.

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

Evans,   M.  J. ;  Stephens,  R,  J. ; Freeman,  G.  (1976c).   Renewal of  pulmonary
     epithelium following oxidant injury.   In:  Bouhuys,  A., ed.  Lung  cells
     in disease.   New York,  NY:   Elsevier Press; pp. 165-178.

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. ;  Johnson, L. V.;  Stephens,  R.  J.;  Freeman, G. (1976a) Renewal of
     the terminal  bronchiolar  epithelium in  the rat following  exposure to N0~
     or 03.  Lab.  Invest. 35: 246-257.
0190JE/A                          10-230                                5/1/84

-------
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.

Fabbri,  L.  M.;  Aizawa,  H.; Alpert, S.  E.;  Walters, E.  H.;  0'Byrne,  P. M. ;
     Gold, B. D.; Nadel, J. A.; Holtzman, M.  0.  (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. (1963) Neurohumoral factors in  injury  from  inhaled  irritants.
     Arch. Environ.  Health 6: 85-92.

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

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.

Fairchild, E.  J. ;  Graham, S. L.  (1963)  Thyroid  influence on the toxicity of
     respiratory irritant  gases,  pzone and  nitrogen dioxide.   J. Pharmacol.
     Exp. Ther. 139:  177-134.

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

Fetner,  R.  H.   (1958) Chromosome  breakage  in Vicia  faba by ozone.   Nature
     (London) 181:  504-505.

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;  AD489014.

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

Fletcher, B. L.; Tappel, A. L. (1973)  Protective effects  of  dietary  alphatoco-
     pherol in  rats  exposed  to toxic  levels of  ozone  and nitrogen  dioxide.
     Environ. Res.  6: 165-175.
0190JE/A                          10-231                               5/1/84

-------
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. ;  Stephens,  R.  J. ;  Coffin, D.  L. ;  Stara, J.  F. (1973)  Changes  in
     dog's lungs after long-term exposure to ozone.  Arch. Environ.  Health  26:
     209-216.

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.
     01s. 106: 563-579.

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. ;  Holman,  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.

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

Fukase,  0. ;  Isomura,  K. ;  Watanabe, H. (1975)   Effect of  ozone on  glutathione
     in  vivo.  Taiki Osen 
-------
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
     Development  Administration;  pp. 1-21.    (ERDA  symposium  series:  43).

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.

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.

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

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.;  Traystman, R.;  Menkes,  H.  (1983c) Effects
     of ozone in peripheral  lung reactivity.   J.  Appl. Physiol.  Respir.  Environ.
     Exercise Physiol.  55:  777-784.

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.   Physiol. Respir.  Environ.  Exercise  Physiol. 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.
     Physiol. Respir.  Environ. Exercise Physiol.  54: 640-646.

Gildemeister, M.   (1921)  Uber Kampfgasvergiftungen.   In.   Laquer, E. ;  Magus,
     R. , eds.  Experimental  and  theoretical  bases for  the therapy  of phosgene
     sickness.   Z. Gesamte Exper. Med. 13: 284-296.
0190JE/A                          10-233                                5/1/84

-------
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, Ohio:   U.S.  Environmental  Protection Agency;  pp. 115-154;
     EPA 600/8-80-014.

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

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

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

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.   Jji vitro studies  utilizing the red cell  mem-
     brane.  Environ.  Res. 9:  179-186.

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

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

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

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

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

Goldstein,  E. ;  Gibson,  J.  B. ; Tassan, C. ;  Lippert,  W. (1978a)   Effect of
     levamisole on alveolar macrophage function.  Am.   Rev. Respir. Dis.  117:
     273.
 0190JE/A                           10-234                               5/1/84

-------
Goldstein, E.;  Tyler, W.  S.;  Hoeprich,  P.  D.;  Eagle,  C.  (1971a) Ozone and the
     antibacterial defense  mechanisms of the murine lung.   Arch.  Intern.  Med.
     128: 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, E. ;  Warshauer,  D.; Lippert,  W.; Tarkington,  B.  (1974) Ozone  and
     nitrogen dioxide exposure. Arch. Environ.  Health 28: 85-90.

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 in vitro exposures.   Environ.  Res. 12:  188-195.

Gordon,  T.;  Amdur,  M.  0.  (1980) Effects  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.

Graham,  J. A.  (1979)  Alteration of  hepatic xenobiotic  metabolism by  ozone.
     Durham, NC:  Duke University.   Ph.D. Dissertation.

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.;  Menzel,  D.  B.;  Mole,  M.  L.; Miller,  F.  J.;  Gardner,  D.  E.
     (1983b)  Influence  of  ozone on  pentobarbital  pharmacokinetics  in mice.
     Toxicol. Lett.   (In Press).

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. (1983a)   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).

Gregory, A.  R. ; Ripperton,  L.  A.;  Miller,  B. (1967)  Effect  of monatal  thymec-
     tomy on the development of ozone tolerance in mice.  Am.  Ind. Hyg. Assoc.
     J. 28:  278-283.

0190JE/A                          10-235                               5/1/84

-------
Grose, E.  C. ;  Gardner,  D.  E. ; Miller, F. J. (1980) Response of ciliated epi-
     theluim 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).

Gross, S.; Melhorn,  P.  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  jj}  vivo and i_n  vitro based  on
     sister  chromatid exchange analysis.  Environ. Res. 18: 336-346.

Hadley,  J.  G.; Gardner,  0.  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:   43).

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.  Part  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.  Part  II:  study of human  subjects  receiving supplemental vitamin  E.
     Environ. Res. 19:  299-305.

Hamelin,  C. ;  Chung,  Y.   S.  (1974)  Ozone  resistance in Escherichia coli.   I.
     Relations with some ONA  repair mechanisms = Resistance  a 1'ozone chez
     Escherichia coli.    I.  Relations avec  certain mecanismes  de  reparation de
     VADN.  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.


0190JE/A                         10-236                                5/1/84

-------
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  E 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.

Hammond,  P.  B.;  Beliles,  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).   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 tracheal  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.  Physiol.  Respir.  Environ. Exercise
     Physio!.  55: 1232-1236.

Holtzman, M.  J.; Fabbri,  L.  M.; O'Byrne,   P. M.; Gold,  B.  D.; Aizawa, H. ;
     Walters,  S.  E.; 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.
0190JE/A                          10-237                                5/1/84

-------
Hu, P.  C.;  Miller, F.  J. ; Daniels, M. J. ; Hatch, G. E. ; Graham, J. A.; Gardner,
     D.  E. ;  Selgrade, 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 jji vitro.  Arch.  Intern. Med. 127: 1059-1063.

Hurst,  D. J. ;  Gardner,  D. E.; Coffin,  D.  L.  (1970) Effect of ozone on acid
     hydrolases of the pulmonary alveolar macrophage.   Res. J. Reticuloendothel.
     See.  8: 288-300.

Hussain,  M.  Z.  ; Cross, C.  £.; 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.

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.

Hyde, D. ; Orthoefer,  J.;  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—observations 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 N0? and 0-.  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.
0190JE/A                          10-238                                5/1/84

-------
Ishidate, M. ,  Jr.,; Yoshikawa,  K.  (1980) Chromosome  aberration  tests with
     Chinese hamster cells  j_n vitro with and without  metabolic activation--a
     comparative study  on mutagens  and carcinogens.   Arch.  Toxicol.  4: 41-44.

Jegier,  Z.  (1973)   Ozone as  an air pollutant.   Can.  J.  Public Health  64:
     161-166.

Johnson, D.  A.  (1980) Ozone inactivation of human alpha 1-proteinase inhibitor.
     Am. Rev. Respir. Dis.  121: 1031-1038.

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 1974
     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) jLn vivo
     damage of  rat  lungs by oxygen metabolites.   Clin.  Invest.  67:  983-993.

Juhos,  L. T. ;  Evans,  M. J.  ; Missenden-Harvey, R.; Furiosi,  N. J.; Lapple,  C.
     E.; Freeman, G. (1978) Limited exposure of rats to H SO  with and without
     0,.  J. Environ. Sci.  Health C13:  33-47.
      i
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. Pharmacol.  48:
     18-28.

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 eind cellular  aspects of carcinogen screening tests;
     June 1979;  Hanover, V/est 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.  Am. Rev. Respir. Dis. 93:  93-102.

Kesner,  \f. ;+Kindya, R.  J. ; Chan, P.  C.  (1979) Inhibition of erythrocyte  membrane
     (Ma +K )-activated  ATPase  by ozone-treated phospholipids. J.  Biol.  Chem.
     254: 2705-2709.
0190JE/A                          10-239                                5/1/84

-------
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 gycloperoxidase-tnediated  system.   Blood  45:  699-707.

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. ;  jjjea^h,  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.

Kulle, T.  J. ; Cooper,  G.  P.   (1975)  Effects of  formaldehyde  and ozone  on  the
     trigeminal   nasal  sensory  system.   Arch.  Environ.  Health  30:  237-243.

Kyei-Aboagye, K.; Hazucha, M. ;  Wyszogrodski,  I.;  Rubinstein,  D.;  Avery, M.  E.
     (1973)  The  effect  of  ozone exposure jj}  vivo on the appearance of lung
     tissue  lipids  in the endobronchial  lavage  of rabbits.   Biochem. Biophys.
     Res.  Commun. 54: 907-913.

La Belle,  C.  W.; Long, J. E.;  Christofano, E.  E. (1955) Synergistic effects of
     aerosols.   Particulates  as carriers of  toxic  vapors.  AMA Arch.  Ind.
     Health  11:  297-304.

Landahl, H.  0. (1950) On the removal of air-borne droplets by the  human respira-
     tory tract.  I.  The lung.  Bull.  Math. Biophys.  12:  43-56.

Larkin, E.G.; Goheen, S.C.; Rao, G.A.  (1983)  Morphology  and  fatty acid compo-
     sition   of  erythrocytes   from monkeys  exposed  to ozone for  one year.
     Environ. Res. 32: 445-454.

Larkin, E.  C. ; Kimzey, S. I..;  Siler, K. (1978) Response of the  rat erythrocyte
     to ozone exposure.  J. Appl. Physio!.  Respir.  Environ.  Exercise Physiol.
     45: 893-898.

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.
0190JE/A                          10-240                                5/1/84

-------
Last, J. A.; Kaizu, T. (1980) Mucus glycoprotein  secretion by trachea!  explants:
     effects of pollutants.  EHP Environ.  Health  Perspect.  35:  131-138.

Last, J. A.;  Greenberg,  D. B. ; Castleman, 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.; Jennings, M.  0.; Schwartz, L. W.;  Cross,  C.  E.  (1977) Glycoprotein
     secretion by  trachea! explants  cultured from rats exposed to ozone.  Am.
     Rev. Respir. Dis. 116: 695-703.

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
     dog.  J. Appl. Physiol. Respir. Environ.  Exercise Physiol.  46:  1108-1109.

Lee, L.-Y.; Djokic, T. D.; Dumont, C.; Graf, P. D.; Nadel,  J. A.  (1980)  Mechan-
     ism 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.

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 beagla.  Arch. Environ. Health 29:102-106.

Ling, N. R.; Kay, J. F.   (1975)  Lymphocyte  Stimulation.   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.; Niegi ,  D. ;  Stephens,  R. L.  (1977) Glucose-6-phosphate
     dehydrogenase response of  postnatal   lungs to NO  and 0  .   In:   Sanders,
     C.   L.;  Schneider,  R.  P.;  Dagle,  G.  E. ;  Ragan, R.  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).

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.


0190JE/A                          10-241                               5/1/84

-------
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.  Physiol.  Respir.
     Environ. Exercise Physiol. 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. 01s. 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  disorders in guinea pigs.   IV.
     Effects  on  respiratory sensitivity to  inhaled acetylcholine.    Am.  Rev.
     Respir.,Dis. 105: 262-267.

Matzen, R.  N.  (1957a) Development of tolerance  to ozone  in reference to pul-
     monary  edema.  Am. J. Physiol. 190: 84-88.

Matzen, R.  N.  (1957b) Effect of vitamin C and hydrocortisone on the pulmonary
     edema produced by ozone in mice.   J.   Appl.  Physiol.  11: 105-109.

McAllen,  S.  J. ;  Chiu, S.  P.; Phalen,  R. F. ;  Rasmussen, R. E. (1981) Effect of
     i_n vivo  ozone exposure  on  i_n vitro pulmonary alveolar  macrophage mobility.
     J. Toxicol.  Environ.  Health 7: 373-381.

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 1972; San  Francisco, CA.   Am.  Ind.  Hyg.  Assoc.  J.
     33(2):  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.   Bull.  Int.  Union Against  Tuberculosis  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.  Ann.   Rev.
     Pharinacol.  10:  379-394.
 0190JE/A                           10-242                               5/1/84

-------
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. II.  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. ;  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.  Dietary vitamin E and  fat:
     effect on  murine lung and  red blood cell glutathione peroxidase activity.
     FASEB Meeting.  1978.

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.  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.

Menzel, D. B.;  Slaughter, R.; Donavan, D.; Bryant, A. (1973) Fatty acid ozonides
     as the toxic agents on ozone inhalation.  Pharmacologist 15:  239.

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,
     D.C.:   U.S.  Environmental Protection Agency,  Office  of Research and
     Development; EPA-600/8-83-002.

Miller, F. J. (1977) A mathematical model  of transport  and  removal of  ozone in
     mammalian  lungs.   Raleigh, NC:   North  Carolina State  University; Ph.D.
     thesis.

Miller, F.  J.  (1979) Biomathematical modeling applications in the evaluation
     of ozone toxicity:  ch. 14.  In:  Lee,  S. D.; Mudd, J. B.,  eds.   Assessing
     toxic effects of environmental pollutants.  Ann Arbor, MI:  Ann Arbor  Science
     Publishers, Inc.; pp.  263-286.

Miller, F.  J. ;  Illing,  J.  W.;  Gardner,  D.  E.  (1978) Effect of  urban  ozone
     levels  on  laboratory-induced respiratory  infections.   Toxicol.  Lett.
     2: 163-169.
0190JE/A                          10-243                               5/1/84

-------
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, S. ;  Ehrlich,  R.  (1958)  Susceptibility  to respiratory infections  of
     animals exposed to ozone.  J. Infect. Dis. 103:  145-149.

Mizoguchi, 1. ;  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.

Montgomery,  M.  R. ;  Niewoehner, D. E.  (1979)  Oxidant-induced alterations  in
     pulmonary microsomal mixed-function oxidation:   acute  effects  of  paraquat
     and ozone.  J.  Environ. Sci. Health C13: 205-219.

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 j_n 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, 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.

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. ; Ongun, A.;  McManus,  T. T.  (1969) Reaction  of ozone
     with amino  acids and  proteins.  Atmos. Environ.  3:  669-681.

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.


0190JE/A                          10-244                                5/1/84

-------
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.

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  mitochondrial  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.; 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. ;  Ospital,  J. J. ;  Hacker,  A.  D.   (1976) Effects of ozone and
     nitrogen  dioxide exposure  on lung  metabolism.   EHP Environ.  Health
     Perspect. 16: 184 (Abstract).

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,  NO:
     Princeton  Scientific  Publishers,  Inc.; pp. 57-73.   (Advances in modern
     environmental toxicology:  v. 5).

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. ; 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.; 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.
0190JE/A                          10-245                                5/1/84

-------
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.  Toxicol.  Appl. Pharmacol. 72: (in press).

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.; Ostu,  H. (1972).   Histopathological changes in the
     respiratory organs  of mice exposed to photochemical  oxidants and  automo-
     bile  exhaust gas.  Osaka-Furitsu Koshu Eisei Kenkyusho Kenkyu Hokoku  Rodo
     Eisei  Hen 10:35-42.

Nasr,  A. N. M.  (1967) Biochemical  aspects of ozone intoxication:   a review.
     J. Occup. Med. 9: 589-597.

National Research  Council. (1977)  Drinking water and health.   Washington,  DC:
     National  Academy of Sciences.

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;  PB190262.

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 Research  Council.  (1977) Toxicology:   ch.  8.   In:   Ozone  and
     other  photochemical  oxidants.   Washington, DC:    National  Academy of
     Sciences,  Committee  on  Medical  and Biologic Effects of  Environmental
     Pollutants; pp. 323-387.

Niewoehner, D.  E. ;  Kleinerman,  J.; Rice,  D.   B.  (1974)  Pathologic changes in
     the peripheral  airways  of young  cigarette  smokers.   N.   Engl.  J.  Med.
     2S1:  755-758.

North Atlantic Treaty Organization. (1974) Air quality  criteria for photochemi-
     cal 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.

Oberst,  F.   W. ;  Comstock,  C.  C.; Hackley, E.  B. (1954)  Inhalation toxicity of
     ninety percent  hydrogen peroxide  vapor.   AMA  Arch. Ind.  Hyg.  Occup.  Med.
     1C: 319-327.

Oomichi, 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. ; 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.
0190JE/A                          10-246                                5/1/84

-------
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.

Osebold, J. W. ; Owens, S. ;_. ; Zee, Y.  C. ; Dotson, W. M. ;  Labarre,  D.  D.  (1979)
     Immunelogical alterations  in the  lungs  of mice  following ozone exposure:
     changes  in  immunoglobulin  levels  and antibody-containing  cells.   Arch.
     Environ.  Health 34: 258-265.

Pack, A.;  Hooper,  M.  B. ; Nixon, W.; and 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.  ; Exley, R. W.; Coffin, D. L. (1972)  Influence  of pollutant  gases
     on benzpyrene  hydroxylase  activity.   Arch. Environ.  Health 25:  439-442.

Palmer, M. S.; Swanson, D. H.; Coffin, D.  L.  (1971)  Effect of 0   on benzpyrene
     hydroxylase  activity in the  Syrian  golden  hamster.  Cancer  Res. 31:
     730-733.

P'an, A.  Y. S.;  Jegier,  Z.  (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;  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.

Penha,  P.  D.; Werthamer, S.  (1974)  Pulmonary lesions  induced  by long-term
     exposure to ozone.  II.  Ultrastructure  observations of pro!iferative  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,  0.; Harder, S.  (1978a) The effect of ozone
     on human immunity:  i_rt vitro responsiveness of  lymphocytes  to  phytohemmag-
     glutinin.  Arch.  Environ. Health  36:  59-63.

Peterson, M. ; Harder,  S. ;  Rummo, N. ;  House,  D.  (1978b)  Effect  of  ozone  on
     leukocyte function in exposed human subjects.   Environ. Res.  15:  485-493.
0190JE/A                          10-247                               5/1/84

-------
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.

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 reexposure.   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)  Effects of  sodium hypochlorite, ozone,
     and ionizing  radiations on pyrimidine  constituents of Escherichia coli.
     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.;  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.  Interna-
     tional  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).

Pryor, W.  A.;  Stanley, J. P.; Blair, E;  Cullen, G. B.  (1976) Autoxidation  of
     polyunsaturated fatty acids.   I.  Effect of ozone  on the  autoxidation of
     neat  methyl  linoleate  and  methyl  linolenate.  Arch. Environ. Health 31:
     201-210.

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  Princeton,  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.

Razumovskii,  S. D.; Zaikov, G.  E. (1972) Effect  of structure of an unsaturated
     compound  on  rate  of its reaction with  ozone.   J.  Gen.  Chem. (USSR)  8:
     468-472.

Reasor, M. J.;  Adams III, G. K.;  Brooks, J. K.;  Rubin,  R. J. (1979)  Enrichment
     of albumin and IgG in the  airway  secretions of  dogs breathing ozone.   J.
     Environ.  Sci. Health C13:  335-346.

0190JE/A                          10-248                                5/1/84

-------
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. (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;
     8 pp; Research Project S-6.

Roehm, J.; Hadley, J. G.; Menzel,  D.  B. (1971a) Antioxidants vs. lung  disease.
     Arch. Intern. Med. 128: 88-93.

Roehm, J. N. ,  Hadley,  J.  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.

Roycroft, J.  H. ;  Gunter, W. B.;  Menzel,  D. B.  (1977)  Ozone  toxicity:  hormone-
     like oxidation  products  from  arachidonic  acid by ozone-catalyzed autoxi-
     dation.   Toxicol.  Lett. 1:  75-82.      (,.
                                            &<
Sachsenmaier,  W. ;  Siebs,  W. ; Tan,  T.  A.  (1t965) Wirkung  von  Ozon  auf
     Mauseacitestumorzellen und  auf  Huknerfibroblasten in  der  Gewebekultur.
     [Effect of ozone on mouse ascites 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.  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.
0190JE/A                          10-249                               5/1/84

-------
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.;  Hirosa,  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.

Sato, S.; Shimura,  S.;  Kawakami, M. ; Hirosa, 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-depleted rats.  Nippon Kyobu Shikkan Gakkai Zasshi  16: 260-268.

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.

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. Physiol. 38: 719-723.

Schlipkoter, H. ;  Bruch, J.  (1973) Functional  and morphological alterations
     caused by exposure to ozone.  Zentralbl. Bakteriol.  Parasitenkd.  Infektion-
     skr.  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-
     mins.  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  Escherichia coli.  J.  Bacteriol.  85:  567-576.
0190JE/A                          10-250                                5/1/84

-------
Belgrade, 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.  (1974) Nutritional influences bn  the toxicity of  environmental
     pollutants.  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-  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).

Shingu, H.  ; Sugiyama, M. ; Watanabe, M.;  Nakajima, T. (1980) Effects of ozone
     and photochemical  oxidants on  interferon production  by  rabbit alveolar
     macrophages.  Bull. Environ. Contain.  Toxicol. 24:  433-438.

Sielczak, M. W. ;  Denas, S.  M.; Abraham,  W.  M.  (1983)  Airway  cell  changes in
     trachea!  lavage of sheep after ozone  exposure.  J.  Toxicol.  Environ.  Health
     11: 545-553.

Sparrow, A. H. ;  Schairn,  L. A. (1974)  Mutagenic response of  Tradescantia  to
     treatment with X-rays,  EMS, DBE, ozone, S02,  N20 and  several insecticides.
     Am. Environ. Mutat. Soc. 445

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. ;  Sloan, M.  F. ; Groth, D.  G. (1976)  Effects of long-term,  low
     level   exposure of  N0?  or 0~  on rat  lungs.   EHP Environ.  Health Perspect.
     16: 178-179.

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.   J.; Freeman, G. (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 0~  for  short  periods.   Exp.  Mol.
     Pathol. 20:  11-23.
0190JE/A                          10-251                                5/1/84

-------
Stephens, R. J. ; Sloan, M. F.; Groth, D. G.; Negi, D.  S.;  Lunan,  K.  D.  (1978)
     Cytologic responses of postnatal rat lungs to 0_ or N0_ 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 (1981)  Hydrogen
     peroxide contracts airway smooth muscle:  a possible endogenous mechanism.
     Respir. Physio!.  45:  333-342.

Stich, H. F.; Wei,  L.; Lam, P. (1978) The need for a mammalian test  in Chinese
     hamster ovary cells  by  Thiol  and hydrazene  compounds.   Mutat.  Res.  68:
     351-365.

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.; Wagner, W.  D.;  Wright, P. G.  (1956) Studies on  ozone toxicity.
     I.  Potentiating effects of exercise and  tolerance  development.  AMA  Arch.
     Ind. Health 14:  158-1S2.

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.   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.  Arch. Environ.  Health 12:698-704.

Sweet,  F. ;  Koo, M.-S. ; Lee,   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 phosphatidyl-
     choline  liposomes and the lytic effect of  products on red blood cells.
     Chem.  Phys. Lipids 12:  153-171.

Tepper,  J.  L. ;  Weiss, B.; Cox, C.  (1982) Microanalysis of ozone depression of
     motor  activity.   Toxicol. Appl.  Pharmacol.  64:  317-326.
0190JE/A                           10-252                                5/1/84

-------
Tapper,  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. ;  Fenters,  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; PB292267.

Thomas,  G. B.;  Fenters, 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.;  Fenters, J.  D.; Ehrlich,  R.;  Gardner, D. E.  (1981b) Effects of
     exposure to ozone  on susceptibility to experimental  tuberculosis.  Toxicol.
     Lett. 9: 11-17.

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 0-.   Arch.  Environ.  Health 24:
     153-159.                                   *

Tremer,  H.  M. ; Falk.  H. L. ;  Kotin, P.  (1959) Effect of  air pollutants  on
     ciliated  mucous  secreting  epithelium.    JNCI  J.  Natl.  Cancer Inst.
     23: 979-994.

Tyson, C.  A.;  Lunan,  K.  D.; Stephens,  R. J.  (1982)  Age-related differences in
     GSH-shuttle  enzymes  in  NO,,-  or  0_-exposed rat  lungs.   Arch.  Environ.
     Health 37: 167-176.       *      J

U.S. Department of Health,   Education and Welfare  (1967)   The  health consequences
     of  smoking.   Public Health  Service  Publication No. 1696, Revised  January
     1968.   Library of Congress  Catalog  No. 68-60025.

U.S. Department of Health,   Education,  and  Welfare (1969)  The  health consequen-
     ces of  smoking,  1969   Supplement.   Supplement  to Public Health Service
     Publication No. 1696-2.   Library  of Congress Catalog No.  68-60025.

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; PB80-124753.
0190JE/A                          10-253                               5/1/84

-------
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-600/8-82-029a.

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.

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;
     Wcishington, DC.  Washington, DC:   International Union of Air  Pollution
     Prevention Associations; paper no.  MB-15E.

Veninga, T.;  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 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.

Wegner, C.  D.   (1982) Characterization of  dynamic respiratory mechanics by mea-
     surement  of pulmonary and respiratory  system  impedance  in  adult bonnet
     monkeys  (Macaca radiata):  including the effects of long-term  exposure  to
     low-level  ozone.  Davis, CA: University of  California; Ph.D. Dissertation.

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.
0190JE/A                          10-254                                5/1/84

-------
Wilkins, E.  S. ;  Fettissoff,  P. (1981) The  effect of air pollutants  on  lung
     surfactant surface tension.  J. Environ. 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.  S. ;  Calabrese,  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.  (In press).

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  a-irradiated nucleic acid consti-
     tuents  in CHO cells.   Mutat. Res. 88: 99-107.

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. ;  Mehlrnan,  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).

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.


0190JE/A                          10-255                               5/1/84

-------
Yokoyama, E. (1969)  A  comparison of the effects  of SO ,  N0? and  0   on  the
     pulmonary ventilation.  Guinea  pig exposure experiments.  Sangyo Igaku
     11:  563-568.

Yokoyama, E. (1972)  The  effects of 0,  on  lung functions.  In: Proceedings  of
     the science  lecture meeting on  air pollution caused by photochemical
     reaction;  Japan Society of Air Pollution, Tokyo; March 1972; 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.  (1972) 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 (FY  1972-1975).  Tokyo,  Japan:   Institute of
     Public Health, Department of Industrial Health; pp. 16-1 - 16-6.

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, Jr., W. E.; 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.
0190JE/A                          10-256                                5/1/84

-------
            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
     tissue^ 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 0,? 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 urn)  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  (PACQ?):   Partial  pressure of carbon  dioxide
     in the air contained in the lung alveoli.

Alveolar  oxygen  partial  pressure (P/iOo)-  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.
019CC/C                            A-l                           May 1984

-------
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 (VQ     .):   Volume of  the conducting airways down to the
     level where, during enr Breathing,  gas  exchange with blood can occur, a
     region probably situated at the entrance of the alveolar ducts.

Arterial oxygen  saturation (SaCk):   Percent saturation of  dissolved oxygen  in
     arterial blood.

Arterial  partial  pressure  of carbon dioxide  (PaCO^):   Partial pressure  of
     dissolved carbon dioxide in arterial blood.

Arterial  partial  pressure  of oxygen (PaO?):    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.

Atropirie:    A poisonous white crystalline alkaloid,  C,7H,,~N03,  from belladonna
     arid  related  plants,  used  to relieve spasms of smootn 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.,
019CC/C                            A-2                            May 1984

-------
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-
     ogy, 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 tne 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 (carbamoylcholine chloride, CgH^ClN-O-)
     that produces constriction of the bronchial smooth muscles.

Carbon dioxide production (VCO-):  Rate of carbon dioxide production by organ-
     isms,  tissues, or cells.  Common units:   ml CO™ (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?.
019CC/C                          ,  A-3                           May 1984

-------
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).

Closinc; capacity (CC):  Closing  volume  plus  residual  volume,  often expressed
     as a ratio of TLC,  i.e.  (CC/TLC%).

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 VENTILATION  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  O'c 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  H^O or ml/cm H^O.  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, D|CO~, D.CO):  Amount of gas (0,,,
     CO, CO,,)  commonly  expressed as  mi  gas CSTPu)  diffusing between alveofar
     gas ana pulmonary  capillary blood  per torr  mean gas pressure difference
     per min,  i.e.,  ml  02/(min-torr).   Synonymous with  transfer  factor and
     diffusion factor.

Dynamic compliance (C,  ):  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.
019CC/C                            A-4                           May 1984

-------
Elastance (E):  The reciprocal of COMPLIANCE;  expressed  in  cm  H?0/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.

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 FEV./FVC.  It is an index of airway obstruction.
                       t>
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.

          FEF? _   nn = mean forced expiratory flow  between 200 ml
                   U    and 1200 ml of the FVC (formerly called the
                        maximum expiratory flow rate  (MEFR).

          FEF?C) 7C.% = mean forced expiratory  flow during the middle
                       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  FEV0 _>
     FEV_ 7t., FEV, „.   These  values are often  expressed  as a  percent of the"
     forced vital capacity, 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.

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.


019CC/C                            A-5                           May 1984

-------
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.

Hematocrit (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 heme+ group.  The heme is made of four pyrrole rings
     and a divalent iron (Fe   -protoporphyrin) which combines reversibly with
     molecular oxygen.

Histamine:   A depressor  airiine  derived  from the ami no 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 P-? is  low  in the environment, whether because of
     decreased barometric  pressure or decreased fractional concentration of
     0,,, the condition is termed environmental hypoxia.  Hypoxia when referring
     to  the  blood is  termed hypoxemia.  Tissues are said to  be  hypoxic when
     their P._ 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.

Kilogram-meter/min (kgm/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.

019CC/C                            A-6                           May  1984

-------
Maximal aerobic  capacity  (max V0~):  The  rate  of oxygen uptake by the body
     during  repetitive  maximal  respiratory effort.  Synonymous with  maximal
     oxygen consumption.

Maximum breathing capacity (MBC):  Maximal volume  cf 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 maximun voluntary ventilation (MVV).
                          »
Maximum expiratory  flow  (V      ):   Forced expiratory flow, related  to  the
     total lung capacity  o:" 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    -,.-0, = instantaneous forced  expiratory flow when the
           max 75%   ,    .   .   -,..<*  f • *.  -riV
                     lung is at  75% of its TLC.
          •
          V    o n = instantaneous forced  expiratory flow when the
           max i.u        volume  is 3 0 liters
Maximum expiratory flow rate (MEFR):  Synonymous with FEF  ~       .

Maximum mid-expiratory  flow rate (MMFR or MMEF):  Synonymous with FEF?c._7C.y.

Maximum ventilation (max Vc):  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 (VV):   Volume  of  air breathed in one  minute.  It  is  a
     product of  tidal  volume  (VT)  and breathing frequency  (fD).   See VENTILA-
     TION.                      '                             B

Minute volume:   Synonymous with minute ventilation.

Mucociliary transport:  The process by which mucus is transported, by  ciliary
     action, from the lungs.

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).
019CC/C                          .  A-7                           May 1984

-------
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 N0? 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 breatrnng ambient air to an inspired mixture of pure 0?.   A
     progressive decrease  of  N« concentration ensues which  may  be analyzed
     into two or more  exponential  components.  Normally,  after 4 min of pure
     ()„ breathing the  fractional N? concentration in expired alveolar gas is
     down to less than 2%.

Normoxia:   A state in which the ambient oxygen pressure is approximately 150 ±
     1C 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 (V0_,  Q0?):   Rate  of oxygen uptake of organisms, tissues,
     or cells.  Common  unit?:   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 Qft_ = (jl 0?/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 (VO ):  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 02 stores
     are constant.

Participates:  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.

Physiological dead  space  (Vn):  Calculated  volume which accounts for  the
     difference between the pressures of C0? 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.


019CC/C                            A-8                            May  1984

-------
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 (SiIverman-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 enlargement of  the air spaces  distal to the
     terminal nonrespiratory bronchiole,  accompanied by  destructive  changes of
     the  alveolar walls.   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.

Resistance flow  (R):   The ratio of the flow-resistive components of pressure
     to simultaneous flow, in cm HpO/liter per sec.  Flow-resistive  components
     of pressure are obtained by subtracting  any  elastic or inertial 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.
019CC/C                          ;  A-9                           May 1984

-------
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 (fn):  The number of breathing cycles per  unit of time.
     Synonymous with breathing frequency (fg)-

Respiratory quotient  (RQ, R):  Quotient of the volume of CO  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 C0? output  to  the 0» uptake  by  the  lungs,  "respiratory
     quotient" being  restricted to the actual metabolic  CO-  output and 0
     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 0? and C0? stores.

Shunt:   Vascular  connection between  circulatory  pathways so that venous blood
     is diverted  into vessels containing  arterialized  blood  (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  nalformations
     are more  important  in  respiratory physiology.   Flow  from  ls?ft 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.

Static lung compliance (C   ):  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., ^  TLC is static recoil pressure measured at
     TLC (maximum recoil pressure).

Sulfur dioxide (S0?):  Colorless gas with pungent odor, released primarily  from
     burning of fossil fuels,  such as  coal, containing sulfur.
019CC/C                            A-10                          May 1984

-------
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..):    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
     barometer 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.

Total pulmonary resistance (R.):  Resistance measured by relating flow-dependent
     transpulmonary pressure to airflow at the mouth.   Represents the  total
     (frictional) resistance of the lung tissue (R..) and the airways (Raw).
     R.  = R   + R+..
      L    aw    ti

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.
019CC/C                            A-11                          May 1984

-------
Transpulmonary pressure  (P.):   Pressure  difference  between airway opening
     (mouth, nares, or  cannula  opening)  and the visceral ploural surface, in
     cm HLO.  Transpulmonary in the sense used  includes  extrapulnonary  s :ruc~
     tures, e.g.,  trachea and extrathoracic  airways.   This usag? 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 thj lungs.  The
     word ventilation sometimes designates  ventila~:ory flow rate  (or  ven\ila-
     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
               Vp = 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 vtntilatijn is
     less than total  ventilation because when a tidal volume of gas leaves the
     alveolar spaces, the last  part  does not get  expelled  from  1 he body but
     occupies the  dead  space,  to be  reinspired with the next ir spirati :>n.
     Thus the volume of alveolar gas actually expelled completely is  equal  to
     the tidal volume minus  the volume of the dead space.  This t^uly conplete
     expiration volume times the ventilatory frequency constitutes the aiveolar
     ventilation.
                          •
Ventilation, dead-space  (Vn):  Ventilation  per  minute of the  physiologic dead
     space  (wasted  ventilation),  BTPS,  defined by the  following equation:

          VD = V£(PaC02 - ?£C02)/(PaC02 - PjCOg)

Ventilation/perfusion ratio (V./Q):  Ratio  of the  alveolar  venti  ation to  th;
     blood perfusion volume flow through the pulmonary parenchyma.  This rat o
     is a  fundamental determinant  of the Q~ and C0» pressure of  the  alveolar
     gas  and  of the  end-capillary  blood.  Throughout the  lungs the  looal
     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 tha point  of
     maximum inspiration.
          U.S. Environmental Protection Agency
          Region V, Library
          230 South Dearborn Street
          Chicago,  Illinois  60604
019CC/C                            A-12                           ^lay 1984

-------