United States
           Environmental Protection
           Agency
                        EPA 6CX 8-82 OObT
                        Jtilv 19fa
                        F null Re}' it
            Research and Development
vvEPA
Health Assessment   Final
Document for         Report
Trichloroethylene

-------
                                  EPA/600/8-82/006F
                                          July 1985
                                         Final Report
Health Assessment Document
                   for
         Trichloroethylene
       U.S. ENVIRONMENTAL PROTECTION AGENCY
          Office of Research and Development
       Office of Health and Environmental Assessment
       Environmental Criteria and Assessment Office
           Research Triangle Park, NC 27711

-------
                                   NOTICE
     This document has been reviewed in accordance with the U.S.  Environmental
Protection Agency's peer and administrative review policies and approved for
presentation and publication.   Mention of trade names or commercial products
does not constitute endorsement or recommendation for use.

-------
                                    PREFACE

     The Office of Health and Environmental Assessment,  in consultation with
an Agency work group,  has prepared this health assessment to serve as a "source
document" for EPA use.   Originally the health assessment was developed for use
by the Office of Air Quality Planning and Standards;  however, at the request
of the Agency Work Group on Solvents, the assessment  scope was expanded to
address multimedia aspects.
     In the development of the assessment document,  the  scientific literature
has been inventoried,  key studies have been evaluated, and summary/conclusions
have been prepared so that the chemical's toxicity and related characteristics
are qualitatively identified.  Observed effect levels and dose-response rela-
tionships are discussed, where appropriate, so that the  nature of the adverse
health responses is placed in perspective with observed  environmental levels.
     Any information regarding sources, emissions, ambient air concentrations,
and public exposure has been included only to give the reader a preliminary
indication of the potential  presence of this substance in the ambient air.
While the available information is presented as accurately as possible, it is
acknowledged to be limited and dependent in some instances on assumption
rather than specific data.  This information is not intended, nor should it be
used, to support any conclusions regarding risks to public health.
     If a review of the health information indicates  that the Agency should
consider regulatory action for this substance, a considerable effort will  be
undertaken to obtain appropriate information regarding sources, emissions, and
ambient air concentrations.   Such data will provide additional information for
drawing regulatory conclusions regarding the extent and  significance of public
exposure to this substance.
                                      iii

-------
                                   CONTENTS
LIST OF TABLES 	     vii
LIST OF FIGURES 	     x
AUTHORS AND REVIEWERS 	     x1i

1.  EXECUTIVE SUMMARY 	     1-1

2.  INTRODUCTION 	     2-1

3.  BACKGROUND INFORMATION 	     3-1
    3.1  PHYSICAL AND CHEMICAL PROPERTIES 	     3-1
    3.2  ENVIRONMENTAL FATE AND TRANSPORT 	     3-3
         3.2.1  Production 	     3-3
         3.2.2  Use 	     3-4
         3.2.3  Emissions 	     3-5
         3.2.4  Persistence 	     3-6
         3.2.5  Degradation 	     3-7
    3.3  LEVELS OF EXPOSURE 	     3-8
         3.3.1  Analytical Methodology	     3-8
         3.3.2  Calibration 	     3-12
         3.3.3  Sampling and Sources of Error	     3-12
    3.4  ECOLOGICAL EFFECTS 	     3-24
    3.5  CRITERIA, STANDARDS, AND REGULATIONS 	     3-26
    3.6  REFERENCES 	     3-27

4.  PHARMACOKINETICS AND METABOLISM 	     4-1
    4.1  ABSORPTION AND DISTRIBUTION 	     4-1
         4.1.1  Dermal Absorption	     4-1
         4.1.2  Oral Absorption 	     4-2
         4.1.3  Pulmonary Absorption 	     4-5
         4.1.4  Tissue Distributions and Concentrations 	     4-10
    4.2  EXCRETION 	     4-14
         4.2.1  Pulmonary Elimination in Man 	     4-14
         4.2.2  Urinary Metabolite Excretion in Man  	     4-17
         4.2.3  Excretion Kinetics in the Rodent  	     4-18
    4.3  MEASURES OF EXPOSURE AND BODY BURDEN 	     4-23
    4.4  METABOLISM	     4-26
         4.4.1  Known Metabolites	     4-26
         4.4.2  Magnitude of TCI Metabolism:  Evidence of Dose-
                 Dependent Metabol ism 	     4-29
         4.4.3  Enzyme Pathways of Biotransformation 	    4-39
         4.4.4  Metabolism and Covalent Binding	    4-48
         4.4.5  TCI Metabolism:  Drug and Other Interactions  	    4-53
         4.4.6  Metabolism and Cellular Toxicity	    4-58
    4.5  REFERENCES  	    4-63
                                        iv

-------
                             CONTENTS (continued)
5.   TOXICOLOGICAL EFFECTS IN MAN AND EXPERIMENTAL ANIMALS 	     5-1
    5.1  INTRODUCTION 	     5-1
    5.2  NEURAL AND BEHAVIORAL EFFECTS 	     5-1
         5.2.1  Human Studies 	     5-1
         5.2.2  Laboratory Animals 	     5-5
    5.3  EFFECTS ON THE CARDIOVASCULAR AND RESPIRATORY SYSTEMS ...     5-10
         5.3.1  Cardiovascular Effects 	     5-10
         5.3.2  Respiratory Effects	     5-13
    5.4  HEPATIC AND RENAL TOXICITY 	     5-14
         5.4.1  Human Studies 	     5-14
         5.4.2  Animal Studies 	     5-15
    5.5  IMMUNOLOGICAL AND HEMATOLOGICAL EFFECTS 	     5-21
    5.6  DERMAL EFFECTS 	     5-23
    5.7  MODIFICATION OF TOXICITY:  SYNERGISM AND ANTAGONISM	     5-23
    5.8  SUMMARY OF TOXICOLOGICAL EFFECTS 	     5-27
    5.9  REFERENCES 	     5-28

6.   TERATOGENICITY, EMBRYOTOXICITY, AND REPRODUCTIVE EFFECTS 	     6-1
    6.1  ANIMAL STUDIES 	     6-2
         6.1.1  Mouse 	     6-2
         6.1.2  Rat	     6-4
         6.1.3  Rabbit 	     6-8
    6.2  FETOTOXICITY IN HUMANS 	     6-10
    6.3  SUMMARY	     6-10
    6.4  REFERENCES 	     6-12

7.   MUTAGENICITY 	     7-1
    7.1  GENE MUTATION STUDIES 	     7-1
         7.1.1  Prokaryotic Test Systems (Bacteria) 	     7-1
         7.1.2  Eukaryotic Test Systems 	     7-9
    7.2  CHROMOSOME ABERRATION STUDIES 	     7-20
         7.2.1  Experimental Animals 	     7-21
         7.2.2  Humans 	     7-24
    7.3  OTHER STUDIES INDICATIVE OF MUTAGENIC ACTIVITY  	     7-26
         7.3.1  Gene Conversion in Yeast 	     7-26
         7.3.2  Sister Chromatid Exchange (SCE) Formation 	     7-27
         7.3.3  Unscheduled DNA Synthesis (UDS)	     7-29
    7.4  EVIDENCE THAT TCI REACHES THE GONADS 	     7-33
    7.5  MUTAGENICITY OF METABOLITES 	     7-34
    7.6  BINDING TO DNA 	     7-36
    7.7  SUMMARY AND CONCLUSIONS  	     7-36
    7.8  REFERENCES 	     7-44

-------
                             CONTENTS (continued)
8.   CARCINOGENICITY	     8-1
    8.1  DESCRIPTION AND ANALYSIS OF ANIMAL STUDIES AND CELL
         TRANSFORMATION STUDIES 	     8-1
         8.1.1  Oral Administration (Gavage):   Rat 	     8-1
         8.1.2  Oral Administration (Gavage):   Mouse 	     8-20
         8.1.3  Inhalation Exposure:  Rats 	     8-41
         8.1.4  Inhalation Exposure:  Hamsters 	     8-47
         8.1.5  Inhalation Exposure:  Mice 	     8-48
         8.1.6  Other Carcinogenicity Evaluations of TCI	     8-55
         8.1.7  TCI Oxide 	     8-55
         8.1.8  Cell Transformation Studies 	     8-61
         8.1.9  Summary of Animal and Cell Transformation
                Studies 	     8-69
    8.2  DESCRIPTION AND ANALYSIS OF EPIDEMIOLOGIC STUDIES 	     8-72
         8.2.1  Axel son et al.  (1978) 	     8-72
         8.2.2  Tola et al.  (1980) 	     8-75
         8.2.3  Malek et al.  (1979) 	     8-78
         8.2.4  Novotna et al.  (1979) 	     8-78
         8.2.5  Hardel 1 et al.  (1981) 	     8-79
         8.2.6  Paddle (1983) 	     8-80
    8.3  RISK ESTIMATES FROM ANIMAL DATA 	     8-81
         8.3.1  Selection of Animal Data Sets  	     8-82
         8.3.2  Interspecies Dose Conversion 	     8-84
         8.3.3  Choice of Risk Model	     8-108
         8.3.4  Calculation of Unit Risk	     8-115
    8.4  RISK ESTIMATION FROM EPIDEMIOLOGIC DATA	     8-126
         8.4.1  Selection of Epidemiologic Data Sets 	     8-126
         8.4.2  Description of Risk Model 	     8-127
         8.4.3  Calculation of Upper Limits of Risk	     8-128
    8.5  RELATIVE CARCINOGENIC POTENCY 	     8-131
         8.5.1  Derivation 	     8-131
         8.5.2  Potency Index 	     8-137
    8.6  SUMMARY	     8-137
         8.6.1  Qualitative Assessment	     8-137
         8.6.2  Quantitative Assessment 	     8-143
    8.7  CONCLUSIONS 	     8-144
    8.8  REFERENCES 	     8-147
                                      vi

-------
                                LIST OF TABLES

Table                                                                 Page


3-1  Physical properties of trichloroethylene 	     3-1
3-2  Stabilizers in trichloroethylene formulations 	     3-2
3-3  Partition coefficients of trichloroethylene and other
      chlorinated hydrocarbons 	     3-4
3-4  U.S.  production capacity for trichloroethylene 	     3-4
3-5  Miscellaneous uses of trichloroethylene 	     3-5
3-6  Estimated half-life of TCI in surface waters 	     3-7
3-7  Ambient air mixing ratios of trichloroethylene 	     3-15
3-8  Mean 1 eve!s of TCI in surface waters 	     3-22

4-1  Recovery of radioactivity for 72 hours after a single
      oral dose of 14C-TCI (200 mg/kg) to female Wistar rats and
      NMRI mice 	     4-3
4-2  Disposition of single doses of 14C-TCI administered by
      gavage to Osborne-Mendel rats and B6C3F1 mice 	     4-4
4-3  Pulmonary uptake of TCI for 25 volunteers exposed to TCI for
      four consecutive 30-minute periods at rest and during
      exercise 	     4-5
4-4  Pulmonary uptake of TCI for 4 subjects at rest exposed to
      70 and 140 ppm TCI for 4-hour periods and 3-hour periods ...     4-8
4-5  Partition coefficients of TCI for various body tissues and
      components of rat and man 	     4-9
4-6  Recovery of radioactivity for 50-hour postinhalation
      exposure of male B6C3F1 mice to 10 or 600 ppm 14C-TCI for
      6 hours 	     4-11
4-7  Recovery of radioactivity for 50-hour postinhalation
      exposure of male Osborne-Mendel rats to 10 or 600 ppm
      14C-TCI for 6 hours 	     4-11
4-8  Organ contents of TCI after daily inhalation exposure of
      200 ppm for 6 hours per day 	     4-14
4-9  Proportion of TCI metabolites (TCA and TCE) in urine of
      rats and mice after single oral doses of 14C-TCI 	     4-21
4-10 Effect of chronic daily oral dosing of TCI (1000 mg/kg)
      on metabolism and metabolite excretion in B6C3F1 mice 	     4-24
4-11 Identification by GC/MS of TCI metabolites in  exhaled
      air and urine of Wistar rats and NMRI mice after a
      single oral dose of 200 mg/kg of 14C-TCI 	     4-28
4-12 Demonstrated metabol ites of TCI 	     4-33
4-13 Jji vitro covalent binding of TCI 	     4-50
4-14 Microsomal bioactivation and covalent binding of aliphatic
      halides to calf thymus DMA	     4-51
4-15 Hepatic and renal macromolecular binding of 1,1,2 (UL-14C)
      metabolite in male B6C3F1 mice and Osborne-Mendel rats
      exposed to 10 or 600 ppm for 6 hours 	     4-54
4-16 jLn vivo alkylation of hepatic DNA by 1,1,2 (UL-14C) TCI in
      male B6C3F1 mice dosed with 1200 mg/kg (14C) TCI by gavage .     4-54

5-1  Effects of TCI exposure on experimental animals 	     5-17
                                      Vll

-------
                          LIST OF TABLES (continued)


Table                                                                 Page

6-1  Summary of animal  studies of fetotoxic and teratogenic
      potential of TCI  	     6-3
6-2  Experimental design of Beliles et al.  study (1980) on
      Sprague-Dawley rats 	     6-6
6-3  Experimental design of Dorfmueller et al.  study (1979) on
      Long-Evans rats	     6-7
6-4  Experimental design of Beliles et al.  study (1980) on
      New Zealand white rabbits 	     6-8
7-1  Mutagenicity of technical-grade TCI in Salmonel1 a/mammalian
      mi crosome assay 	     7-4
7-2  Mutagenicity of trichloroethylene and epichlorohydrin in
      S. pombe by means of the host-mediated assay in B6C3F1 mice.     7-11
7-3  Mutagenicity of technical-grade TCI at the his 1-7 locus of
      Saccharomyces cerevisiae 	     7-13
7-4  Mutagemcity of reagent-grade TCI at the ilv locus in
      Saccharomyces cerevisiae 	     7-15
7-5  Mutagenicity of TCI at the ilv locus in Saccharomyces
      cerevisiae D7 	     7-17
7-6  Mutagenicity of TCI in the dominant lethal assay using NMRI
      mice 	     7-23
7-7  Numerical chromosome aberrations in TCI workers 	     7-25
7-8  Unscheduled DMA synthesis in WI-38 cells 	     7-30
7-9  Summary of tests for mutagenicity of TCI 	     7-38

8-1  TCI carcinogenicity bioassays in animals 	     8-2
8-2  Analysis of primary tumors in male rats 	     8-8
8-3  Dosage and observation schedule - TCI chronic study on
      Osborne-Mendel rats 	     8-13
8-4  Mean body weights, food consumption, and survival - TCI
      chronic study - male rats 	     8-14
8-5  Mean body weights, food consumption, and survival - TCI
        chronic study - female rats 	     8-15
8-6  Analysis of primary tumors in male mice 	     8-24
8-7  Analysis of primary tumors in female mice 	     8-25
8-8  Dosage and observation schedule - TCI chronic study on
        B6C3F1 mice  	     8-27
8-9  Mean body weights, food consumption, and survival - TCI
        chronic study - male mice  	     8-29
8-10 Mean body weights, food consumption and survival - TCI
        chronic study - female mice 	     8-30
8-11 Incidence of hepatocellular  carcinomas in B6C3F1 mice
        receiving TCI by oral gavage  	     8-34
8-12 Mouse skin  bioassays of TCI  and TCI oxide 	     8-38
8-13 Subcutaneous injection of TCI 	     8-39
8-14 Carcinogenicity of TCI by intragastric feeding in male
        and female HA: ICR Swiss mice  	     8-40
8-15 Estimation  of  probability of survival of  rats  	     8-43
8-16 Estimation  of  probability of survival of  hamsters  	     8-48
8-17 Estimation  of  probability of survival of  mice  	    8-49


                                     viii

-------
                          LIST OF TABLES (continued)


Table                                                                 page


8-18 Statistical analysis of the incidence of primary hepato-
      cellular neoplasms in control and TCI-treated B6C3F1 mice ..     8-54
8-19 Summary of negative carcinogenic!ty data for TCI 	     8-56
8-20 Chronic skin application in mice 	     8-58
8-21 Chronic subcutaneous injection in mice 	     8-59
8-22 Jji vitro transformation of Fischer rat embryo cell  cultures
      by TCI 	     8-63
8-23 Transformation and survival of Syrian hamster embryo cells
      treated with diverse chloroalkene oxides 	     8-70
8-24 A cohort study of mortality among men exposed to TCI in the
      1950s and 1960s and followed through December 1975 	     8-74
8-25 Comparison of workers exposed to TCI with the total Finnish
      population for mortality from all causes and for cancer
      mortality	     8-77
8-26 Incidence of hepatocellular carcinomas in male and female
      B6C3F1 mice in the NTP (1982) and NCI (1976) gavage
      studi es 	     8-83
8-27 Disposition of 14C-TCI 72 hours after single oral doses to
      male Osborne-Mendel and Wistar-derived rats and to male
      B6C3F1 and Swiss mice 	     8-88
8-28 Metabolism of TCI in B6C3F1 mice:  Effect of chronic dosing .     8-90
8-29 Disposition of 14C-TCI radioactivity for 71 hours after
      single oral dose (200 mg/kg) to rats and mice (NMRI) 	     8-92
8-30 Disposition of 14C-TCI 50 hours after inhalation exposure
      of male B6C3F1 mice 	     8-96
8-31 Disposition of 14C-TCI 50 hours after inhalation exposure
      of male Osborne-Mendel rats 	     8-96
8-32 Covalent binding to liver and kidney macromolecules of
      14C-TCI in male B6C3F1 mice and Osborne-Mendel rats 	     8-101
8-33 Dose conversion for the NTP bioassay (1982) in B6C3F1 mice ..     8-105
8-34 Dose conversion for the NCI bioassay (1976) in B6C3F1 mice ..     8-107
8-35 Slope estimates (qf) based on extrapolation from data in
      male and female    mice 	     8-116
8-36 Number of workers with cancer deaths 	     8-129
8-37 Relative carcinogenic potencies among 54 chemicals  evaluated
      by the Carcinogen Assessment Group as suspect human car-
      cinogens 	     8-133
                                      ix

-------
                                LIST OF FIGURES


Figure                                                                Page


4-1  Predicted partial pressure of TCI in alveolar air and
     tissue groups during and after an 8-hr exposure of
     100 ppm	     4-6
4-2  The relationship between the concentration of TCI in
      arterial blood and alveolar air at the end of inhalation
      exposures 	     4-8
4-3  Predicted partial pressure of TCI in fatty tissue for a
      repeated exposure to 100 ppm, 5 days, 6 hr per day 	     4-13
4-4  Elimination curves of TCI and kinetic model for transfer
      of TCI in the body 	     4-15
4-5  Whole blood levels (ug/ml) of TCI and metabolites after
      single gavage dose of TCI (1000 mg/kg) in corn
      oil to male Osborne-Mendel rats and B6C3F1 mice 	     4-20
4-6  Dose-TCI metabolite excretion relationship in male
      B6C3F1 mice and Osborne-Mendel rats given single
      intragastric doses 14C-TCI in corn oil 	     4-22
4-7  Relationship between TCI dose and the amount of total
      urinary metabolite excreted per day by mice in each
      group 	     4-25
4-8  Relationship between environmental TCI concentration
      and urinary excretion of TCI metabolites in human urine 	     4-31
4-9  Relationship between administered single oral doses of
      14C-TCI to rats and mice and amount of dose metabolized
      in 24 hr, expressed as mg/kg b.w., as calculated from
      14C-radioactivity excreted in urine, feces, and
      expired air 	     4-38
4-10 Postulated scheme for the metabolism of TCI to chloral
     and to trichloroethylene epoxide and its metabolites 	     4-40
4-11 Metaboli sm of chloral hydrate  	     4-41
4-12 Postulated scheme for the metabolism of TCI based on
      urinary metabolite profiles of rats and mice 	     4-42
4-13 Dose-effect relationship between chronic daily oral TCI
      dose and increase in liver weight of mice after six
      weeks.  Relationship between  liver weight increase with
      chronic dosing  at increasing  levels and total urinary
      metabolite (TCA and TCE) excreted per day by mice
      at various dose 1 evel s  	     4-60
4-14 Dose-effect relationship between chronic daily oral TCI
      dose and inhibition of  liver  glucose-6-phosphatase
      activity (G6P)  of mice  after  six weeks.  Relationship
      between inhibition of  liver G6P activity with
      chronic dosing  at increasing  levels and total urinary
      metabolite (TCA and TCE) excreted per day by mice
      at various dose levels  	    4-61
 8-1  Growth  curves  for  rats  administered TCI  in corn  oil by
       gavage 	
 8-2  Survival  curves  for  rats  administered TCI in corn  oil by
       gavage  	     8-7

-------
                          LIST OF FIGURES (continued)


Figure                                                                Page


8-3  Product-limit estimates of probability of survival -
      TCI-treated male rats 	     8-16
8-4  Product-limit estimates of probability of survival -
      TCI-treated female rats 	     8-17
8-5  Growth curves for mice administered TCI in corn oil by
      gavage ..;	     8-21
8-6  Survival curves for mice administered TCI in corn oil by
      gavage 	     8-22
8-7  Product-limited estimates of probability of survival -
      TCI-treated male mice 	     8-31
8-8  Product-limited estimates of probability of survival -
      TCI-treated female mice 	     8-32
8-9  TCI histogram of chamber concentration measurements T-l
      (100 ppm)  	     8-44
8-10 TCI histogram of chamber concentration measurements T-l
      (300 ppm)  	     8-45
8-11 TCI histogram of chamber concentration measurements T-3
      (600 ppm)  	     8-46
8-12 Incidence of malignant lymphomas in female mice 	     8-51
8-13 Cell transformation assays on vinyl chloride and TCI in
      BHK-21/C1  13 IQ vitro 	     8-66
8-14 Structures  and stability of chloroalkene oxides 	     8-71
8-15 Histogram representing the frequency distribution of the
      potency indices of 54 suspect carcinogens evaluated by the
      Carcinogen Assessment Group 	     8-132
                                      xi

-------
                             AUTHORS AND REVIEWERS
     The principal  authors of this document are:

Larry D. Anderson,  Carcinogen Assessment Group, U.S.  Environmental Protection
     Agency, Washington, D.C.

Steven Bayard, Carcinogen Assessment Group, U.S.  Environmental Protection
     Agency, Washington, D.C.

Vernon Benignus, Health Effects Research Laboratory,  U.S. Environmental
     Protection Agency, Research Triangle Park, N.C.

Chao W.  Chen, Carcinogen Assessment Group, U.S. Environmental Protection
     Agency, Washington, D.C.

I. W. F. Davidson,  Department of Physiology/Pharmacology, The Bowman Gray
     School of Medicine, Wake Forest University,  Winston-Salem, North Carolina.

John R.  Fowle III,  Reproductive Effects Assessment Group, U.S. Environmental
     Protection Agency, Washington, D.C.

Herman J. Gibb, Carcinogen Assessment Group, U.S. Environmental Protection
     Agency, Washington, D.C.

Mark M.  Greenberg,  Environmental Criteria and Assessment Office, U.S. Environ-
     mental Protection Agency, Research Triangle Park, North Carolina.

Jean C.  Parker, Carcinogen Assessment Group, U.S. Environmental Protection
     Agency, Washington, D.C.

     The following individuals reviewed earlier drafts of this document and
submitted valuable comments.

Dr. Mildred Christian
Argus Laboratories, Inc.
Perkasie, Pennsylvania  18944

Dr. Herbert Cornish
Dept. of Environmental and Industrial Health
University of Michigan
Ypsilanti, Michigan  48197

Dr.  I. W. F. Davidson
Dept. of Physiology/Pharmacology
The Bowman Gray School of Medicine
300 S.  Hawthorne Road
Winston-Salem, North Carolina  27103

Dr. John Egle
Dept. of Pharmacology
Virginia Commonwealth University
Richmond, Virginia  23298
                                      xii

-------
Dr. Thomas Haley
National Center for Toxicological Research
Jefferson, Arkansas  72079

Dr. Rudolph J. Jaeger
Institute of Environmental Medicine
New York University Medical Center
New York, New York  10016

Dr. John G.  Keller
P.O. Box 12763
Research Triangle Park, North Carolina  27709

Dr. Norman Trieff
Dept.  of Preventive Medicine
University of Texas Medical Branch
Galveston, Texas  77550

Dr. Benjamin Van Duuren
Institute of Environmental Medicine
New York University Medical Center
New York, New York  10016

Dr. James Withey
Food Directorate
Bureau of Food Chem.
Tunney's Pasture
Ottawa, Canada

Participating Members of the Carcinogen Assessment Group
Roy E. Albert, M.D.,  Chairman
Elizabeth L. Anderson, Ph.D.
Larry D. Anderson, Ph.D.
Steven Bayard, Ph.D.
David L. Bayliss, M.S.
Robert P. Beliles, Ph.D.
James Cogliano, Ph.D.
Chao W.  Chen, Ph.D.
Margaret M.L. Chu, Ph.D.
I.W.F. Davidson, Ph.D. (consultant)
Herman J. Gibb, B.S., M.P.H.
Bernard H. Haberman, D.V.M., M.S.
Charalingayya B. Hiremath,  Ph.D.
Robert E. McGaughy, Ph.D.
Jean C. Parker, Ph.D.
Charles H. Ris, M.S., P.E.
Dharm V.  Singh, D.V.M., Ph.D.
Todd W. Thorslund, Sc.D.
Participating Members of the Reproductive Effects Assessment Group

John R.  Fowle III, Ph.D.
K.S. Lavappa, Ph.D.
Sheila L.  Rosenthal,  Ph.D.
Carol N.  Sakai,  Ph.D.
Daniel S.  Straus, Ph.D.  (consultant)
Lawrence R.  Valcovic, Ph.D.
Vicki Vaughan-Dellarco,  Ph.D.
Peter E.  Voytek, Ph.D.
                                     xm

-------
The Environmental  Mutagen Information Center (EMIC) in Oak Ridge,  Tennessee,
kindly identified literature bearing on the mutagenicity of TCI.   Their initial
report and subsequent updates were used to obtain papers from which the
mutagenicity assessment was written.

Members of the Agency Work Group on Solvents
Elizabeth L.  Anderson
Charles H. Ris
Jean C. Parker
Mark M. Greenberg
Cynthia Sonich
Steve Lutkenhoff
Arnold Edelman
James A.  Stewart
Paul Price
William Lappenbush
Hugh Spitzer
David R.  Patrick
Lois Jacob
Josephine Brecher
Mike Ruggiero
Charles Delos
Jan Jablonski
Richard Johnson
Priscilla Holtzclaw
Office of Health and Environmental Assessment
Office of Health and Environmental Assessment
Office of Health and Environmental Assessment
Office of Health and Environmental Assessment
Office of Health and Environmental Assessment
Office of Health and Environmental Assessment
Office of Toxic Substances
Office of Toxic Substances
Office of Toxic Substances
Office of Drinking Water
Consumer Product Safety Commission
Office of Air Quality Planning and Standards
Office of General Enforcement
Office of Water Regulations and Standards
Office of Water Regulations and Standards
Office of Water Regulations and Standards
Office of Solid Waste
Office of Pesticide Programs
Office of Emergency and Remedial Response
                                      xiv

-------
                             1.  EXECUTIVE SUMMARY
     Trichloroethylene (1,1,2-trichloroethylene) (TCI) is a solvent  widely
used in the industrial degreasing of metals.   It has no known  natural  sources.
Current U.S. production is estimated at about 130,000 metric tons  per  year.
Of the TCI used in the United States, 80 to 95 percent evaporates  to the
atmosphere.  In addition to the workplace, TCI is found in a variety of  urban
and nonurban areas of the United States and other regions of the world.   It
has been measured in ambient air and water.  An average ambient  air  concentra-
tion of about 1 part per billion (ppb) would  be expected for some  large  urban
centers.  Concentrations as high as 32 ppb have been measured  in urban centers
in the United States, and 47 ppb has been measured in urban Tokyo.   Ambient
air concentrations of TCI are greatly influenced by the rate and geographic
distribution of emissions and the rate of decomposition and transposition  in
the atmosphere.  The average mixing ratio in  the troposphere of  the  northern
hemisphere is 11 to 17 parts per trillion (ppt).  Reaction with  hydroxyl  radi-
cals is the principal mechanism by which TCI  is scavenged from the atmosphere.
     TCI has been detected in both natural and municipal  waters  in the United
States.  Up to 403 ppb has been measured in some surface and subsurface  waters.
In finished drinking water, concentrations have ranged from 1  to as  much as 32
ppb.  There is no direct evidence of bioaccumulation of TCI in the food  chain.
Few studies have been made of the ecological  consequences of TCI in  the  envi-
ronment.
     The pharmacokinetics and metabolism of TCI have been studied  in man as
well as in animals.  Inhalation is the principal  route of concern  by which TCI
enters the body.  Ingestion of drinking water contaminated with  TCI  is another
important concern.  The extent of TCI absorption after oral  ingestion  is  virtual'
ly complete; with air exposure, the amount of TCI absorbed increases in  propor-
tion to its concentration in inspired air, the respiratory rate, and duration
of exposure.  TCI distributes widely into body tissues.   TCI is  eliminated by
two major processes, pulmonary excretion of unchanged TCI and  liver  metabolism
to urinary metabolites.  At air concentrations of 500 ppm (2690  mg/m3) or  less,
humans are estimated to metabolize between 60 and 90 percent of  absorbed TCI.
The proportion of the TCI dose metabolized in rodents has been demonstrated to

                                      1-1

-------
decrease at higher doses;  that  is,  metabolism  becomes  nonlinear and approaches
saturation.  However,  metabolism is linearly proportional to the inhaled concen-
tration dose in man,  at least up to about  300  ppm  (1614 mg/m^).  There  is  no
evidence that TCI metabolism in man is  saturation-dependent.   Studies have not
been made of TCI metabolism after oral  exposure  in  man.  At the levels  found or
expected in drinking  water, virtually  all  TCI  is expected to be absorbed and
metabolized.  There is persuasive experimental evidence that the metabolic
pathways for TCI are  qualitatively  similar in  mice,  rats, and  humans.   In  these
species, the principal urinary  metabolites of  TCI that have been identified
are trichloroethanol  (TCE) and  its  glucuronide,  and  trichloroacetic acid  (TCA).
Minor metabolites have also been identified for  each of these  species.   In the
liver, TCI is first metabolized to  chloral  hydrate  and to a reactive epoxide
(TCI oxide) by a microsomal P45Q system.   Reactive  intermediate metabolites such
as TCI epoxide covalently  bind  to cellular macromolecules, principally  protein,
and to a much lesser  extent, DNA.  Hepatotoxicity of TCI has been  shown to cor-
relate with the amount of  metabolism.   Metabolism of TCI is enhanced by drugs
(e.g., barbiturates,  oral  antidiabetic  agents) and  by  other xenobiotics (e.g.,
PCBs) which induce the hepatic  microsomal  P^Q metabolizing system.  Heightened
toxicity can result from drug interactions known to  occur with ethanol, barbi-
turates, disulfiram,  and Warfarin.
     Excluding carcinogenicity  as an end  point,  toxicity testing in experimen-
tal animals, coupled  with  limited human data derived principally from excessive
exposure, suggests that long-term exposure of  humans to environmental  (ambient)
levels of TCI is not  likely to  represent  a health concern.  Behavioral  and psy-
chological effects, particularly as they  affect  psychomotor performance,  have
been reported at levels of 200 ppm (1076  mg/m^)  in  some, but not all, experi-
mental and epidemiologic studies.  There  is no information that such impair-
ment, representative of dysfunction of the central  nervous system  (CNS),  would
occur during chronic, low-level exposures.  At levels  found or expected in the
ambient environment,  such an effect would be unlikely. Similarly, dysfunction
of the liver or kidneys would not be likely during  or as a result  of environ-
mental exposures.
     Teratogenicity is another health end point  for which  available data  from
experimental animals suggest that the conceptus  is  not uniquely susceptible to
TCI.  Exposure of rats, mice, and rabbits, during  gestation, to  levels  (300 ppm

                                      1-2

-------
or 1614 mg/m^) greatly in excess of those generally found in the environment,
has not been observed to result in any teratogenic effects.   The teratogenic
potential of TCI for humans cannot be directly extrapolated  from the observa-
tions in the animal studies.  However, the animal  studies suggest that at low
ambient levels that do not cause maternal toxicity, TCI would not pose a signi-
ficant hazard to the developing conceptus.
     With respect to the mutagenic potential  of TCI, available data provide
suggestive evidence that commercial-grade TCI is a weakly-active, indirect
mutagen, causing effects in a number of different  test systems representing a
wide evolutionary range of organisms.  Thus,  commercial TCI  may have the poten-
tial to cause weak or borderline increases above the spontaneous level  of muta-
genic effects in exposed human tissue.  Adverse effects in the testes of mice
exposed to commercial TCI suggest that TCI could cause similar effects  in
humans.  A conclusion about the mutagenic potential of pure  TCI cannot  be made.
If TCI is mutagenic, the available data suggest that it would be a very weak,
indirect mutagen.
     The evidence reviewed in this document for the carcinogenicity of  TCI
in experimental  animals includes:  increases  in the incidence of hepatocellu-
lar carcinomas in male and female B6C3F1 mice (three studies); malignant lym-
phomas in female HantNMRI mice; and renal adenocarcinomas in male Fischer 344
rats.  Statistically significant increases of malignant liver tumors were
observed in both male and female B6C3F1 mice.  A gavage study of purified TCI
in both sexes of B6C3F1 mice was conducted with basically the same experimental
design as a gavage study of technical grade TCI in male and  female B6C3F1 mice
to evaluate the role of epoxide stabilizers in TCI in the induction of  hepato-
cellular carcinomas.  Similar carcinogenic responses were observed for  the
purified epoxide-free and for stabilized TCI.  The other studies provide some
additional support to the overall body of evidence, particularly since  two of
them were carried out in a different species  or strain.  However,  the third
study in B6C3F1 mice, an inhalation study in  which an increase in  liver tumors
was seen, was weakened by deficiencies in its conduct.   In the inhalation study
in Han:NMRI mice, a 30 percent incidence of spontaneous lymphoma in control
mice, and the possibility of an indirect effect in treated mice,  made these
results difficult to interpret.  A borderline response  was observed in  the
gavage study of Fischer 344 rats.

                                      1-3

-------
     A gavage study with male and female Osborne-Mendel  rats was  negative;
however, this study may be considered  inconclusive  because  of  high mortality
in the treatment groups.  Exposure to  airborne  concentrations  of  TCI  did  not
result in a carcinogenic effect  in Han:Wist  rats  and  Syrian hamsters, although
higher dose levels of TCI probably could have been  given  to the test  animals.
     Other long-term animal  studies did not  suggest a carcinogenic potential
for TCI.  Most of these studies,  however,  were  not  specifically designed  to
evaluate carcinogenicity, involved only small treatment  groups, lacked  suffi-
cient doses of TCI, or were of insufficient  duration  in  view of the expected
latency period for cancer induction.
     TCI induced malignant tumors of the liver  in both male and female  B6C3F1
mice in multiple studies.  This  constitutes  a signal  that TCI  might be  carcino-
genic in man.  While kidney tumors observed  in  the  NTP rat  study  are  not  strong
indications of a response in a second  species because of  the small number of
animals responding (3 of 49 animals) and because  of the  high mortality, the
statistical significance after mortality corrections  suggests  that a  carcino-
genic effect may be taking place.
     The U.S. Environmental  Protection Agency's Proposed  Guidelines for Carci-
nogen Assessment (U.S. EPA, 1984), take the  position  that the  mouse-liver-
tumor-only response, when other conditions for  a  classification of "sufficient"
evidence in animal studies are met, should be considered  as "sufficient"  evi-
dence of carcinogenicity.  Thus,  based on  EPA's proposed  cancer guidelines,
the overall evidence for TCI would result  in a  classification  of  B2,  i.e., a
probable human carcinogen.
     The TCI carcinogenicity results could be classified  under the criteria of
the International Agency for Research  on Cancer (IARC) as either  "sufficient"
or "limited" depending on which of the differing  current  scientific views
about chlorinated organic compound induction of liver tumors in mice  is chosen.
Since there are no adequate epidemiologic data  in humans, the  overall ranking
of TCI under the criteria of the IARC  depends primarily  upon the  position taken
regarding the mouse liver tumor.   Thus, the overall ranking could be  either
Group 2B or Group 3.  The more conservative public  health view would  regard
TCI as a probable human carcinogen (Group 2B),  but  there  is also  scientific
sentiment for regarding TCI as an agent that cannot be classified as  to its
carcinogenicity for humans (Group 3).

                                      1-4

-------
     Additional carcinogenicity studies on purified TCI  (no detectable  epox-
ides) are currently in progress.  These studies and others  published recently
will be incorporated into any updates of this Health Assessment  Document.
     Four sets of gavage bioassay data on hepatocellular carcinomas  in  male and
female mice provide a basis to calculate an upper-bound  carcinogenic slope
estimate for TCI using a linearized multistage model.  The  development  of these
risk estimates is for the purpose of evaluating the "what-if"  question:   If TCI
is carcinogenic in humans, what is the possible magnitude of the public health
impact?  Since upper-bound potency estimates calculated  on  the basis of these
data sets are comparable, ranging from 5.8 x 10~3 to 1.9 x  10"2  mg/kg/day,
the geometric mean, 1.1 x 10~2 mg/kg/day is used to calculate  the incremental
lifetime cancer risk (i.e., unit risk) due to .a unit.exposure  of TCI in drinking
water and in air.  The upper-bound estimate of the cancer risk due to 1 yg/L
of TCI in drinking water is 3.2 x 10~?.  The upper-bound estimate of the cancer
risk due to 1 pg/m3 of TCI in air is 1.3 x 10~6.  The  upper-bound nature of
these estimates is such that the true risk is not likely to exceed this value
and may be lower.  The carcinogenic potential of TCI is  generally considered to
reside in cellular reactive intermediate metabolites,  and therefore  metabolic
and pharmacokinetic factors have been used in the calculation  of the drinking
water and air unit risks.
     None of the epidemiologic studies reviewed in this  report provide  positive
evidence from which to estimate a unit risk for exposure to TCI.  As an alter-
native, an upper-bound estimate has been calculated from one negative study.
The calculation on the basis of human data results in  a  greater  risk estimate
than the estimate from animal data, and thus does not  appear to  contradict the
risk estimate calculated from the animal  data.
     Expressed in terms of relative potency, TCI ranks in the  lowest quartile
among the 54 suspect or known human carcinogens evaluated by EPA's Carcinogen
Assessment Group.
                                      1-5

-------
                      RECOMMENDATIONS FOR FURTHER STUDIES

     Areas for which incomplete information  is  available, and  which  should  be
considered in formulating research  needs, are presented below.  The  needs
listed, however, are not necessarily  in  the  order of  relative  priority.
     1.  Teratogenicity and Reproductive Effects.  There are few animal
studies via inhalation that adequately assess the teratogenic  and reproductive
effects potential  of TCI.  Uncertainty about these biological  end points  in
humans should serve as a stimulus for future research.
     2.  Neurobehavioral Toxicity.   There have  been few animal studies  of the
effects of TCI on  the nervous system and behavior.  Most end points  studied
have been relatively insensitive.  Further studies are  warranted on  more  sen-
sitive end points.
                                      1-6

-------
                               2.   INTRODUCTION
     Trichloroethylene (TCI) is a member of a family of unsaturated chlorinated
aliphatic compounds.  Trichloroethylene,  though  a water and solid waste con-
taminant, is primarily of interest in situations  involving ambient air exposure.
It is released into ambient air as a result of evaporative  losses  during pro-
duction, storage, and/or use.  It has no  known natural sources.  It is photo-
chemical ly reactive in the troposphere and is removed by scavenging mechanisms,
principally via hydroxyl  radicals.
     The scientific  data base concerning the effects of TCI  on  humans  is
limited. Effects  on humans have generally been ascertained from studies invol-
ving individuals  occupationally or accidentally exposed.   During such exposures,
the concentrations associated with adverse effects on human health were either
unknown or  far in excess of concentrations found or expected in ambient air.
Controlled  TCI exposure  studies have been  directed toward elucidating the
effects  on  the CMS,  effects  on  clinical  chemistries,  and pharmacokinetic
parameters of TCI exposure.
     The available  epidemiologic  studies have not  been  able to adequately
assess  the  overall  impact of  TCI  on human health.    It  has therefore been
necessary to rely greatly on animal  studies to derive indications of potential
harmful  effects.   Although animal  data cannot always be extrapolated to humans,
indications of probable or likely effects in animal  species increase confidence
that similar effects may occur in humans.
     This document is intended to provide an evaluation of the scientific data
base concerning  TCI.   The publications cited  in  this document represent a
majority of the  known scientific references to TCI.   Reports which had little
or no bearing  on  the  issues discussed were not cited.  Some topic  areas, such
as effects on terrestrial ecosystems, are only briefly discussed.   Such topics
reflect the nature of the scientific data base.  On-going  literature  searches
have been conducted through 1985,  resulting in the inclusion of selected refer-
ences in this document.   However,  the basic literature search for Chapter 8 is
current through  1982; the  basic  literature search  for Chapter  7  is  current
through 1983.   The Agency is aware of additional  carcinogenicity and mutagenicity
information that  has  subsequently become  available.   This information will be
considered for incorporation in any future updates of this document.
                                    2-1

-------
                           3.   BACKGROUND INFORMATION
3.1  PHYSICAL AND CHEMICAL PROPERTIES
     Trichloroethylene (TCI) is a  colorless,  highly volatile liquid that is
miscible  with a variety of solvents.  Some of its  important physical proper-
ties are shown in Table 3-1.

             TABLE 3-1.  PHYSICAL PROPERTIES OF TRICHLOROETHYLENE

     Molecular weight                             131.39
     Boiling point                                87°C
     Vapor pressure                               94 mm Hg @ 30°C
     Vapor specific gravity                       4.55 at boiling point
                                                  (air = 1)
     Solubility @ 20°C in water                   0.107 g percent
     At a temperature  of  25°C and a pressure of 1 atmosphere, one part-per-
                                                     3
million (ppm) of the vapor is equivalent to 5.38 mg/m  in air.
     Trichloroethylene  is  photochemically  reactive  and  autooxidizes  upon
catalysis by free radicals (McNeil!,  1979).   Autooxidation is greatly acceler-
ated by high temperatures and exposure to ultraviolet radiation (McNeill, 1979).
Some of its degradation products, e.g., hydrochloric  acid (HC1),  are corrosive
to metals.   These decomposition products include dichloroacetyl chloride, phos-
gene, carbon monoxide, hexachlorobutene, and HC1 (McNeill, 1979.)
     Decomposition is catalyzed when TCI comes in contact with aluminum metal.
The HC1 produced reacts with aluminum to yield aluminum chloride (A1C13) which
catalyzes formation of hexachlorobutene (McNeill, 1979).   Sufficient quantities
of aluminum have been reported to cause violent decomposition of TCI (Metz and
Roedy, 1949).
     While TCI  is  nonflammable under ordinary conditions, mixtures of TCI  and
oxygen will ignite at 25°C when the TCI concentration is  between 8 to 10 percent
in air (McNeill, 1979).
     To inhibit  its  decomposition, stabilizers  are added  to  commercial  formu-
lations of TCI (McNeill, 1979).   Commonly used stabilizers in vapor degreasing
grades of TCI are neutral  inhibitors and free radical scavengers (McNeill, 1979).
                                    3-1

-------
Epoxides (including epichlorohydrin,  a  known  animal  carcinogen) are added to

scavenge free HC1  or A1C1,.   Antioxidants are  usually added in quantities less
                         O
than 1 percent by weight (Hardie, 1964).   Fuller (1976) reported that inhibited

grades of TCI are stable  up to 130°C  in the presence of air, moisture, light,

and common construction metals.   At higher temperatures,  inhibitors are less

effective and corrosion of metals can result.   Stabilizers that have been used

in TCI formulations are shown in Table 3-2.
TABLE 3-2.  STABILIZERS IN TRICHLOROETHYLENE FORMULATIONS (Beckers, 1972; U.S.
      EPA, 1975; Hardie, cited in:  Kirk Othmer, 1975; Schlossburg, 1980)
Amyl alcohol
Propanol
Diethyl amine
Triethyl amine
Dipropylamine
Diisopropylamine
Diethanolamine
Morpholine
N-Methy1morphol i ne
Aniline
Acetone
Ethyl acetate
Borate esters
Ethylene oxide
1,2 Propylene oxide
1,2 Epoxy butene
Cyclohexene oxide
Propylene oxide
Butadiene dioxide
Styrene oxide
Pentene oxide
2,3 Epoxy 1-propenol
3, Methoxy-1,2 epoxy propane
3, Ethoxy-1,2 epoxy
Stearates
2-Methyl-l,2 epoxy propanol
Epoxy cyclopentanol
Epichlorohydrin
Tetrahydrofuran
Tetrahydropyran
1,4-Dioxane
Dioxalane
Trioxane
Alkoxyaldehyde hydrazones
Methyl ethyl ketone
Nitromethanes
Nitropropanes
Phenol
o-Cresol
Thymol
p-tert-Butylphenol
p-tert-Amylphenol
Isoeuganol
Pyrrole
n-Methyl pyrrole
n-Ethyl pyrrole
(2-pyrryl) Trimethylsilane
Glycidyl acetate
Isocyanates
Thiazoles
      Trichloroethylene  decomposes at  temperatures  up  to  700°C.   Substances  found

 in  the decomposition mixture  include dichloroethylene,  tetrachloroethylene,
 carbon  tetrachloride, chloroform, and methyl  chloride (Hardie,  1964).   Formation

 of  phosgene,  a highly  toxic  gas, was observed when TCI came in contact with

 iron, copper, zinc, or aluminum  over the temperature range  250°C to 600°C

 (Sjoberg,  1952).   When TCI came in contact with zinc at 450°C, 69 mg phosgene

 per gram  of TCI was  produced.
                                     3-2

-------
     Trichloroethylene is not readily hydrolyzed by water (McNeil!, 1979).   In
the presence of  strong  alkali  (e.g., sodium hydroxide) TCI can decompose to
dichloroacetylene, an explosive,  flammable,  and highly toxic compound (Humphrey
and McClelland, 1944).   Noweir et al. (1973) recommended that TCI  be stored in
cans or dark glass bottles to minimize decomposition.
     Because of  its  lipophilic properties,  TCI can be expected to partition
selectively in the body's adipose tissue.   The relationships between such parti-
tioning and manifestation of its toxic properties are discussed in Chapter 4.
The unsymmetrical nature  of  the  molecule and the electron-withdrawing effect
of the chlorine  substituents (leading to destabilization of the charge at the
double bond) have been proposed as factors  responsible for the supposed muta-
genicity of TCI (Bonse and Henschler, 1976).
     Measurements of the lipophilic potential of TCI have been made by a number
of investigators  (Waters  et al., 1977; Mapleson, 1963; Powell, 1947; Sato and
Nakajima,  1979).  Mapleson (1963) reported  an  oil/water partition coefficient
of 960:1 at 37°C.  Waters et al. (1977) reported the oil/water coefficient as
900:1.  Sato and Nakajima (1979)  reported an olive oil/water partition coeffi-
cient of 552:1  at 37°C  and an olive oil/blood partition coefficient of 76:1.
The relationship of TCI  to other compounds with respect to partition coeffici-
ents  is shown  in Table  3-3 (Sato and Nakajima, 1979).   These values indicate
that TCI is highly lipophilic.
3.2  ENVIRONMENTAL FATE AND TRANSPORT
3.2.1  Production
     Trichloroethylene has been declining in production in recent years because
of restrictions on emissions (Federal Register, 1977) and substitution by other
solvents.  According  to  statistics  published by the U.S. International Trade
Commission (1982), 117,363 metric tons of TCI were produced in 1981.
     Recent estimates of world production of TCI are not available.  Such figures
for 1973 indicated 700,000 metric tons were produced (McConnell et al., 1975).
About two-thirds was produced outside the United States.
     In 1977, TCI was produced by four American manufacturers (SRI, 1982).  The
producers and production capacities are shown in Table 3-4.
     Most of  the TCI  produced commercially  is  derived from  ethylene or
dichloroethane  (EDC)  (McNeill,  1979).   Dichloroethane,  produced  by
chlorination of ethylene, can be further chlorinated to TCI and

                                    3-3

-------
       TABLE 3-3.   PARTITION  COEFFICIENTS  OF  TRICHLOROETHYLENE  AND  OTHER
              CHLORINATED  HYDROCARBONS  (Sato  and  Nakajima,  1979)


Solvent
Dichloromethane
Chloroform
Carbon tetrachloride
1,2-Dichloroethane (EDC)
Methyl chloroform
Tetrachl oroethyl ene
Trichloroethylene


Oil /water
21
115
1444
40
383
4458
552


Oil /blood
16
39
150
23
108
146
76
Log P
(partition
coefficient
1.32
2.06
3.16
1.60
2.58
3.65
2.74
    TABLE 3-4.   U.S.  PRODUCTION CAPACITY FOR TRICHLOROETHYLENE (SRI,  1982)
     Producer
Site
Annual capacity, Jan.  1981
      (metric tons)
PPG Industries
Dow Chemical
Diamond Shamrock
Ethyl Corporation
Lake Charles, LA
Freeport, TX
Deer Park, TX
Baton Rouge, LA
91,000
54,000
23,000
20,000
tetrachloroethylene at 280° to  450°C.   Catalysts include potassium chloride,
aluminum chloride, fuller's earth,  graphite,  activated carbon,  and activated
charcoal (McNeill, 1979).   Maximum  conversion to TCI  (75%  of  EDC feed) is
achieved at a chlorine-to-EDC  ratio  of 1.7:1 (McNeill,  1979).
     Oxychlorination of ethylene or EDC to TCI is achieved at 425°C and at 20
to 30 psi  (McNeill,  1979).  Common  catalysts are mixtures  of  potassium and
cupric chlorides.   Conversion  efficiency is between  85  and 90 percent.

3.2.2  Use
     It has  been  estimated  that from 80 to 95 percent  of the TCI produced in
the United States  is  used in  the degreasing  of  fabricated metal  parts  (SRI,
1978; McNeill, 1979; Chemical  and Engineering News,  1975).  The remaining 5 to
20 percent is divided equally between exports and miscellaneous  applications.
Use in  vapor degreasing is  estimated to account  for 87 percent of production,
according to  industry trade statistics  (Chemical Marketing Reporter, 1978).
                                    3-4

-------
     In addition to its use as a vapor degreasing agent, TCI has a wide variety
of other uses  (Table  3-5).   About 4,500 to  6,800  metric tons are consumed
annually in the  production  of polyvinyl  chloride, in which TCI is used as a
chain-transfer agent (McNeil 1, 1979).

               TABLE 3-5.   MISCELLANEOUS  USES OF TRICHLOROETHYLENE
                      (McNeil!, 1979;  Waters et al.,  1977)

         Adhesive formulations
         Paint-stripping formulations
         Low-temperature heat-transfer medium
         Carrier solvent in industrial paint systems
         Solvent in textile dyeing and finishing
     Uses of TCI  that  have been discontinued in the United States because of
environmental and health restrictions include use as an inhalation anesthetic
(Page and Arthur, 1978;  Waters  et al.,  1977), use in fumigant mixtures (Farm
Chemicals Handbook,  1976),  and  use as an extractant in the decaffeination of
coffee.   Use in anesthetic procedures may be continuing to some extent in other
countries (Cundy, 1976).
     Declining consumption of TCI from the early 1970's has been offset by the
increased use  of substitute solvent, e.g., methyl  chloroform  and methylene
chloride (U.S.  EPA,  1979).

3.2.3  Emissions
     There are no known natural  sources of TCI.   Graedel and Allara (1976) con-
cluded that atmospheric chemical production of TCI was negligible.  Because of
its dispersive  use  pattern in degreasing operations, most of the TCI used is
expected to be emitted to the atmosphere.  Emissions estimates (U.S.  EPA, 1979)
for 1976, pertaining to vapor degreasing, cold cleaning, and miscellaneous use
categories suggest that 130,000 metric tons were discharged to the atmosphere.
     Solvent emissions from  vapor  degreasing occur  primarily during  unloading
and loading of the degreaser (Bucon et al., 1978).  Daily emissions of a single
spray degreasing  booth may vary from a  few pounds  to  1,300 pounds (Bucon et
al., 1978).   In uncontrolled degreasing operations,  as much as 2.8 kg/hr can be
emitted, based on 200,000 pounds of metal cleaned per day (Bucon et al., 1978).
     Singh et al. (1979)  estimated global emissions  of TCI for 1977 at 435,000
metric tons  ±  50  percent.   Estimates were  derived from production  figures and
usage patterns.
                                    3-5

-------
3.2.4  Persistence
     Many processes occur  in  the  troposphere which can alter the atmospheric
levels of TCI.  Once  emitted  into the troposphere, vertical  and horizontal
mixing occurs. Transport is highly dependent upon the  length of time TCI remains
in the troposphere, which  is  determined largely by the  extent to which TCI
reacts with hydroxyl  free  radicals (-OH),  the principal  scavenging mechanism
for TCI and many other halogenated compounds in the atmosphere.
     Edney et al.  (1983), on the basis of the observed rate of reaction of TCI
with  -OH  in a reaction cell,  calculated an  atmospheric  lifetime for TCI of
about 54  hours.   An -OH  concentration  in the  troposphere of  10  molecules/cm
was assumed.
     The average northern hemisphere background mixing ratio is between 11 and
17 ppt  (Singh et  al., 1979),  which  suggests  that  TCI is short-lived in the
troposphere.   Singh et al.  (1979) estimated a residence time of about 2 weeks,
an estimate  based on  a seasonally-averaged  -OH concentration of 4 x  10
molecules/cm  and  a  rate constant (National  Bureau of Standards,  1978), at
265°K, of 2.30  x  10"12.  Singh and coworkers (1979) estimated that, given an
                            6             3
•OH  concentration  of  1 x 10  molecules/cm , 20 percent of the TCI in ambient
air  can  be destroyed  each day.   Crutzen et  al.  (1978) estimated a residence
time  of  11 days for TCI.  Derwent and Eggleton  (1978) estimated the lifetime
at 0.04 year  (^ 15 days).  The percentage of the ground level  injection of TCI
that was  estimated to  survive free radical attack was  0.4 percent.
      Seasonal  variations in  -OH  concentrations, important for  longer-lived
species,  are  not  expected  to  play a  significant role  in  the tropospheric sur-
vival  of TCI.  Altshuller (1980)  calculated  that in January  (when  -OH  levels
and  solar flux are low)  0.6 day would  be required for  the photochemical  decom-
position  of  1 percent of the  ambient TCI.   In July (when -OH levels and solar
flux are  high), only  0.07  day  is  required.
      Trichloroethylene was observed  to have  a slow decomposition rate  in dilute
aqueous  solution  (Billing  et  al.,  1975).  Its half-life  (in  the  dark)  was  10.7
months  at ambient temperatures.   When  the solution was exposed to  sunlight,  75
percent  of the TCI decomposed  in 12 months,  compared  with  54 percent  for  the
reaction in the dark.   A correction  for  the  amount of  TCI that volatilized into
the  air space above the solution  was not employed.  On the  other hand,  Pearson
and  McConnell (1975),  in their determinations of the  hydrolytic  decomposition,
extrapolated  to zero  volatilization.  Their  estimate  was  a  half-life  of  30
months.

                                     3-6

-------
     The major route of removal  of TCI from water is volatilization.   Volatil-
ization of TCI  from water has been investigated by several  authors.   Oil ling
et al.  (1975) have  shown  that the  loss of TCI from an agitated dilute aqueous
solution occurs  exponentially with an evaporative half-life  of 21 ± 3 minutes.
Scherb (1978) measured the volatilization of TCI from an aeration channel in a
wastewater treatment plant and found the half-life to be 3 hours.  The rate of
volatilization of TCI from surface waters in the environment has been reported
by Zoeteman et al.  (1980). TCI was found to have a half-life of 1 to 4 days in
the Rhine River  and 30 to 40 days in a tidal estuary.  Smith et al.  (1980) have
shown that the  rate of volatilization of TCI and other low-molecular-weight
compounds from  various  bodies  of waters is a function  of reaeration rates.
From the work of  Smith et al.  (1980), it is estimated that  the half-life for
TCI in surface waters ranges from 3 hours,  for rapidly moving shallow streams,
to 10 days or longer for ponds and lakes (Table  3-6).

           TABLE 3-6.  ESTIMATED HALF-LIFE OF TCI IN SURFACE WATERS

     Water Type                                        TCI  half-life (days)
       Pond                                                    11
       Lake                                                  4 to 12
       River                                                 1 to 12

Source:  Calculated from information in Smith et  al.  (1980).

3.2.5  Degradation
     The transformation products of the decay of TCI have  been investigated in
chamber studies  simulating atmospheric conditions.   Two principal transformation
products have been  observed,  phosgene and dichloroacetyl  chloride.   Dahlberg
(1969) observed  that both were produced when TCI was photooxidized in air.  Gay
et al. (1976) also found formyl  chloride, in addition to phosgene and dichloro-
acetyl chloride, when TCI was photooxidized in the presence  of nitrogen dioxide
(N02).  A maximum observed consumption rate for  TCI of 70  percent per hour was
reported.  In this system, dichloroacetyl chloride, identified by long-path in-
frared  spectroscopy,  was the principal  product.   In a smog chamber system
containing 2.5 ppm  TCI  and  0.5  ppm NO,,, TCI disappeared at  the rate of about
25 percent per hour (Dimitriades et al., 1983).   Ozone built up to 0.5 ppm.  The
presence of equal  amounts of tetrachloroethylene and TCI in  a chamber containing
N0  did not alter the consumption rate of TCI above TCI controls.
                                    3-7

-------
     Edney et al.  (1983)  observed  a number of degradation products when «OH
radicals reacted with TCI in a 700 liter cell coupled to an FTIR spectrometer
and scanning interferometer.   Reaction products  observed were  phosgene,  dichlo-
roacetyl chloride,  formyl  chloride, carbon monoxide, and nitric acid.   The  rate
                                 -12   3
constant of reaction was 3.6  x 10    cm /sec.
     Lillian et al. (1976),  finding phosgene formation from the photooxidation
of the  TCI  congener,  tetrachloroethylene,  considered that phosgene formation
also was likely from  other  chloroethylenes.  Singh (1978) reported that -OH-
catalyzed photodecomposition of TCI, as occurs in the troposphere, can result
in phosgene as the  principal  transformation product.
     Under simulated atmospheric conditions in  the presence of nitric oxide
(NO) and N02, TCI (10 ppm) was very short-lived  (less than 3.5 hours) (Dilling
et al., 1976).  A  consumption rate up to  30  percent  per hour was reported.
The high reactivity  of  TCI  and its catalytic role in the formation of photo-
chemical smog  have been  discussed (U.S.   EPA,  1978a;  National Academy of
Sciences, 1977).   Free  radicals  (e.g., -RO,,), are  formed when TCI is photo-
decomposed,  and  these in turn react  with  NO by oxidizing it  to  NO^.   Ac-
cording to  the  following  reaction  sequence, N02 results  in the formation of
ozone (03):

                    (1)  N02  ——	» NO + 0
                    (2)  0 + 02 + M (third body) 	> 03 + M
     Because chamber  studies  simulate some atmospheric reactions believed to
occur in the environment, further studies of greater complexity would be needed
to determine the extent and pathways by which TCI is photodecomposed.  Phosgene
and dichloroacetyl  chloride,  though likely to be  formed  in  the  troposphere,
also are subject to transformation and scavenging processes.
3.3  LEVELS OF EXPOSURE
3.3.1  Analytical Methodology
     Of the analytical methods and procedures available for detecting and quan-
titating  halogenated  hydrocarbons  present in ambient air,  water,  soil,  foods,
etc.,  two systems are commonly  used:   gas chromatography/mass spectrometry
                                    3-8

-------
(GC-MS) and gas chromatography with electron capture detection (GC-ECD).   Each
system has the capability of detecting TCI concentrations as low as a few parts
per trillion by volume.   The usefulness of GC-ECD is that the apparatus can be
used  in  the field to provide quasi-continuous measurements  by intermittent
sampling (every 15 to 20 minutes).   Such use can circumvent decomposition that
results when samples of TCI are stored for long periods before assay (Pellizzari,
1974; Lovelock, 1974b).
3.3.1.1  Gas Chromatography with Electron Capture Detection.   The GC-EC metho-
dology has been reviewed by Pellizzari (1974) and by Lovelock (1974b).   The EC
detector is  specific  in that halogenated compounds are quantified while non-
halogenated hydrocarbons are  not  detected.   Thus, high background  levels of
non-halogenated hydrocarbons  in ambient air  or water samples do not interfere
with measurements of  halogenated  compounds.   In a complex mixture  in  which
several compounds may have similar retention times, alteration of the operating
parameters of the GC-EC system will usually provide separation of the components.
     In the method  developed by Rasmussen et al. (1977), the detection limit
for TCI was  0.3  ppt.   With four samples of ambient air, the percent standard
deviation was  23.0  by electronic area  integration  (+ 5  percent by peak height
measurement).  The  system,  which  employs  a freezeout concentration loop, has
the capability of measuring various  halocarbons in 500-ml aliquots of air.
Total analysis time is  40 minutes.  The  GC  column consisted of a 6-mm x 3-m
stainless  steel  column  packed with  10  percent SF-96 on 100/120  mesh-
Chromosorb W.
     A similar analytical procedure was employed by Harsch and Cronn (1978) to
measure levels of halocarbons in  30-ml samples of low-pressure stratospheric
air.  For  TCI, the  detection  limit was 1  ppt and the precision (percent  stan-
dard deviation)  for 10  samples with a mixing ratio of 22 ppt was 10 percent.
     Gas chromatography with electron capture detection also has been the method
of choice by Singh and coworkers for measuring the ambient air levels of a wide
variety of  halocarbons.   Singh  et al. (1979) maintained  the  EC  detector at
325°C.  The  GC column used was 6-ft x 1/8-in.  stainless steel, containing 20
percent SP2100,  0.1 percent Carbowax  1500 on 100/120 mesh Supelcoport.   Singh
et al.  (1979b) recommended two columns for  separation  of TCI:  (a) silicone
oil 10 percent DC 200 on 30/60 mesh Chromosorb W in a 6-ft x 1/4-in. stainless
steel column or (b)  20 percent SP-2100 on 80/100 mesh Supelcoport in a 15-foot
x 1/8-inch column.
                                    3-9

-------
     The applicability of GC-ECD to analysis of TCI and other halohydrocarbons
purged from water samples has been discussed by Bellar et al. (1979).   Precon-
centration on a Tenax and silica gel  trap was recommended.
     Singh et al. (1979a) avoided the use of Tenax traps for preconcentrating
organics in air  samples  because of artifacts.   They  thought that oxygen or
ozone was oxidizing  the  Tenax monomer and producing  electron-absorbing oxy-
genated species, thus interfering with EC detection of the organics of interest.
     Makide et al.  (1980) used GC, coupled with constant current ECD,  to measure
TCI in ambient air at sites in Japan.   An extremely low detection limit of <0.2
ppt (<1.36 x 10   mg/m ) was reported.   Sampling canisters were made of stain-
less steel and were  polished  electrochemically.  Preconcentration was carried
out on  a silicone OV-101 column  at -40°C.  Column  temperature was raised  at a
rate of 5°C per minute up to 70°C for separation of components.
     A  direct headspace technique  for use with  EC  detection  has  been  reported
by Piet and  coworkers (1978).  This technique avoids the need for preconcen-
tration traps when water samples are purged.   An overall detection limit below
1 ug/1  was  reported.   With isothermal operation of the GC, a detection limit
for TCI of 0.05 ug/1 was reported.  An OV225 capillary column was used.
     Dietz and  Singley  (1979) described a headspace  method  applicable  to EC
detection of volatile organics  from water samples.  A standard  deviation  of 5
percent was reported for routine analyses of drinking, natural, and industrial
waters.  Concentrations ranging from 0.1 ppb to several ppm of TCI were detec-
ted.   In this  method, the water sample  is  sealed  in a vial  until  organics
equilibrate between  the  water and vial headspace.   A major advantage is that
only relatively  volatile,  water-insoluble  compounds tend to partition  in the
headspace.  In  a comparison of  this method with the purge  and trap  technique,
Dietz  and  Singley  (1979)  obtained comparable results.   However,  the chromato-
grams  from the purge-trap method were of poorer quality.  The headspace method
was recommended  for  routine sample analyses of  volatile halocarbons.
     Wang and  Lenahan (1984)  described  a closed-loop  system  that combines the
techniques of both gas stripping and headspace, with  a reported  detection  limit
of about 0.2 H9/1-
3.3.1.2 Gas Chromatography-Mass Spectrometry.  While GC-ECD  relies on  retention
time and ionization  efficiency for compound  identification,  the  equally sensi-
tive GC-MS  system  identifies compounds  by their characteristic  mass  spectra.
However, the  expense, bulk,  and demands on  the operator preclude routine  use
                                    3-10

-------
of GC-MS in field measurements.   Often, GC-MS is used to verify results obtained
by GC-ECD.
     Comparisons between the precision afforded by GC-MS versus GC-ECD have been
made by Cronn and co-workers (1976).   In these comparisons, the detection limit
was defined based  on  a peak height three times the noise  level with a 75-ml
sample loop.  With  TCI  at an average mixing  ratio of 128  ppt, the detection
limit (GC-MS) was 23 ppt (with 3 doped samples of zero air).   The percent stan-
dard deviation was  10.   More recent data from Cronn and Harsch (1979) show a
slightly lower detection  limit  (16 ppt) for  a 100-ml freezeout sample.  The
sensitivity of the  GC-MS  technique was achieved using direct coupling of the
gas chromatograph and mass  spectrometer and selective ion  monitoring (at mass
95 for TCI).
     Pellizzari  et  al.  (1976) used Tenax GC to absorb TCI  and  other halogena-
ted hydrocarbons from  ambient  air.   Recovery was made by  thermal desorption
and helium  purging  into a freezeout trap.  The estimated detection limit when
high resolution GC-MS  is  used is 1.92 ppt (0.01 x 10  mg/m  ).  Accuracy was
reported at ±30 percent.   Included among the inherent analytical  errors  were
(1) the ability to accurately determine the breakthrough volume,  (2)  the  percent
recovery from the  sampling  cartridge  after a period of storage, and (3) the
reproducibility of  thermal  desorption from the cartridge and its introduction
into the analytical system.   To minimize loss of sample,  cartridge samplers
should be enclosed  in cartridge  holders and placed in a second container that
can be sealed, protected from light, and stored at 0°C.   The advantages repor-
ted for Tenax include (1) low water retention, (2) high thermal stability,  and
(3) low background levels (Pellizzari, 1974;  Pellizzari, 1975;  Pellizzari,  1977;
Pellizzari et al., 1976).
     In an  analytical  system using Tenax GC  as  the  absorbent, Krost et al.
(1982) reported an  estimated detection limit for TCI of 2 ppt.  Analysis  of
the breakthrough volumes at various temperatures suggests that Tenax GC is  not
a particularly effective absorbent for TCI.
     Resolution of  TCI  from other hydrocarbons was reported to be effective
when Carbopak C-HT, a  graphitized thermal  carbon black treated with hydrogen
at 1,000°C  was  used as the adsorbent  in a GC column (Knoll et al., 1979).
Flame ionization was used in detection.  Interference between TCI and ethylene
dichloride was reported with Porapak T.
                                    3-11

-------
     Gas chromatography-mass spectrometry has been more widely used in measure-
ments of volatile  organic  compounds  in aqueous samples.  In this regard, gas
purging followed by trapping on an absorbent has been extensively used in com-
bination with GC-MS detection.   A review of this system, as applied to TCI and
other volatile organics, has been made by Bellar et al. (1979).   A combination
of Tenax GC  (60/80 mesh)  and silica gel was recommended.   Purging was accom-
plished at a water temperature of about 22°C.  In general, the method is appli-
cable to compounds  that have a low solubility  in water and a vapor pressure
greater than water at ambient temperatures.  For well-defined samples when the
probability of  unexpected  compounds  is low, the EC  detector  could be used.
The detection limit of this system was reported to be in the 0.1 to 1 ug/L range.
Desorption onto the head of the column can be accomplished in one step by heat-
ing.

3.3.2  Calibration
     Both static and dynamic procedures have been used to calibrate the instru-
mentation used  for measurement of TCI.   Because of the very low  levels  of TCI
found or expected in ambient air, the generation of known low-level concentra-
tions of TCI is critical in instrument calibration.
     Singh and  coworkers  (1979,  1977) employed permeation tubes to calibrate
their GC-EC instrumentation.  A tube constructed of FEP Teflon (wall thickness
of 0.03  in.)  was  found to be satisfactory for TCI.   Errors in the permeation
rate were less than 5 percent (Singh et al., 1979).   It was considered the best,
most accurate  means of  generating primary  standards.   Since primary  standards
could not be  used  in field operations,  secondary standards were  generated and
compared to  them.   The  permeation rate at  the  95-percent  confidence  level was
246.0 ± 20.1 mg/min  (Singh et al., 1977).
     Primary and  secondary standards  prepared  by multiple dilutions  were used
by Cronn and coworkers  (1976, 1977a,b).

3.3.3  Sampling and  Sources of Error
     Common  approaches  used to sample  ambient  air for trace gas analysis are
presented below (National  Academy of Sciences,  1978).
     (1)  Pump-pressure samples:  A  mechanical pump is used to  fill a  stain-
less steel or glass  container to a positive  pressure relative to the surround-
ing atmosphere.
                                    3-12

-------
     (2)  Ambient pressure samples:   An evacuated chamber is opened and allowed
to fill until  it has reached ambient pressure at the sampling  location.  If
filling is conducted at high altitude, the container may easily become contami-
nated when returned to ground level.
     (3)  Adsorption on molecular sieves, activated charcoal, or other adsor-
bents.
     (4)  Cryogenic samples:  Air is  pumped into a container and  liquefied,
and a partial  vacuum is created which allows more air to enter.  This method
allows for the collection of several  thousand liters of air.
     Singh and coworkers (1979) employed electropolished stainless steel  vessels
in their  sampling  of tropospheric  air.   Vessels were checked for  background
contamination until it was reduced to less than 2 to 3 percent of the expected
background concentration of a given trace substance.  Typically, vessels that
maintained positive pressure were not affected by room contamination.
     Sampling and  storage of ambient  air collected at stratospheric  altitudes
present more of a problem.   Because  of the low pressures in the sampling vessels,
contamination may  easily  result when  vessels are returned  to  ground level.
Cronn and Robinson (1978) and Harsch and Cronn (1978) reported that contamina-
tion of low-pressure  stainless  steel  vessels precluded use of  data  for TCI.
Whole-air samples  were  collected at  stratospheric altitudes.    Contamination
occurred during the sample-transfer  procedure because of high  mixing ratios  of
halocarbons in laboratory air.
     To overcome contamination of low-pressure samples,  Harsch and Cronn (1978)
developed a  low-pressure sample-transfer technique.  A  vacuum transfer proce-
dure employing a  freezeout  loop (Rasmussen et al., 1977) was found  to yield
results with  sufficient sensitivity  and with minimal  contamination problems.
A leak-tight  system was considered essential.  System components were silver-
soldered.
     Sample collection and handling  and preservation of water  samples have been
discussed by  Kopfler et al. (1976) and  by Bellar et al. (1974).  Narrow-mouth
vials with a total volume in excess  of 50 ml were considered acceptable (Bellar
et al., 1974).  It was recommended that the vials be filled to overflowing and
each covered with a Teflon-faced silicone rubber septum.  Because of the pres-
ence of chlorine in some drinking water supplies and the chlorination of organic
materials, an increase in the concentration of some halocarbons may occur also
during storage (Kopfler et al., 1976).
                                    3-13

-------
     Bellar et al.  (1974) reported that free chlorine can result in the forma-
tion of TCI and other halohydrocarbons.  In measurements of TCI in water from
a sewage-treatment plant, they found an increase of 1.2 ug TCI/L in the water
after chlorination.   Dietz and Singley (1979) collected water samples in 125-mL
serum vials that had  been cleaned by detergent wash, distilled water  rinse,
dichromic acid wash, distilled water rinse, acetone rinse, hexane rinse, ace-
tone rinse, distilled water rinse, and oven-drying at 150°C.  Vials were com-
pletely filled and then sealed with Teflon-faced septa and aluminum crimp seals.
Samples were stored at 4°C.   Minimum loss  occurred during storage at 4°C up to
28 days.
     The stability of  stored  samples  of TCI present in ambient air or breath
samples from  exposed  individuals  has  been  investigated  by  Pasquini  (1978).
The mean percent  loss  of TCI in  alveolar  breath  samples collected in glass
tubes at room  temperature was 59.3 ± 15.3  percent over 169 hours.  The mean
decay rate  of  TCI in dilution air samples was 4.7 ± 3.2 percent over the same
time interval.  Moisture, temperature, and tube surface and condition can alter
greatly vapor  retention.  For TCI, tubes stored at 37°C had high vapor reten-
tion rates.  When breath was exhaled  into the tubes at 37°C, water vapor was
not condensed from the sample.  Additional  experiments with TCI indicated that
si 1 iconized tubes showed a lower decay rate than non-si 1 iconized tubes.
     Renberg (1978) judged XAD-4  the most  suitable of several possible adsor-
bents for volatile halogenated hydrocarbons in water, including TCI.  This method
results in  an  extract concentrated enough  for both chemical determination  and
small-scale biological tests.
     Kummert et al.  (1978)  have  suggested  a direct  aqueous  injection high-
pressure liquid chromatography technique for analysis of TCI in natural waters.
The technique  has been  successfully applied to an actual water  pollution case
for tetrachloroethylene.  The detection limit for TCI was reported as 1 umol/£
for a 10-ml sample.
3.3.3.1  Air-Mixing Ratios.   Measurements  of ambient  air mixing ratios of  TCI
have been made at a variety of sites in the United States and abroad. As shown
in Table 3-7,  urban areas have mixing ratios in the low ppb range.
     Singh et  al. (1979)  reported  an average mixing ratio in the northern  hemi-
sphere  of  11  to 17 ppt,  with the average  in the  southern  hemisphere  of less
than 3  ppt.   Over the  Pacific Northwest, an average  tropospheric  mixing ratio
of 20 ± 3.9 ppt was found (Cronn  et al., 1977a).  Harsch (1977) reported levels
                                    3-14

-------
TABLE 3-7.   AMBIENT AIR-MIXING RATIOS OF TRICHLOROETHYLENE
Location
ARIZONA
Phoenix
Grand Canyon*
ARKANSAS
Helena
El Dorado*
CALIFORNIA
£ Stanford Hills
Ul
Los Angeles
Los Angeles
Point Reyes
Palm Springs
Yosemite
Mill Valley
Riverside
Menlo Park
Upland
Badger Pass
Date of
measurements
04/23-05/06 1979
11/28-12/05 1977
11/30 1976
06/13-07/08 1978
11/23-11/30 1975
04/28-05/04 1976
04/09-04/21 1979
12/01-12/12 1975
05/05-05/11 1976
05/12-05/18
01/11-01-27 1977
04/25-05/04 1977

08/13-09/23 1977
05/05-05/13 1977
Mixing ratio (ppb)
Maximum Minimum Average
3.0696
0
< 0.03
13
3.09
1.77
1.702
0.064
0.446
0.022
0.090
0.476

0.64
0.016
0.0117
0
< 0.03
0.13
0.006
0.026
0.363
0
0.006
0.008
0.009
0.057

0
0.001
0.4835 ±
0
< 0.03
3.1 ± 3.
0.093 ±
0.315 ±
0.399 ±
0.024 ±
0.035 ±
0.015 ±
0.026 ±
0.266 ±

0.080 ±
0.012 ±
.5869


8
0.32
.310
.302
0.014
0.050
0.003
0.094
0.094

0.22
0.002
Reference
Singh et al. , 1981
Pellizzari, 1979a
Battell e, 1977
Pellizzari, 1979a
Singh et al., 1979a
Singh et al. , 1979a
Singh et al., 1981
Singh et al. , 1979
Singh et al., 1979
Singh et al., 1979
Singh et al., 1979
Singh et al., 1979
Singh et al., 1979
Pellizzari, 1979a
Pellizzari, 1979a

-------
TABLE 3-7.   (continued)
Location
Point Arena
Point Arena
San Jose
Oakland
La Jolla*
DELAWARE
Delaware City
£ LOUISIANA
CT)
Baton Rouge*
Lake Charles
MARYLAND
Baltimore
NEW JERSEY
Rutherford*
Somerset *
South Amboy
Date of
measurements
05/23-05/30 1977
08/30-09/05 1978
08/21-08/27 1978
06/28-07/10 1979

07/08-07/10 1974
11/18 1976
05/20 1977
12/06 1976
06/26 1978
07/11-07/12 1974
05/01 1978
12/29 1979
07/18-07/26 1978
01/27 1979
12/29 1979
Mixing ratio
Maximum Minimum
0.012
0.034
2.192
1.5582
3.9
0.56
0.65
2.1
< 0.05
8.6
0.28
8.5
0.010
0.008
0.048
0.0141
0
< 0.05
0
0.073
< 0.05
0
0
0
(ppb)
Average
0.011 ± 0.001
0.017 ± 0.007
0.417 ± 0.425
0.1876 ± 0.2697
1.3 ± 1.2
0.35
0.4 ± 0.23
1.6 ± 0.79
_
1.15
0.17 ± 0.17
0.08 ± 0.28
0.4 ± 1.4
Reference
Pellizzari, 1979a
Pellizzari, 1979a
Pellizzari, 1979a
Singh et al., 1981
Su and Goldberg, 1976
Lillian et al., 1975
Pellizzari et al., 1979
Battelle, 1977






>
Battelle, 1977; Pellizzari,
1979b
Pellizzari et al., 1979
Bozzelli and Kebbekus,
Bozzelli and Kebbekus,
Bozzelli and Kebbekus,

1983
1979
1983

-------
                                              TABLE 3-7.   (continued)
co
Location
Batso
Seagirt
Newark
Sandy Hook
Elizabeth
Fords*
Middlesex*
Edison*
Carlstadt*
Camden
Bayonne
Boundbrook*
Bridgewater*
Rahway*
Date of
measurements
02/26 1979
12/29 1979
06/18-06/19 1974
03/23 1976
12/29 1979
07/02-07/05 1974
09/15 1978
12/29 1979
03/26 1976
09/27 1978
07/23-07/28 1978
03/24 1976
09/24 1978
09/28-09/30 1978
04/03-10/24 1979
March-Dec. 1973
03/26 1976
09/20 1978
07/17 1978
08/05 1978
09/20-09/22 1978
Mixing ratio (ppb)
Maximum Minimum Average
1.3
2.8
1.9
0.8
6.4
3.1
0.36
6.1
9
3.5
8.8
4.1
6.9
18
0
< 0.05
0
< 0.05
0
0
0
0.54
3
0
< 0.05
0
0
3.4
0.08 ± 0.28
0.26
0.25
0.34
0.76
1.9 ± 1.1
0.22 ± 0.11
2.7 ± 2.1
6.4 ± 3.1
0.42 ± 0.87
0.92
2.4 ± 1.8
0.35 ± 0.29
11 ± 7.8
Reference
Bozzelli and Kebbekus, 1983
Lillian et al., 1975
Bozzelli and Kebbekus, 1983
Lillian et al. , 1975
Bozzelli and Kebbekus, 1983
Pellizzari, 1977;
Pellizzari et al., 1979
Bozzelli and Kebbekus, 1979
Pellizzari, 1978;
Pellizzari, 1979;
Pellizzari, 1977b
Pellizzari et al. , 1979
Bozzelli and Kebbekus, 1983
Bozzelli and Kebbekus, 1983
Pellizzari, 1977;
Pellizzari et al. , 1979
Bozzelli and Kebbekus, 1979
Pellizzari et al. , 1979

-------
                                                 TABLE 3-7.  (continued)
co
i
00
Location
NEW YORK
NYC
Whiteface Mtns.
Niagara Falls/Buffalo
OHIO
Wilmington
NEVADA
Reese River
KANSAS
Jetmar
WASHINGTON
Pullman
Pacific Northwest
PACIFIC OCEAN
(lat 37°N)
PANAMA CANAL ZONE
Date of
measurements
06/27-06/28 1974
09/16-09/19 1974
Fall, 1978

07/16-07/26 1974
05/14-05/20 1977
06/01-06/07 1978
December 1974-
February 1975
March 1976
April 1977

Mixing ratio (ppb)
Maximum Minimum Average
1.1 0.11 0.71
0.35 < 0.05 0.10
0.11 0.05 0.10

0.63 < 0.05 0.19
0.016 0.001 0.012 ± 0.002
0.021 0.007 0.013 ± 0.004
< 0.005
0.020 + 0.0039
0.0124

Reference
Lillian et
Lillian et
Pellizzari

Pellizzari
Singh et al
Singh et al

al., 1975
al., 1975
et al . , 1979

et al., 1979
., 1979
., 1979
Grimsrud and
Rasmussen, 1975
Cronn et al
Cronn et al

., 1977a
., 1977b

     (lat 9°N)
July 1977
0.015
Cronn and Robinson,

1978

-------
TABLE 3-7.  (continued)
Location
NORTH ATLANTIC OCEAN
WESTERN IRELAND
ENGLAND
JAPAN
Tokyo
Various sites
WESTERN EUROPE
£ COLORADO
VO
Denver*
MISSOURI
St. Louis*
TEXAS
Aldine
El Paso
Freeport
Houston
Date of
measurements
October 1973
June-July 1974
March 1972
09/10-10/27 1975
Summer, 1979
1976
06/16-06/26 1980
05/30 1980
06/08 1980

06/22 1977
10/20 1977
04/05 1978
05/01 1978
08/09 1976
11/10 1976
07/27 1976
05/24/1980
Mixing ratio
Maximum Minimum
.005 4 x 10"4
47.1 0.8
0.038 0.005
8.5 < 0.02
0.41 0.028
0.24 0.019

0.029 0
0.3 0
0.029 0
0.33 0
(ppb)
Average
< 0.005 ± 0.0657
0.015 ± 0.0121
.002
3.5 ± 3.7
0.014 ± 0.010
0.2 ± 0.11
0.11 ± 0.063

0.009 ± 0.017
0.15 ± 0.11
0.001 ± 0.004
0.12 ± 0.098
Reference
Lovelock, 1974a
Lovelock, 1974a
Murray and Riley, 1973a
Tada et al . , 1976
Makide, et al. , 1980
Correia et al. , 1977
Singh et al. , 1980
Singh et al., 1980

Pellizzari et al., 1979
Pellizzari, 1979a
Battelle, 1977b;
Pellizzari et al. , 1979
Pellizzari et al., 1979;
Singh et al. , 1980

-------
                                                     TABLE 3-7.  (continued)
u>
i
ro
O
Location
WASHINGTON
Auburn*
WEST VIRGINIA*
Nitro
West Belle
VIRGINIA*
Front Royal e
Date of
measurements
01/10-01/11 1977
09/27 1977
11/18 1977
09/27 1977
11/18 1977
09/29 1977
11/16 1977
Mixing ratio
Maximum Minimum
0.83 0.19
0.067 0
0 0
0.009 0
(ppb)
Average
0.64 ± 0.29
0.032 ± 0.035
0
0.003 ± 0.004
Reference
Battelle, 1977b
Pellizzari, 1978
Pellizzari, 1978
Pellizzari, 1978
    *Data obtained from summary report of Brodzinsky and Singh, 1982.

-------
between 22 ppb  and 8.7 ppb in more  than 20 different indoor  environments.
Brodzinsky and Singh (1982), in their review of monitoring data, reported that
background concentrations average about 30 ppt.  A maximum level of 87 ppb was
reported for Newark, New Jersey.
3.3.3.2  Water.  Trichloroethylene has been detected  in  a  number of raw water
sources, finished drinking waters, and subsurface waters in the United States.
The occurrence  of  TCI  in waters of the United  States  has been  reviewed by the
U.S. Environmental  Protection Agency (1979).
3.3.3.2.1  Drinking water.   Dowty et al.  (1975) detected TCI in finished drink-
ing water in the New Orleans area.   It was  reported to pass through the water
treatment plant without  alteration.   It  was also reported to be incompletely
removed after passage  of  water through a charcoal filtering unit.   Detection
was made by GC-MS.
     In a comprehensive surface water monitoring  effort  for TCI, Ewing et al.
(1977) have shown  TCI  to be a common low-level (ppb) contaminant of  surface
waters in the vicinity of urban/industrial areas.   TCI was reported at concen-
trations of 1 |jg/L (ppb)  or greater in  72  of  179 surface  water samples from
U.S. urban/industrial  areas  from  23  states (Table 3-8).   Ewing also reported
samples of drinking water which were not  included in this total.
     A concentration of 8.5 ppb was found in the drinking water of the town of
Grand Island, New  York (Coleman  et al.,   1976)  using  headspace GC-EC.  In 10
lake water samples  from this area, an average  concentration of 11.9 ppb ±2.5
percent was found.   Coleman et al.  (1976) detected TCI in the drinking water
of  four cities  in  1975.   The concentrations were:  0.1 ppb (Cincinnati);  0.3
ppb (Miami);  <0.1 ppb (Ottumwa, Iowa);  and 0.5 ppb (Philadelphia).   No TCI was
detected in drinking water  samples for Seattle.   Identification and quantifi-
cation were made by GC-MS using the purge/trap technique.
     As evidenced by the differences reported in the individual studies above,
the frequency of contamination and the amount of TCI  found in potable waters
depend in part on the source of water being monitored.
     The surface water supplies of 133 cities using surface water supplies have
been sampled during federal  surveys.   Thirty-two percent of the finished waters
serving these cities were  found  to contain TCI.  The  concentration of TCI in
these water supplies ranged from  0.06 to  3.2 ug/L.  The  average concentration
in positive water supplies was  0.47 ug/L.  Thirty percent of all  supplies  con-
tained a level  of  TCI  which was  below I  ug/L.  The remaining two percent of
the supplies contained a concentration between 1 to 4  ug/L.

                                    3-21

-------
                TABLE 3-8.  MEAN LEVELS OF TCI IN SURFACE WATERS'
Surface water samples
State
A1 abama
Delaware
Illinois
Indiana
Kentucky
Louisiana
Michigan
Minnesota
New Jersey
New York
Pennsylvania
Tennessee
Texas
West Virginia
All Sites
Positive
for TCI
1
4
17
1
1
1
1
2
14
10
14
1
4
1
72
Total
7
10
25
4
3
7
8
9
15
20
21
6
8
3
179
Percent
positive
for TCI
14
40
68
25
33
14
13
22
93
50
67
17
50
3
40
Mean levels
(ug/L)
All b
samples '
0.5
1.2
0.6

0.7
0.6
23.9
1.9
4.0
1.6
2.3
0.9
1.2
0.7
2.7
Positive
samples only
1.0
2.3
3.2
1.0
1.0
1.0
188
7
4.3
2.6
3.2
3.0
1.8
1.0
5.9
aSource:   Ewing et al.  (1977);  drinking water samples not reported in these data.

bMean levels were calculated by substituting a value of one-half the reporting
 limit of 1 ppb (1 ug/L) for samples reported as having non-detectable (i.e.,
 less than 1 ug/L).
GLevels of TCI in sample from the following states (number of sites per state
 in parentheses) were reported as not detectable (i.e., less than 1 ug/L):

 California (9)                         North Carolina (1)
 Iowa (2)                               Oregon (3)
 Mississippi (1)                        Washington (2)
 Missouri (3)                           Wisconsin (6)
                                    3-22

-------
     In 25 cities that draw their water from underground sources, water supplies
were sampled for  TCI  during the NOMS, NORS,  SRI  and EPA Region V  surveys.
Thirty-six percent of these finished waters contained detectable concentrations
of this chemical.   The  median level detected was 0.31 pg/L with a range from
0.11 to 53.0 ug/L.
     The Community Water Supply Survey examined an additional 330 ground water
systems selected randomly from across the United States.   Four percent of these
systems contained a detectable  level of TCI.  Seventy percent of these detec-
table levels of TCI were between 0.5 to 5 ug/L.   TCI was found in finished water
from small and large systems alike.
     In analyses  of drinking waters  from sites  near  producer and user facili-
ties and  from  a  background site, Battelle (1977) found the following concen-
trations in 1977:   19 ppb (Freeport, Texas); 0.4 ppb (Baton Rouge,  Louisiana);
0.1 ppb  (Lake  Charles,  Louisiana);  32 ppb (Des Moines, Iowa);  and 22  ppb
(Helena, Arkansas).
     Boateng et al.  (1984) have described the geology,  hydrology, and chemistry
of both shallow and deep aquifers in relation to the occurrence of  TCI in two
production wells.
3.3.3.2.2  Rivers, treatment plant effluents.   Up to 403 ppb TCI was detected
by Battelle in surface  waters near specific manufacturing facilities (1977).
Trichloroethylene was detected  in  both  industrial  wastewater and river water
receiving the waste (Jungclaus et al.,  1978).
     Murray and Riley (1973a) detected TCI in surface sea water in the eastern
Atlantic ocean.
     Bellar et al. (1974) found an increased concentration of TCI after chlori-
nation treatment  of  the effluent from a  sewage-treatment  facility.  Before
chlorination,  the concentration detected  by GC-MS was 8.6 ppb;  after chlorina-
tion, the concentration increased to 9.6  ppb.
     Case reports describing TCI contamination of well  waters and water supplies
have been reviewed elsewhere  (U.S.  EPA, 1979).  Concentrations as high as 330
ppb have been reported.
     Correia et al.  (1977)  measured TCI  in river,  canal, and sea waters  that
receive effluent  from production and user sites.   Values ranged from 0.02 to
74 ug/L (ppb w/w), showing the effect of  these additions on background levels.
Dilution effects  and losses to the surrounding air were noted.
                                    3-23

-------
3.3.3.3  Soil and Sediment.   Battelle  (1977)  reported the presence of TCI in
soil samples in  the  vicinity  of producers and users at levels up to 5.6 ppb.
In  freshwater  sediment taken  near  the Hooker  Chemical  facility at Taft,
Louisiana, up to  300  ppb  TCI  was detected (Battelle, 1977).  Trichloroethy-
lene was detected in the soil  in the Love Canal  area in Niagara Falls, New York
(U.S. EPA, 1979).
     Pearson and McConnell (1975) found up to 9.9 ppb in sediment from Liverpool
Bay, England.  Murray and Riley (1973b) found TCI in river mud.
     Rogers and McFarlane (1981) reported that aluminum (Al  )-saturated mont-
morillonite clay sorbed about 17 percent of the TCI applied (100, 500, and 1,000
ppb).  The organic carbon content was assumed to be zero.   There was no apparent
absorption when calcium-saturated montmorillonite clay was used.  With two silty
clay loams  (organic  carbon  content = 2 percent), less than 6 percent  of  the
TCI available was sorbed.   When soil sorption was normalized,  based on the soil
organic  carbon  contents,  a correlation was  found  between water solubility,
sorption constant, and the partitioning coefficient.
3.4  ECOLOGICAL EFFECTS
     Trichloroethylene  has  been  tested for acute  toxicity  in  a  limited  number
of  aquatic  species.   This  section presents observed levels reported to  result
in  adverse  effects under laboratory conditions.  Such parameters of toxicity
are  not easily extrapolated  to  environmental  situations.   Test populations
themselves  may not be representative of the entire  species  in which  susceptibil-
ity  of  various lifestages to  the  test  substance may vary considerably.
     Guidelines for the  utilization of these data in the development of criteria
levels  for  TCI in  water  are discussed  elsewhere (U.S. EPA,  1980).
     The  toxicity  of TCI to  fish and  other aquatic organisms has been  gauged
principally by  flow-through and  static testing methods  (Committee  on Methods
for Toxicity with Aquatic Organisms,  1975).   The flow-through  method exposes
the organism(s)  continuously to  a constant concentration  of  TCI while  oxygen
 is  continuously replenished and  waste  products are  removed.   A  static test,  on
 the other hand, exposes  the organism(s) to the added initial  concentration only.
 Both types of tests are commonly used as initial  indications of the potential
 of  substances to  cause  adverse effects.
                                     3-24

-------
     Alexander et al.  (1978)  used both  flow-through  and  static  methods  to  in-
vestigate the acute  toxicity  of several chlorinated solvents, including TCI,
to adult  fathead  minnows  (Pimephales promelas).   Studies  were  conducted in
accordance with test  methods  described by the  U.S.  Environmental  Protection
Agency (1975).  The  lethal concentration of each solvent tested to produce  50
percent mortality (LCcn) was lower in the flow-through than in the static-type
experiment.   The  flow-through method was considered  the  better  of  the two  for
volatile  solvents such  as TCI.  The 96-hr  LC50  concentration  for TCI was 40.7
mg/L in the flow-through test.  The 95 percent confidence band was between 31.4
and 71.8 mg/L.  Of four solvents tested (perchloroethylene, methylene chloride,
methyl chloroform, and TCI), TCI had the second highest toxicity (after perchlo-
roethylene).
     In acute static  experiments  with bluegills,  TCI was  reported to have  a
96-h LC5Q value of 44.7 mg/L (U.S.  EPA, 1978b).   With Daphnia magna as the test
organism  in a 48-hr  static  experiment, the LC5Q for TCI was 85.2 mg/L.   In a
chronic test with this  organism, no adverse effects  were found  at  the highest
test concentrations  of 10 mg/L.
     Lay et al.  (1984) evaluated the effect of TCI  on Daphnia magna and phyto-
plankton  species  in  a natural  pond.   The  pond  was compartmentalized and two
compartments  each were  used  for a single high dose of 110 ng/L, a single low
dose of 25 ng/L,  and controls.   Complete elimination of Daphnia magna in the
high dose groups was noted at 24 and 72 hours  after dosing.  In the low dosage
groups, TCI was less toxic to Daphnia.   However, a  significant decrease in pop-
ulation was noted in both treated compartments from day 3 to day 28 of the ex-
periment.   The  relative abundance  and absolute number  of  all phytoplankton
individuals counted  increased with the increase of  TCI concentrations.
     Investigation of the bioconcentration potential  of TCI for bluegill indi-
cated that the  bioconcentration factor (BCF)  was 17.  The half-life of this
compound  in tissues  was less  than 1 day  (U.S.  EPA,  1980;  U.S.  EPA,  1978b).
These data suggest  that  residues  of TCI  at  exposure  concentrations  below
those directly toxic to aquatic life are not likely (U.S. EPA, 1980).
     Few data are available  relating  to effects of  TCI  on saltwater aquatic
life.  Borthwick  (1977) exposed sheepshead minnows and grass  shrimp to 20 and
2 mg/L, respectively,  and noted erratic swimming,  uncontrolled movement,  and
loss of equilibrium  after several  minutes.
                                    3-25

-------
3.5  CRITERIA, STANDARDS, AND REGULATIONS
     The current occupational standard for TCI established by the Occupational
Safety and Health Administration  in 1972 (General Industry Standards, OSHA,
                           3
1978) is 100  ppm  (525 mg/m ) as a  time-weighted  average (TWA) over an 8-hr
period.   More  recently,  the National Institute for  Occupational  Safety and
Health (NIOSH) (Page and Arthur, 1978) has noted that the current standard ap-
pears to be inadequate for  protection of workers  from the carcinogenic poten-
tial of TCI.   A level of 25 ppm, as a TWA, was the recommended standard proposed
by NIOSH, based on  engineering  control capabilities.  The American  Council  of
Governmental  Industrial Hygienists has established a TLV® of 50 ppm.
                                    3-26

-------
3.6  REFERENCES
Alexander, H. C. ,  W.  M.  McCarty, and  E.  A.  Bartlett.  1978.  Toxicity of per-
     chloroethylene, trichloroethyl ene,  1,1,1-trichloroethane, and  methylene
     chloride to fathead minnows.  Bull. Environ. Contam. Toxicol. 20:344-352.

Altshuller, A.  P.  1980. Lifetimes of organic molecules in the troposphere and
     lower stratosphere.   Adv. Environ. Sci. Technol.  10:181-219.

Battelle Columbus Laboratories. 1977.  Environmental Monitoring Near  Industrial
     Sites:  Trichloroethylene.  Prepared for the U.S. Environmental  Protection
     Agency, PB 273203, August.

Beckers, N.  L.  1972.  Stabilization  of Chlorinated  Hydrocarbons.   Patent No.
     3,682,830,  U.S. Patent Office,  Washington, DC, August 8.

Bellar, T. A.,  W.  L.  Budde,  and J.  W.  Eichelberger. 1979. The Identification
     and Measurement of Volatile Organic Compounds in Aqueous Environmental
     Samples.   Chapter 4.  In:   Monitoring Toxic Substances.   D.  Schvetzle,
     ed.,  ACS Symposium Series 94, American Chemical Society.

Bellar, T. A.,  J.  J.  Lichtenberg, and R.  C.  Kroner.  1974.  The occurrence of
     organohalides  in chlorinated drinking  waters.  J.  Am.  Water Works Assoc.
     66: 703-706.

Boateng, K. P.,  P. C.  Evers,  and S.  M.  Testa. 1984. Ground water contamination
     of two production wells: a case history.  Ground Water Mom't. Rev. 4:24-31.

Bonse, G., and D. Henschler.  1976. Chemical reactivity, biotransformation, and
     toxicity of polychlorinated aliphatic compounds.  CRC Crit. Rev.  Toxicol.
     4:395-409.

Borthwick, P. W. 1977.  Results of toxicity tests with  fishes  and macroinverte-
     brates.  Data sheets available from U.S. Environmental Protection Agency,
     Environmental Research Laboratory, Gulf Breeze, FL.

Bozzelli,  J. W.  and B.  B.  Kebbekus.  1979. Analysis of  Selected Volatile Organic
     Substances in Ambient Air.  Prepared for New Jersey  Department  of Environ-
     mental Protection by New Jersey Institute of Technology, NTIS PB  80-944694.

Bozzelli,  J.  W.  and B.  B. Kebbekus. 1983.  Volatile Organic Compounds in the
     Ambient Atmosphere of the New  Jersey,  New  York  Area.   Prepared for the
     U.S.   Environmental  Protection  Agency  by the  New Jersey Institute  of
     Technology.

Brodzinsky, R. and H.  B.  Singh. 1982.  Volatile Organic Chemicals in  the Atmos-
     phere:  An assessment of Available Data.  Final Report,  SRI International,
     Menlo Park,  California,  prepared  for the U.S.  Environmental  Protection
     Agency.

Bucon, H.  W. , J.  F. Macko, and H. J.  Taback. 1978. Volatile  Organic  Compound
     (VOC) Species  Data Manual.   EPA-450/3-78-119.   Report prepared for the
     U.S.  Environmental Protection Agency, December.
                                    3-27

-------
Chemical and Engineering News. 1975. 19 May.  pp. 41-43.

Chemical Marketing  Reporter.  1978.  Chemical  Profile for Trichloroethylene,
     26 June.

Coleman, W. E.,  R.  D. Lingg, R.  G. Melton, and F. C. Kopfler. 1976. The Occur-
     rence  of Volatile Organics  in  Five  Drinking Water Supplies  Using  Gas
     Chromatography/Mass Spectrometry.  Chapter  21.   In:   Identification and
     Analysis of Organic  Pollutants in Water.   L.  H.  Keith,  ed. ,  Ann Arbor
     Science.

Committee  on Methods  for  Toxicity Tests with Aquatic Organisms:   Methods for
     Acute  Toxicity Tests  with  Fish, Macroinvertebrates,  and Amphibians.
     1975.  Ecol. Res. Series, EPA-600/3-75-009.

Correia, Y., G.  J.  Martens, F. H. van Mensch, and B. P. Whim. 1977. The occur-
     rence of trichloroethylene, tetrachloroethylene, and 1,1,1-trichloroethane
     in Western Europe in air and water.  Atmos. Environ. 11:1113-1116.

Cronn,  D.  R., and D.  E. Harsch.  1979.  Determination of atmospheric halocarbon
     concentrations  by gas chromatography-mass  spectrometry.  Anal.  Lett.
     12(B14):  1489-1496.

Cronn,  D.  R. , and  E. Robinson.  1978. Determination  of  Trace  Gases in  Learjet
     and V-2 Whole-air Samples  Collected  During  the Intertropical  Convergence
     Zone  Study.   Report  submitted to NASA,  Washington State University,
     August.

Cronn,  D.  R. ,  R.  A. Rasmussen,  and E.  Robinson. 1976.  Report  for Phase I.
     Report prepared for  U.S.E.P.A.,  Contract  No.  R0804033-01,  Washington
     State University, 23 August.

Cronn,  D.  R., R. A.  Rasmussen, E. Robinson, and  D.  E.  Harsch. 1977a. Halogenated
     compound identification  and measurement in  the troposphere and low stra-
     tosphere.  J.  Geophy.  Res. 82:5935-5943.

Cronn,  D.  R., R. A.  Rasmussen,  and E.  Robinson.  1977b.  Measurement of Tropos-
     pheric Halocarbons  by Gas  Chromatography/Mass  Spectrometry.  Report for
     Phase  II of EPA Grant  No. R0804033-02, October.

Crutzen, P.  A., I.  S. A.  Isaksen,  and J.  R.  McAfee. 1978.  The impact of the
     chlorocarbon industry on the  ozone  layer.   J.  Geophy.  Res.  83:345-362.

Cundy,  J.  M. 1976.  Trichloroethylene  in  anesthetic practice.   Anesthesia
     31:950 (Letter).

Dahlberg,  J.  A. 1969. The  non-sensitized photooxidation of  trichloroethylene
     in air.  Acta  Chem.  Scand.  23:3081-3090.

Derwent,  R.  G., and A. E. J.  Eggleton. 1978.  Halocarbon lifetimes and concen-
     tration  distributions calculated  using a  two-dimensional  tropospheric
     model.  Atmos.  Environ.  12:1261-1269.

Dietz,  E.  A., and K.  F. Singley.  1979.  Determination of chlorinated hydrocarbons
      in water  by  headspace gas  chromatography.   Anal. Chem. 51:1809-1814.


                                     3-28

-------
Dilling, W.  L.,  C.  J.  Bredeweg,  and N.  B.  Tefertiller.  1976.  Simulated atmos-
     pheric photodecomposition  rates of methylene chloride,  1,1,1-trichloro-
     ethane,  trichloroethylene,   tetrachloroethylene,  and  other  compounds.
     Environ. Sci. Techno!. 10:351-356.

Dilling, W.  L. ,  N.  B.  Tefertiller,  and G. J. Kallos. 1975. Evaporation  rates
     and reactivities of methylene chloride, chloroform, 1,1,1-trichloroethane,
     trichloroethylene, tetrachloroethylene,  and other chlorinated  compounds
     in dilute aqueous solutions.  Environ. Sci. Technol. 9:833-838.

Dimitriades, B.,  B.  W.  Gay, Jr.,  R.  R.  Arnts,  and R.  L.  Seila.  1983.  Photo-
     chemical reactivity of perchloroethylene:  A  new appraisal.   0. Air Poll.
     Cont.  Assoc.  33(b):575-587.

Dowty, B.  J. ,  D.  R. Carlisle, and J.  L.  Laseter.  1975.  New Orleans drinking
     water sources tested by gas chromatography-mass spectrometry.   Occurrence
     and origin of aromatics and halogenated aliphatic hydrocarbons.   Environ.
     Sci. Tech. 9:762-765.

Edney, E.,  S. Mitchell, and J. Bufalini. 1983. Atmospheric  chemistry of  several
     toxic compounds.  EPA-600/S3-82-092, U.S. Environmental  Protection  Agency.

Ewing, B.  B. ,  E.  S. K. Chian, J.  C. Cook, C. A.  Evans, P. K. Hopke,  and E.
     G. Perkins.  1977.  Monitoring to Detect Previously Unrecognized  Mutants  in
     Surface Waters.   EPA-560/6-77-015.   Report  prepared by the  University of
     Illinois for the U.S. Environmental Protection Agency, July.

Farm Chemicals Handbook,  1976.   Cited in:  Waters, E. M.,  H.  B.  Gerstner, and
     J.  E.  Huff.  1977. Trichloroethylene.   I.   An  overview.  J. Toxicol.
     Environ. Health 2:671-707.

Federal Register. 1977. 42(131):35314-35316, 8 July.

Fuller,  B.  B.  1976. Air  pollution assessment of  trichloroethylene.   MITRE
     Technical Report No.  MTR-7142.

Gay, B.  W. ,  P.  L. Hanst,  J. J. Bufalini, and R. C. Noonan. 1976.  Atmospheric
     oxidation of chlorinated ethylenes.  Environ.  Sci.  Technol. 10:58-67.

General  Industry  Standards.  1978.  OSHA Safety and Health  Standards, (29 CFR
     1910)  United States  Department of Labor Occupational  Safety and  Health
     Administration OSHA 2206, revised November 7.

Graedel, T.  E. ,  and D.  L. Allara. 1976. Tropospheric halocarbons:   Estimates
     of atmospheric chemical production.  Atmos. Environ. 10:385-388.

Grimsrud, E. P.,  and R. A. Rasmussen. 1975. Survey and analysis of halocarbons
     in the atmosphere by gas chromatography-mass  spectrometry.  Atmos.  Environ.
     9:1014-1017.

Hardie, D.  W.  F.  1964.  Chlorocarbons and Chlorohydrocarbons:  Trichloroethy-
     lene.   In:   Kirk-Othmer's  Encyclopedia  of Chemical Technology.   Vol. 5.
     Second edition, pp.  183-195.
                                    3-29

-------
Harsch, D. 1977.  Study of Halocarbon Concentrations  in  Indoor Environments.
     Report  to  U.S.   Environmental  Protection  Agency.   Contract  No.
     WA-6-99-2922-J.  July 7.

Harsch, D. E. ,  and  D.  R. Cronn. 1978.  Low-pressure sample-transfer technique
     for analysis of stratospheric  air  samples.   J.  Chromat.  Sci.  16:363-367.

Jungclaus, G.  A., V. Lopez-Avila, and R.  A.  Hites.  1978.  Organic compounds in
     an industrial  wastewater.   A case study  of  their environmental  impact.
     Environ. Sci. Technol.  12:88-96.

Knoll, J.  E.,  M.  A. Smith,  and  M. Rodney  Midget.  1979.  Evaluation of emission
     test methods for  halogenated hydrocarbons.   Vol.  I.   EPA-600/4-79-025,
     March.

Kopfler,  F.  C. ,  R. G.  Melton,  R. D. Lingg,  and  W. E. Coleman.  1976.  In:
     Identification and Analysis of Organic  Pollutants in Water.   L. W.  Keith,
     ed., Ann Arbor Science, pp. 87-104.

Krost,  K.  J. ,  E.  D.  Pellizzari, S. G.  Walburn,  and S. Hubbard.  1982
     Collection  and analysis of  hazardous organic emissions.   Anal.  Chem.
     54:810-817.

Kummert,  R. ,  E.  Molnar-Kubica,  and W.  Giger. 1978. Trace  determination of
     tetrachloroethylene  in natural  water  by direct  aqueous  injection,
     high-pressure  liquid chromatography.  Anal.  Chem. 50:1637-1639.

Lay, J.  P.;  W.  Schauerte, and W.  Klein.  1984. Effects of trichloroethylene on
     the  population dynamics of phyto- and  zooplankton  in compartments of a
     natural pond.  Environ. Pollut. 33:75-91.

Lillian,  D. ,  H.  B.  Singh,  A.  Appleby,  L.  Lobban,  R.  Arnts,  R.  Gumpert,
     R. Hague,  J. Toomey,  J. Kazazis,  M. Antell,  D.  Hansen,  and B.  Scott.
     1975. Atmosphere  fates of  halogenated compounds.   Environ. Sci.  Technol.
     9:1042-1048.

Lillian,  D. ,  H.  B.  Singh,  and  A. Appleby. 1976.  Gas  chromatographic  analysis
     of  ambient halogenated compounds.   J.  Air  Pollut.  Control  Assoc.  26:
     141-143.

Lovelock,  J.  P.  1974a.  Atmospheric  halocarbons  and  stratospheric  ozone.
     Nature 252:  292-294.

Lovelock,  J.  F. 1974b. The  electron capture detector:  Theory and practice.
     J. Chromat.  99:3-12.

Makide, Y.,  Y.  Kanai,  and T. Tominaga.  1980.  Background atmospheric  concentra-
     tions  of  halogenated  hydrocarbons  in  Japan.   Bull. Chem.  Soc.  Japan
     53:2681-2682.

Mapleson,  W.  W. 1963.  An electric  analogue  for uptake and exchange of  inert
     gases and other agents.  J.  Appl.  Physiol.  18:197-204.
                                     3-30

-------
McNeil!, W. C. 1979. Chlorocarbons and Chlorohydrocarbons.  Trichloroethylene.
     In:  Kirk-Othmer's Encyclopedia of Chemical Technology.  Volume  5.   Third
     Edition, pp. 745-753.

Metz,  L. ,  and A.  Roedy.  1979. Chem. Ing. Technile.  21:191, 1949.  Cited in:
     Kirk-Othmer's Encyclopedia of Chemical Technology.  Volume 5.  Third Edi-
     tion.

Murray, A. J. , and J. P.  Riley. 1973a. Occurrence of some chlorinated aliphatic
     hydrocarbons in the environment.  Nature 242:37-38.

Murray, A. J., and J. P.  Riley. 1973b. The determination of chlorinated aliphatic
     hydrocarbons in air, natural waters, marine  organisms,  and  sediments.
     Anal. Chim.  Acta 65:261-270.

National Academy  of  Sciences.  1977.  Ozone and  Other Photochemical Oxidants.
     Committee on Medical  and Biologic Effects of  Environmental  Pollutants.

National Academy of Sciences. 1978. Nonfluorinated Halomethanes in the Environ-
     ment.  Panel on Low Molecular Weight-Halogenated Hydrocarbons.   Coordinat-
     ing Committee  for Scientific and Technical Assessments of Environmental
     Pollutants.

National Bureau  of  Standards.  1978.  Reaction Rate and Photochemical  Data for
     Atmospheric Chemistry-1977.  R. F. Hampson, Jr. and D. Garvin, eds.,  May.

Noweir, M., E. A. Pfitzer, and T. F. Hatch. 1973.  Decomposition of chlorinated
     hydrocarbons.  A review.  Am. Ind. Hyg.  Assoc. J. 33:454-460.

Page, N.  P. ,  and J.  L.  Arthur.  1978.  Special  Occupational Hazard Review of
     Trichloroethylene,  NIOSH, January.

Parsons, F., P. R. Wood,  and J. DeMarco. 1984.  Transformations of  tetrachloro-
     ethene and  trichloroethene  in microcosms  and groundwater.  J. Am. Water
     Works Assoc. 76:56-59.

Pasquini,  D.  A.  1978. Evaluation of  glass  sampling tubes for  industrial
     breath analysis.  Am. Ind. Hyg. Assoc. J.  39:55-62.

Pearson, C.  R. ,  and  G. McConnell.  1975.  Chlorinated C1  and  C2  hydrocarbons in
     the marine environment.   Proc. Roy. Soc. London B. 189:305-332.

Pellizzari,  E.  D.  1974.   Electron capture detection in gas chromatography.
     J. Chromatogr.   98:323-361.

Pellizzari,  E. D.  1975.  Development of  Analytical  Techniques for  Measuring
     Ambient  Atmospheric   Carcinogenic  Vapors   EPA-600/2-75-075,   November.

Pellizzari, E. D. ,  J.  E.  Bunch, R. E. Berkley and J. McRae. 1976.  Collection
     and analysis of trace organic vapor pollutants  in ambient  atmospheres.
     The  performance of  Tenax  GC cartridge sampler for  hazardous vapors.
     Anal. Lett.  9:45-63.

Pellizzari, E. D. 1977. The  Measurement  of Carcinogenic Vapors  in  Ambient At-
     mospheres.  EPA-600/7-77-055, June.


                                    3-31

-------
PeTHzzari, E. D.  1977.  Analysis  of Organic Air  Pollutants by Gas  Chromato-
     graphy and  Mass Spectrometry.   EPA 600/2-77-100.   Research Triangle
     Institute, Research Triangle Park, North Carolina.

Pellizzari, E. D.  1978.  Quantification of Chlorinated Hydrocarbons in Pre-
     viously  Collected  Air Samples.   EPA 450/3-78-112.  Research  Triangle
     Institute, Research Triangle Park, North Carolina.

Pellizzari, E. D., M. D.  Erickson,  and R. A. Zweidinger.  1979. Formulation of
     a  Preliminary Assessment of  Halogenated  Organic Compounds   in Man  and
     Environmental Media.   EPA-560/13-79-006,  U.S. Environmental  Protection
     Agency, July.

Pellizzari, E. D. 1979a. Information on the Characteristics of Ambient Organic
     Vapors in Areas  of High Chemical  Production.   Report  prepared for the
     U.S.  Environmental  Protection Agency  by  Research Triangle  Institute,
     Research Triangle Park, North Carolina.

Pellizzari, E.  D.  1979b. Organic  Screening  in Lake Charles, LA,  Using  Gas
     Chromatography  Mass  Spectrometry Computer  Techniques.   EPA  Contract
     #68-02-2714,  Research  Triangle Institute, Research  Triangle Park, North
     Carolina.

Piet, G.  J. ,  P.  Slingerland, F.  E. deGrunt, M. P.  M.  V.  d. Heuvel, and B. C.
     J.  Zoeteman.  1978. Determination of very volatile  halogenated organic
     compounds in  water by means  of  direct  headspace analysis.   Anal.  Lett.
     All:437-448.

Powell,  J.  F. 1947.  Solubility or distribution coefficient of trichloroethy-
     lene  in  water, whole  blood, and  plasma.   Brit. J.  Indust. Med. 4:233-236.

Rasmussen,  R.  A., D.  E.  Harsch,  P. H. Sweany,  J.  P. Krasnec,  and D. R.  Cronn.
     1977.  Determination of atmospheric halocarbons by a  temperature-programmed
     gas  chromatographic  freezeout  concentration  method.   J.  Air Pollut.
     Control  Assoc.  27:579-581.

Renberg,  L.  1978. Determination  of volatile halogenated hydrocarbons in water
     with  XAD-4  resin.   Anal.  Chem. 50:1836-1838.

Rogers, R. D. and J. C.  McFarlane.  1981. Sorption of carbon tetrachloride,
     ethylene dibromide, and trichloroethylene on soil and clay.  Env. Monit.
     Asses. 1:155-162.

Sato,  A., and T. Nakajima. 1979.  A structure-activity relationship of chlori-
      nated hydrocarbons.  Arch.  Environ.  Health  34:69-75.

Scherb, K. 1978.  Studies  on the  evaporation of several  low-molecular weight
      chlorinated hydrocarbons from a  flowing channel.   Muench eeitr.  abwasser-
      fisch-flussboil.  30:235-248.

 Schlossberg,   L.   1980.   Letter to  Paul  Price,  U.S.  Environmental  Protection
      Agency,  Washington, DC, June  26.

 Singh,  H. B.   1978. Discussions concerning trichloroethylene.  Atmos. Environ.
      12:1809.


                                     3-32

-------
Singh, H. B., L. Salas, D. Lillian, R. R. Arnts, and A. Appleby. 1977. Genera-
     tion of  accurate halocarbon  primary standards with  permeation  tubes.
     Environ.  Sci.  Techno!. 11:511-513.

Singh, H. B. ,  L.  J.  Salas, H.  Shigeishi,  A.  J. Smith, E.  Scribner,  and L.
     A.  Cavanagh.   1979.  Atmospheric  Distributions,  Sources,  and  Sinks of
     Selected Halocarbons, Hydrocarbons, SF6, and N20.  EPA-600/3-79-107, U.S.
     Environmental  Protection Agency, November.

Singh, H. B. ,  L.  J.  Salas, R.  Stiles,  and H.  Shigeishi.   1980.  Atmospheric
     Measurements of Selected Hazardous Organic Chemicals.  Second Year  Interim
     Report, SRI International, Menlo Park, California.

Singh, H. B. ,  L.  J.  Salas, A. J. Smith, and H. Shigeishi. 1981. Measurements
     of some potentially  hazardous organic chemicals  in  urban environments.
     Atmos.  Environ.  15:601-612.

Sjb'berg, B.  1952.  Thermal  decomposition  of chlorinated hydrocarbons.   Svensk.
     Kern. Tid.  64:63-79.

Smith, J. H., D. C. Bomberger,  Jr.,  and  D.  L.  Haynes.  1980.  Prediction of  the
     volatilization rates  of high-volatility  chemicals from  natural  water
     bodies.  Environ. Sci. Techno!.  14:1332-1337.

SRI International.  1978.  Directory of Chemical  Producers.

SRI International.  1982.  Chemical Economics Handbook, January.

SRI International. 1981.  In:   Chemical  Economics  Handbook.   Current  Indicator
     Supplemental Data 632.3001J; 632.3001V; 632.3001W, Menlo  Park, CA.

Su, C. W. and E. D. Goldberg. 1976. Environmental Concentrations and Fluxes of
     Some Hydrocarbons.   In:   Marine Pollutant Transfer.   H.  L. Windom, ed.,
     pp 353-374.

Tada, T., T. Ohta, and I.  Mizoguchi.  1976. Behavior of  chlorinated hydrocarbons
     in urban air.   Ann.  Rep. Tokyo Metr. Res.  Lab.  P.H.  27:242-246.

U.S.  Environmental  Protection  Agency.  1975.  Preliminary Study  of  Selected
     Potential Environmental Contaminants - Optical Brighteners, Methyl  Chloro-
     form,  Trichloroethylene,  Tetrachloroethylene,   Ion  Exchange  Resins.
     EPA-560/ 2-75-002, July.

U.S.  Environmental Protection Agency.  1978a.  Air Quality  Criteria  for Ozone
     and  Other  Photochemical  Oxidants.  EPA-600/8-78-004, April,   pp. 5-52.

U.S.  Environmental Protection Agency.  1978b.  In-Depth  Studies on  Health and
     Environmental Impacts  of Selected Water  Pollutants.   Contract  No. 68-01-
     4646, Duluth, MN.

U.  S. Environmental  Protection Agency. 1979. An  Assessment of the Need for
     Limitations on Trichloroethylene,  Methyl Chloroform, and Perchloroethy-
     lene.  EPA-560/ 11-79-009, July.
                                    3-33

-------
U.S.  Environmental Protection  Agency.  1980. Ambient Water  Quality Criteria:
     Trichloroethylene.  EPA-440/5-80-077, October.

U.S.   International  Trade  Commission.  1982.  Synthetic Organic  Chemicals.
     United States Production and Sales.

Wang, T. and  R.  Lenahan. 1984.  Determination of volatile  halocarbons in water
     by purged closed-loop gas chromatography.  Bull.  Environ. Contam.  Toxicol
     32:429-438.

Waters, E. M. ,  H.  B.  Gerstner, and  J. E.  Huff.  1977.  Trichloroethylene.   I.
     An overview.  J.  Toxicol. Environ. Health 2:671-707, 1977.

Zoeteman,  B.  C.  J. ,  K. Harmsen, J.  B. H.  J.  Linders,  C.  F.  H.  Morra, and W.
     Slooff.   1980.  Persistent organic pollutants  in  river  water and  ground
     water of the Netherlands.  Chemosphere 9:231-249.
                                     3-34

-------
                      4.   PHARMACOKINETICS AND METABOLISM
4.1  ABSORPTION AND DISTRIBUTION
     Trichloroethylene is a highly volatile liquid whose vapors exert pharma-
cologic effects common to general  inhalation anesthetics.   Although it is no
longer widely used as an anesthetic or analgesic agent,  the highly effective
fat-solvent properties of TCI are  the basis of its numerous industrial uses.
At work sites where this compound  is manufactured or used,  the principal  mode
of entry into the body is inhalation and lung absorption of TCI vapor.  Dermal
absorption occurs with direct liquid contact in some degreasing applications
and consumer uses (e.g., paint stripping).   Oral absorption is also an impor-
tant route of absorption.  Although sparingly soluble in water (1.0 gm/£,
25°C), TCI is present in the water supplies of many of our  cities and in  their
atmospheres, and it has been identified in measurable amounts in the food
chain.  Hence, ordinary sources in the nonindustrial environment include  air,
food, and drinking water.  Indeed, TCI has been detected in notable amounts in
the breath of healthy humans and in postmortem human tissue samples (Conkle
et al., 1975; McConnell et al., 1975; Barkley et al., 1980).

4.1.1  Dermal Absorption
     Skin absorption from TCI vapor exposure is negligible, although direct
skin contact with the liquid by immersion of the hands,  or  partial body immer-
sion, may result in significant absorption.  Uptake through the skin is related
to the type of skin exposed and the area of skin exposure as well as to the
duration of skin exposure.  Tsuruta (1978) estimated skin absorption of TCI in
mice by j_n vivo and T_n vitro techniques.   The percutaneous  absorption rate of
                                                                2
TCI in mice was reported to be between 59.8 and 92.4 umol/min/cm  or between
                       2
7.82 and 12.1 ug/min/cm .  From this study, Tsuruta concluded that direct
contact with TCI, even if the percutaneous absorption rate  of TCI is different
between human skin and mouse skin, may have the potential  to contribute impor-
tantly to the overall exposure to  TCI.  However, Stewart and Dodd (1964)  have
experimentally estimated skin absorption of TCI in volunteers by noting its
appearance and concentration in lung alveolar air after thumb immersion.   The
mean peak concentration of TCI in  expired air 30 minutes after immersion was
only 0.5 ppm,indicating that the rate of absorption through the skin is very
                                    4-1

-------
slow, even after allowing for the possibility of lipid storage and for metab-
olism to urinary metabolites (see below).   In addition, Sato and Nakajima
(1978) demonstrated that for skin exposure in man,  a large proportion of
absorbed TCI is eliminated unchanged through the lungs.  These authors also
concluded that TCI would rarely be absorbed through the human skin in toxic
amounts during normal industrial  use.
     Recently Jakobson et al.  (1982) carried out dermal absorption studies on
                                                        2
TCI with guinea pigs.  Liquid contact (skin area, 3.1 cm ) was maintained for
up to 6 hours and solvent concentration was monitored in blood during and
following dermal applications.    Blood concentration (reflecting absorption
rate) increased rapidly, peaking at 0.5 hour (0.8 ug/ml blood), and then
decreased despite continuing exposure (0.46 ug/ml blood after 6 hours).   This
pattern was characteristic of halogenated hydrocarbon solvents of low water
solubility.  Of 12 solvents tested, TCI produced the second lowest blood
concentrations from dermal absorption.

4.1.2  Oral Absorption
     TCI is an uncharged, nonpolar, and highly lipophilic compound (Table 4-1)
and consequently can be expected to readily cross the gastrointestinal mucosal
barrier by passive diffusion.   In man, absorption through the gastrointestinal
mucosa is extensive, as documented by the numerous cases of poisoning by oral
ingestion reported over the years (Defalque, 1961; Vyskocil and Polak, 1963;
Wiecko, 1966; Vignoli et al., 1970; Migdal et al., 1971).  In animals, the
oral LD5Q in mice, rats and dogs are similar, being 3.2, 4.9 and 2.8 g/kg,
respectively (Klaasen and Plaa, 1966, 1967; Smyth et al., 1969).
     Daniel (1963) dosed rats by stomach-tube with   Cl-labeled TCI liquid (40
to 60 mg/kg b.w.) and recovered 90 to 95 percent of the label in expired air
and urine, suggesting virtually complete absorption by the route.  Similar
results have been reported recently by Dekant et al. (1984) and Prout et al.
                                                                      14
(1984).  Dekant et al. dosed by gavage both rats and mice with 200 mg   C-TCI/kg
in corn oil vehicle  and recovered 93 to 98 percent of the radiolabel in expired
                                                      14
air and urine (Table 4-1).  Prout et al. administered   C-TCI in corn oil
vehicle as single intragastric doses of 10, 500, 100, and 2000 mg/kg to both
rats and mice and recovered 91 to 98 percent of the doses in expired air and
urine  (Table 4-2).   Peak blood levels occurred at about 1 hour in mice and 3
hours  in rats,  indicating rapid absorption from the GI tract.
                                    4-2

-------
TABLE 4-1.  RECOVERY OF RADIOACTIVITY FOR 72 HOURS AFTER A SINGLE ORAL DOSE OF
            14C-TCI (200 mg/kg) TO FEMALE WISTAR RATS AND NMRI MICE
As % Recovered Radioactivity*

Exhaled air
Unchanged, 14C-TCI
14C02
Urine
Cage rinse
Feces
Carcass, tissues

Time
(hr)
0-72
0-72
0-12
12 - 24
24 - 48
48 - 72
0-72
0-72
Rats
52.0
1.9
53.9
22.0
17.0
2.0
0.2
0.2
41.4
1.8
2.9
Mice
11.0
6.0
17.0
30.0
39.0
6.0
1.0
0.1
76.3
4.9
2.0
*N = 2 rats and 3 mice; total recovered radioactivity was 93 to 98% of the
 administered radioactivity given the rats and mice.  The average weight
 of the rats was 240 gm and the mice 25.5 gm.
Source:  DeKant et al.  (1984).

     Withey et al. (1983) demonstrated in rats that gastrointestinal absorption
of TCI is considerably more rapid in water solution than when given in corn
oil as a vehicle.  Following an 18 mg/kg intragastric dose in 5 ml water or
corn oil to fasting rats (400 g), the peak blood concentration (5.6 minutes
for aqueous solution) averaged 15 times higher for water than for corn oil
solution (14.7 versus 1.0 ug/ml); however, the peak blood concentration was
reached faster for water solution than for oil solution, which exhibited a
second delayed peak 80 minutes post absorption.  Comparison of the area under
the blood concentration curve (AUC) during the absorptive phase reflects the
rate and extent of absorption.  The ratio of the AUCs for 5 hours after dosing
was 218, water:corn oil.   These results illustrate the effect of solution
                                    4-3

-------
           TABLE 4-2.   DISPOSITION OF SINGLE  DOSES  OF 14C-TCI ADMINISTERED BY GAVAGE TO
                                 OSBORNE-MENDEL  RATS  AND B6C3F1 MICE
% Dose Excreted **
Dose, Unchanged
mg/kg 14C-TCI
Osborne-Mendel
10
500
1000
f 2000
B6C3F1 Mice*
10
500
1000
2000
Rats*
1.3
42.7
56.2
77.8
3.9
6.0
17.5
13.6
Metabolized
14C02
10.2
3.7
2.9
1.4
11.7
8.9
7.6
7.8
Urine
69.4
43.6
31.1
14.0
59.4
63.7
46.4
47.5
Feces
8.3
4.3
3.1
3.4
17.2
14.5
15.0
19.2
Carcass
8.4
3.8
2.4
1.3
4.8
4.4
8.6
3.8
Total
96.3
55.4
39.5
20.1
93.1
91.5
77.6
78.3
Recovery
97.6
98.1
95.7
97.9
97.0
97.5
95.1
91.9
 *Means of groups of 4 animals; rat weights 180 to 200 gm and mouse 25 to 32 gm.

**During 72-hr period after dosing.

Source:  Prout et al. (1984).

-------
medium on TCI intestinal absorption; the much slower absorption of TCI dis-
solved in corn oil is probably due to the high lipid solubility (and poor
aqueous solubility) of this compound which remains as a depot in the oil.
Hence, absorption is limited to the rate of absorption of the corn oil itself,
or the corn oil effectively acts as a slow-release dosage form.

4.1.3  Pulmonary Absorption
4.1.3.1 Man.  Inhaled TCI rapidly equilibrates across the lung alveolar endo-
thelium.   The blood/gas partition coefficient of 9.92 at 37°C is comparable to
the anesthetic gases—chloroform, diethylether, and methoxyfluorane--but its
lipid solubility is considerably higher (olive oil/gas coefficient, 960 at
37°C) (Powell, 1947; Eger and Larson, 1964).   Thus, pulmonary uptake of TCI is
rapid and extensive with considerable distribution into lipid of body tissues
(Table 4-3).

  TABLE 4-3.  PULMONARY UPTAKE OF TCI FOR 25  VOLUNTEERS EXPOSED  TO TCI FOR FOUR
     CONSECUTIVE 30-MIN PERIODS AT REST AND DURING EXERCISE (50  TO 150 WATTS)

   Exposure
 concentration,                            Amount inhaled,     Amount taken up,*
mg/m3
540
1080
540
1080
Conditions
rest
rest
Exercise (50w)
Exercise (50w)
mg
149 ± 6
313 ± 15
395 ± 6
803 ± 12
mg
79 ± 4
156 ± 9
160 ± 5
308 ± 16
 540 and 1080 mg/m3 = 100 and 200 ppm,  respectively.
Calculated from difference in inspired air and expired air x ventilation volume
 x time.   Retention value averaged = 45 percent.
Source:   Astrand and Ovrum (1976).
                                    4-5

-------
     Pulmonary  uptake is proportional to  the magnitude of the  gas  partial
pressure difference  between alveoli and pulmonary venous blood.  This differ-
ence diminishes when all body tissues achieve equilibrium with the alveolar
(arterial) partial pressure.  Initially,  uptake is rapid, but  about 8 hours of
exposure is  required for complete tissue  equilibrium to be achieved (Fernandez
et al., 1977; Sato et al.; 1977; Monster  et al.,  1976, 1979; Muller et al.,
1974).  Figure  4-1 shows that equilibrium is achieved most rapidly (4 to 6
hours) with  the rich blood-perfused tissues.  The poorly perfused  fat tissues
(FG) require longer  than 8 hours for equilibrium.
                                                            FG
                                                        — MG
                                                            VRG               —
                                                        	ALV. AIR AND ART. BLOOD
                246

                   EXPOSURE, hours
246
 POST EXPOSURE, hours
         Figure 4-1. Predicted partial pressure of TCI in alveolar air and tissue groups during and
         after an 8 hour exposure of 100 ppm.
         Source: Fernandez et al. (1977).
                                        4-6

-------
     At equilibrium with a given concentration of TCI in inspired air, the
difference between the inspired air concentration of TCI and alveolar concen-
tration of TCI provides a measure of gas uptake.   With appropriate accounting
for ventilation (about 4 to 8 L/min at rest) and duration of exposure, the
amount taken up by the body, or retention (dose;  body burden), can be calcu-
lated.  At alveolar and tissue equilibrium, the retention reflects disposition
within the body by tissue storage, metabolism, or excretion by routes other
than the lungs.  Retention of 36 to 75 percent has been reported for TCI by
various investigators (Soucek and Vlachova, 1960; Bartonicek, 1962; Nomiyama
and Nomiyama, 1971, 1974a,b), with higher values (70 to 75 percent) reported
from more recent studies (Fernandez et al., 1975; Monster et al., 1976, 1979)
(Tables 4-3 and 4-4).  The retention value is independent of the inspired TCI
concentration.  In contrast, the amount (dose) of TCI absorbed into the body
is directly proportional to the TCI concentration in the inspired air (Fernandez
et al., 1975; Monster et al., 1976, 1979; Muller et al., 1974).   The body
burden of TCI also increases with duration of exposure, repeated daily exposures
and exercise (increased ventilation) at a given inhaled air concentration
(Astrand and Ovrum, 1976; Monster et al., 1976, 1979).   These relationships
between inspired air concentration, pulmonary uptake and body burden are
illustrated by the data of Astrand and Ovrum and of Monster et al . in Tables
4-3 and 4-4.  Astrand and Gamberale (1978) exposed volunteers to 100 and 200
ppm (ventilation volume 9.3 L/min) for 70 minutes and observed a linear rela-
tionship between pulmonary uptake (Q) and alveolar concentration (A) divided
by inspired air concentration (I)

                               Q = 0.72       +  74-91-
However, as about eight or more hours are required for complete body equilibra
tion, this relationship only approximates steady-state inhalation conditions.
     The blood concentration of TCI during inhalation and in the elimination
phase after exposure closely parallels alveolar gas concentration (Figure 4-1
and Figure 4-2) (Muller et al., 1974; Monster et al., 1976, 1979; Stewart
et al., 1974; Sato et al., 1977; Fernandez et al., 1977).   The kinetics of
tissue uptake, accumulation, and equilibration of TCI is a function of blood
concentration, the blood flow/tissue mass, and the relative solubilities of
TCI in tissues (Table 4-5).   Therefore,  TCI most rapidly attains equilibrium
                                    4-7

-------
     TABLE 4-4.  PULMONARY UPTAKE  OF TCI FOR  4 SUBJECTS  AT REST EXPOSED TO
     70  AND 140 PPM  TCI  FOR 4-HOUR PERIODS AND 3-HOUR PERIODS, PLUS TWO 30-
                           MIN PERIODS OF WORK  (100 WATTS)
Exposure concentration
and conditions

70 ppm,
70 ppm,
140 ppm,
140 ppm,

rest
rest + work
rest
rest + work
A
320
500
740
1050
Dose, mg
Subject
B
430
450
790
1100
C
330
470
710
970
D
470
660
790
1100
Mean
mg

390
520
755
1055
Retention  value average = 38 percent.

Source:  Monster (1976).
               12
               10
             E  8

             u

             I  6
            oc
            <
                                 I      I     I      I
                                TRICHLOROETHYLENE
                                                       I
                     100
300         500         700

   AVERAGE CONC. mg/m3
900
                Figure 4-2. The relationship between the concentration of TCI in
                arterial blood and alveolar air at the end of inhalation exposures (30
                min periods, for 15 male subjects). Symbols stand for one exposure
                period for a subject at rest, and at various levels of exercise. The
                regression line is y = —0.489 + 0.014 X.
                Source: Astrand and Ovrum (1976).
                                         4-8

-------
    TABLE 4-5.   PARTITION COEFFICIENTS OF TCI  FOR VARIOUS BODY TISSUES AND
                           COMPONENTS OF RAT AND MAN
Rat
Blood/air
Lung/blood
Heart/blood
Kidney/blood
Liver/blood
Muscle/blood
Brain/blood
Testes/blood
Spleen/blood
Fat/blood
Man
Blood/air
Fat/air
Lecithin/air
Cholesterol/air
Cholesterol oleate/air
Triolein/air
Mean values
25.82
1.03
1.10
1.55
1.69
0.63
1.29
0.71
1.15
25.59

9.92
674.4
387.9
52.15
261.93
848.24
Conversion factors:
        1 ppm vapor in air = 5.38 mg/m3 at 25°C, 760 torr.
Source:   Adapted from Sato et al. (1977).

by passive diffusion into the vessel-rich group (VRG) of tissues (brain,
heart, kidneys, liver, and endocrine and digestive systems), more slowly with
the lean mass (MG) (muscle and skin), and last with adipose tissue, i.e., fat
groups (FG).  As determined from elimination kinetics following exposure, TCI
distributes from blood into these three major compartments at approximate rate
constants of 17 hr'1 (t, , 2.4 min) for VRG, 1.7 hr'1 (t^, 25 min) for MG, and
0.2 hr"1 (tj,, 3.4 hr) for FG (Fernandez et al., 1975, 1977; Sato et al.,
1977).  Though the lean muscle mass is 50 percent of the body volume and
adipose tissue is 20 percent, saturation and desaturation proceed more rapidly
from the MG compartment (tt , 25 min) than from the FG compartment (tj , 3.4 hr)
because of the considerably greater solubility of TCI in lipid (Table 4-1).
Interperson variations in TCI uptake (dose) are therefore influenced primarily
by lean body mass and secondarily by adipose tissue mass (Monster et al.,
1976, 1979).

                                    4-9

-------
4.1.3.2 Animals.   Stott et al.  (1982) have determined the pulmonary uptake of
        	                                                 o
TCI in mice and rats exposed to 10 and 600 ppm (54 and 3228 mg/m ).   The
                                                           14
animals were exposed in a 30-L glass inhalation chamber to   OKI, the concen-
tration of which was maintained within 10 percent of target.   At termination
of exposure (6 hours), the cumulative pulmonary uptake was estimated by deter-
mining the radioactivity of carcass, and of TCI and metabolites in expired
air, urine and feces.  The data for mice and rats are given in Tables 4-6 and
4-7, respectively.  For mice,  the body burdens for 6-hr exposure to 10 and
600 ppm (54 and 3228 mg/m3) were 78.5 and 3138 umol equivalents/kg (10.31 and
412.3 mg/kg; or 0.24 and 9.48 mg/mouse), respectively.  For rats, amounts were
35.8 and 1075 umol equivalents/kg (4.7 and 141.3 mg/kg; or 1.03 and 31.09
mg/rat), respectively.  For these species the 6-hr pulmonary uptake was not
directly proportional to inspired air concentration level, as observed in man.
Furthermore, total uptake of TCI, rat versus mouse, is more closely related to
                                           p/o
their relative body surface areas—(220/23) '  = 4.5-fold--than it is to their
relative body weights, 9.6-fold.  Presumably, this difference is primarily
related to the overriding influence of metabolism of TCI on uptake rather than
to differences in other factors such as ventilation, minute volume, and body
weight.

4.1.4  Tissue Distribution and Concentrations
     TCI distributes  to all body  tissues and appears  in  sweat and saliva
(Bartonicek, 1962).   It crosses the blood-brain barrier  (anesthetic effect)
and the placental barrier.  Helliwell and Mutton  (1950),  investigating the
distribution in pregnant  sheep and  goats, found that  TCI  appeared  in  the  fetal
circulation within  2  minutes of exposure; at equilibrium,  the fetal/maternal
blood  concentration  ratio was  unity.  The partition coefficients  for  various
tissues of rat and  man are  given  in  Table 4-5  (from Sato et  al.,  1977).
4.1.4.1  Man.  There  is  little  direct analytical  information  of tissue  levels
after  various  exposure concentrations of  TCI.  Tissue concentrations  are
expected  to be proportional to  exposure  concentrations and to duration  of
exposure.   Fernandez et  al.  (1977),  from  their mathematical  model  of  TCI
 uptake, distribution, and excretion,  based  on  experimental data,  have pre-
dicted the partial  pressure of TCI  in the three  major compartmental  tissue
 groups, vessel-rich group (VRG),  muscle group  (MG),  and adipose tissue (FG),
 during and after an 8-hr inhalation exposure of  100 ppm (538 mg/m ).   Figure 4-2
                                     4-10

-------
   TABLE 4-6.  RECOVERY OF RADIOACTIVITY FOR 50-HOUR POSTINHALATION EXPOSURE OF
                MALE B6C3F1 MICE TO 10 OR 600 PPM 14C-TCI FOR 6 HOURS1


10 ppm
umol-eq TCI/
kg body weight
Expired
1,1,2-TCI
C02
Urine
Feces
Cage wash3
Skin
Liver
Ki dney
Carcass
Total body burden
Total metabolized

0.617
7.38
58.1
2.92
1.10
2.33
2.40
0.310
1.50
78.5
77.9

(0.182)
(0.383)
(18.8)
(0.542)
(0.446)
(1.13)
(0.401)
(0.310)
(0.316)
(17.9)
(18.1)

%2

0.79
9.4
74.0
3.7
1.4
3.0
3.1
0.39
1.9

99.2

600 ppm
umol-eq TCI/
kg body weight

76.2
299
2278
115
85.5
56.3
72.2
11.1
146
3138
3062

(27.3)
(37.5)
(505.0)
(24.6)
(65.5)
(11.9)
(10.4)
(3.18)
(57.9)
(616.0)
(592.0)

%2

2.4
9.5
72.6
3.7
2.7
1.8
2.3
0.35
4.6

97.6
1Average of four animals (±SD)
Percentage of recovered radioactivity.
3Primarily due to urine.
Source:   Stott et al., 1982.
  TABLE 4-7.   RECOVERY OF RADIOACTIVITY FOR 50-HOUR POSTINHALATION EXPOSURE OF
          MALE OSBORNE-MENDEL RATS TO 10 OR 600 PPM 14C~TCI FOR 6 HOURS1

Expired
1,1,2-TCI
C02
Urine
Feces
Cage wash3
Skin
Liver
Kidney
Carcass
Total body burden
Total metabolized
10 ppm
umol-eq TCI/
kg body weight

0.778 (0.103)
1.71 (0.295)
22.6 (5.96)
2.52 (0.253)
0.357 (0.204)
3.78 (1.21)
1.01 (0.085)
0.131 (0.009)
2.82 (0.898)
35.8 (5.64)
35.0 (5.64)

%2

2.1
4.8
63.1
7.0
1.0
10.6
2.8
0.37
7.9

97.8
600 ppm
umol-eq TCI/
kg body weight

227 (6.40)
31.2 (2.10)
594 (79.8)
38.3 (3.47)
15.1 (10.7)
76.2 (30.3)
16.3 (1.13)
1.72 (0.083)
75.1 (27.7)
1075 (87.2)
847 (81.6)

%2

21.1
2.9
55.3
3.6
1.4
7.1
1.6
0.16
7.0

78.9
1Average of four animals (±SD)
Percentage of recovered radioactivity.
Primarily due to urine.
Source:   Stott et al. ,  1982.
                                    4-11

-------
shows that VRG and MG are nearly always in equilibrium with arterial  blood
concentration, but the concentration of TCI in FG increases only very slowly
and continues to increase beyond the 8-hr exposure.   During desaturation,  TCI
in FG exhibits a long residence (about 40 hours) for disappearance with detec-
table FG concentrations remaining even after 70 hours.   Because of the long
residence of TCI in FG, repeated daily exposure (6 hr/day;  100 ppm) results in
an accumulated concentration, as TCI from new exposures adds to residual
concentration from previous exposures, until steady-state is reached.  This
accumulation of TCI in FG is shown in Figure 4-3.  Adipose tissue levels of
TCI reach a steady-state "plateau" (dependent on the inhalation concentration
and daily period of exposure) and thereafter remain relatively constant.  For
a 6-hr/day, 100-ppm (538 mg/m ) exposure, Fernandez et al.  (1977) predict that
FG peak plateau concentration occurs 5 to 7 days after initial daily exposure.
4.1.4.2  Rodents.  Following intravenous administration of TCI to Wistar rats,
Withey and Collins (1980) found a whole-body half-time of disappearance from
the rat of 2 to 6 minutes (dependent on dose) but a disappearance half-time of
215 minutes for adipose tissue, indicative of the propensity of TCI to remain
in adipose tissue as compared to other tissues.  Pfaffenberger et al. (1980),
after dosing rats by gavage for 31 days (1 and 10 mg/day),  found that blood
serum levels of TCI were not detectable (analyzed 2 hours after last dosing);
but adipose tissue levels averaged 0.28 and 20.0 ug TCI/g, respectively.
Three days after last dosing, 1 ug TCI/g fat still remained in adipose tissues.
     Savolainen et al. (1977) analyzed tissue levels of TCI after exposure of
                                                     3
rats 6 hours per day for 5 days to 200 ppm (1076 mg/m ) TCI.  Table 4-8 gives
the values found.  Seventeen hours after exposure on day four, blood and
adipose tissue levels were 0.35 and 0.23 mol/g, respectively, indicating long
storage in adipose tissue, though other tissues contained virtually  no TCI.
With exposure on day five, tissue levels in brain, lungs, liver, fat, and
blood reached a steady-state within 2 to 3 hours.  Partition coefficients were
adipose tissue/blood, 10.4; brain/blood, 1.3; lungs/blood, 0.8; and  liver/
blood, 0.5.  The lower value,for  liver possibly  reflects metabolism  occurring
in this organ, while the high value for adipose  tissue reflects high solubility
of TCI in  lipids.
                                    4-12

-------
oo
                                                                    REPEATED EXPOSURE 100 ppm
                                                                         (5 days, 5 hr. day)
                         Mo
Th         FT

EXPOSURE, days
                     Figure 4-3. Predicted partial pressure of TCI in fatty tissue for a repeated exposure to
                     100 ppm, 5 days, 6 hours per day.

                     Source: Fernandez et al. (1977).

-------
  TABLE 4-8.   ORGAN CONTENTS OF TCI  AFTER DAILY INHALATION EXPOSURE OF 200 PPM
                               FOR 6 HOURS PER DAY
Time (hr of
exposure) on
fifth day

0 [17 hr after
day 4- expo-
sure]
2

3

4

6

Cerebellum

0


11.7
± 4.2
8.8
± 2.1
7.6
± 0.5
9.5
± 2.5
Cerebrum

0


9.9
± 2.7
7.3
± 2.2
7.2
±1.7
7.4
± 2.1
Lungs

Liver
nmol/g
0.08 0.04


4.9
± 0.3
5.5
± 1.4
5.8
± 1.1
5.6
± 0.5


3.6

5.5
± 1.7
2.5
± 1.4
2.4
± 0.2
Peri renal
fat

0.23
± 0.09

65.9
±1.2
69.3
± 3.3
69.5
± 6.3
75.4
± 14.9
Blood

0.35
± 0.1

7.5
±1.6
6.6
± 0.2
6.0
± 0.9
6.8
±1.2
Values are mean of 2 determinations ± range.
Source:   Savolainen et al.  (1977).

4.2  EXCRETION
     The total elimination of absorbed TCI involves two major processes,
pulmonary excretion of unchanged TCI and hepatic biotransformation to urinary
metabolites.   The details of hepatic biotransformation of TCI are reviewed
below.  In man, the principal metabolites are trichloroethanol (TCE), TCE-
glucuronide,  and trichloroacetic acid (TCA).   These metabolites are excreted
in urine, and in less significant quantities by other excretion routes, such
as the bile (Bartonicek, 1962).

4.2.1 Pulmonary Elimination in Man—Pulmonary elimination of unchanged TCI
after exposure occurs nearly as an inverted reproduction of its uptake.
Alveolar concentration parallels blood concentration; both follow exponential
concentration decay curves showing three major components with approximate
half-times of 2 to 3 minutes, 30 minutes, and 3.5 to 5 hours (Figures 4-1 and
4-4), usually represented by VRG, MG, and FG compartments, respectively
                                    4-14

-------
  0.5




  0.1

 0.05




 0.01

0.006




0.001
3     5

   1

  0.5




  0.1

 0.05




 0.01
T     I      I     \     I     T

                 EXPIRED AIR
                                                I     I     I     I
                                0.0126»-°-1795t      _
                                             I      I     I     I
        I     I     I
                                                   8     9    10
                 I     T     I     I    I     I    I
                                                        I	I
                            4     6B

                                HOURS
                                                 10
                                                                                    *Aax
                        Figure 4-4. Elimination curves of TCI and kinetic model for transfer of TCI in the body. A a x =
                        pulmonary excretion (ventilation  x air/blood partition coefficient x blood concentration);  bx =
                        metabolic clearance (urine) x blood concentration; V-), N/2, V"3 volumes of VRG, MG and FG com-
                        partments respectively.

                        Source: Sato et al. (1977).

-------
(Fernandez et al.,  1977; Sato et al., 1977; Muller et al.,  1974; Nomiyama and
Nomiyama, 1971; Monster et al., 1979).
     Sato et al.  (1977) developed a physiological kinetic model for TCI excre-
tion in man, based  on experimental  findings from controlled exposures of
volunteers.   Four male medical students (avg.  b.w.:  61.6 kg) inhaled 100 ppm
         3
(538 mg/m )  for 4 hours.  After cessation of exposure, the blood and exhaled
air concentrations  of TCI and the concentration of urinary metabolites were
determined during the time-course of 10-hr postexposure.  Figure 4-4 shows the
results obtained.   The time-course of concentration decay (C) TCI in blood and
exhaled air  is the  sum of three exponentials:
                            -23.6+         -1.484-          .0.184-
               C .   = 0.213E       + 0.029E       + 0.0126E
                air
                              -16.7t         -1.714-         -0.204-
               Cblood = °-115E       + °-449E       + 0.255E

These reflect first-order excretion from three compartments (VRG, MG, and FG,
respectively) of the model (Figure 4-4).   As indicated in the model, C -
                                                                      al T*
represents pulmonary clearance only, but C. ,   . represents whole-body clearance
by the pulmonary system and metabolism.  The kinetic parameters were determined
as follows:   V (volume of FRG, MG,  and FG compartments), 8, 31, and 10/£,
respectively; metabolic clearance,  104 £/hr; partition coefficients (blood/air,
VRG/ blood,  MG/blood, FG/blood), 10, 1.5, 1.0, 67; ta  of FG, 3.4 hours.
Similar tri-exponential curves for decay of blood and exhaled air concentra-
tions of TCI have been reported by other investigators (Stewart et al., 1970,
1974; Nomiyama and  Nomiyama, 1974; Fernandez et al., 1975,  1977; Monster,
1976).
     The long half-time of elimination from adipose tissue (3.5 to 5 hours)
explains why, after a single short exposure (e.g., 4 to 6 hours), exhaled air
contains a notable  concentration of TCI 18 hours later, and also why body
burden of TCI occurs with repeated, fluctuating daily exposures until steady-
state is reached (Section 4.1.4).  Several investigators (Fernandez et al.,
1977; Monster et al., 1979) have shown that, for volunteers exposed to 70 to
                        3
100 ppm (377 to 538 mg/m ) 4 hours daily for 5 days, the concentration of TCI
in blood and exhaled air was two to three times higher 18 hours after the last
exposure than at the comparable time after a single exposure.  TCI accumulation
was highest in those volunteers with a proportionally larger ratio of adipose
to lean muscle mass.
                                    4-16

-------
4.2.2 Urinary Metabolite Excretion in Man
     The urinary metabolites of TCI (total amount, excretion ratios, and
excretion time-course) have always been of considerable interest over the
years as quantitative indices of exposure and body burden.   Recent investiga-
tions have determined that the half-time for renal elimination of TCE and of
TCE-glucuronide approximates 10 hours (range 7 to 14 hours) after TCI exposure
(Nomiyama and Nomiyama, 1971; Muller et al., 1974; Sato et al., 1977; Monster
et al., 1976, 1979).   This value is in agreement with that observed after
ingestion of TCE itself (Breimer et al., 1974; Muller et al., 1974).  As might
be predicted from this relatively short half-time of renal  elimination, blood
concentrations of TCE and TCE-glucuronide after repeated daily TCI exposures
reach a "plateau concentration" (with peaks for each daily exposure) that is
dependent on the inhaled air concentration of TCI (Monster et al., 1979;
Muller et al., 1974;  Fernandez et al., 1977).   Consequently, the same daily
amount of TCE and TCE-glucuronide is excreted into the urine during repeated
daily exposures to the same air concentration of TCI.
     In contrast to the short half-time of TCE elimination, the half-time of
renal elimination of TCA produced during TCI exposure approximates 52 hours
(range 35 to 70 hours) (Soucek and Vlachova, 1960; Nomiyama and Nomiyama,
1971; Muller et al.,  1974; Sato et al., 1977;  Monster et al., 1976).  Similar
values have been observed when TCA is directly administered to men (82 hours,
Paykoc and Powell, 1945; 51 hours, Muller et al., 1974).   Distribution of TCA
in the body is restricted to the extracellular water space (Paykoc and Powell,
1945); it is very tightly and extensively bound to plasma protein and excluded
from erythrocytes, so that plasma concentration is twice as high as whole blood
(Soucek and Vlachova, 1960; Monster et al., 1979).  As a consequence of the
long half-time of TCA, and of the constant metabolic formation of TCA from
TCE, blood concentrations of TCA smoothly rise during a single TCI exposure,
reaching a maximum 24 to 48 hours postexposure before declining exponentially
for 10 to 15 days.  The blood concentration of TCA during the rising phase and
at maximum is proportional to the TCI inspired air concentration and the
duration of the exposure, i.e., to the retained TCI dose (Soucek and Vlachova,
1960; Monster et al., 1976; Sato et al., 1977; Fernandez et al., 1977).
During repeated daily exposures, TCA accumulates in the blood exponentially,
and declines only 24 to 48 hours after the last exposure (Muller et al., 1974;
Fernandez et al., 1977; Monster et al., 1979).  Consequently the daily amount
                                    4-17

-------
of TCA excreted into the urine increases during repeated exposure and then
decreases slowly over a period of 15 to 20 days.   Several  investigators have
documented an apparent increase in the half-time of renal  elimination of TCA
(to 100+ hours) with repeated daily exposures to TCI.   This phenomenon is
explained by continuous formation of TCA from accumulated tissue stores of TCI
and TCE, and by the very tight binding of TCA to plasma protein.
     The renal elimination kinetics of TCE and TCA readily explain the urinary
excretion ratio time-course of these metabolites.   The ratio of daily urinary
metabolites, TCE/TCA (mg/day), is highest (2:1) 24 hours after a single TCI
exposure or after repeated daily exposures (Soucek and Vlachova, 1960; Muller
et al., 1972, 1974; Fernandez et al., 1975; Nomiyama and Nomiyama, 1971, 1977;
Sato et al., 1977; Monster et al., 1976, 1979).  Thereafter, the relationship
of TCE/TCA progressively decreases from day to day and becomes less than unity
at about the sixth day following exposure.  Several additional factors are
known to influence the excretion ratio, TCE/TCA.   The ratio is reported to be
markedly greater in males than females for the first 24 hours after exposure
(Nomiyama and Nomiyama, 1971, 1977), and TCA urinary excretion is reported to
vary diurnally (Soucek and Vlachova, 1960; Monster et al., 1979).  There is no
evidence in man that the elimination of TCI by metabolism is saturable, at
least up to 300 ppm (1584 mg/m ) inhalation concentrations, but saturation may
occur at higher levels (Section 4.4.2).

4.2.3 Excretion Kinetics in the Rodent
     Withey and Collins (1980) determined the  kinetics of distribution and
elimination of TCI from blood of Wistar rats after intravenous  (i.v.) admini-
stration of 3, 6, 9, 12, or 15 mg/kg of TCI given  in 1 ml water interjugularly.
For the three  lower doses, the blood decay curves  exhibited two components of
exponential disappearance and best fitted a first-order two-compartment model,
with average  kinetic parameters of K  , 0.18 min   ; V., 158 ml;  k,„ and  k~,,
0.059 and 0.39, respectively.  For the highest dose (15 mg/kg), these  investi-
gators  found  that their data best fitted  a first-order three-compartment
model.  They  suggest that the shift  from  two-  to  three-compartment kinetics
may have occurred either because of  a dose-related alteration  in  the  kinetic
mechanisms  of uptake,  distribution,  metabolism, and elimination,  or that
three-compartment  kinetics were  followed  at all dose  levels,  but  with  lower
doses,  the  third  exponential component,  representing  the  third compartment,
                                    4-18

-------
was obscured by the limit of analytical sensitivity.  The kinetic values for
the three-compartment model (using a dose of 15 mg/kg) were k , 0.174 min  ;
Vd, 111 ml; k12, k21, k13:  0.183, 0.164, and 0.051, respectively.  The half-
time, tj , for disappearance of TCI from peri renal fat was 217 minutes.  Withey
and Collins (1980) state that they found no evidence in their kinetic analysis
of the disposition of i.v. bolus injections of TCI into the rat of nonlinear
or dose-dependent Michaelis-Menten kinetics, although the first-order kinetic
parameters obtained were clearly dose-dependent.   These workers suggested that
a dose of 15 mg/kg TCI to the rat is below hepatic metabolism saturation.
However,  other investigators have found clear evidence of dose-dependent
Michaelis-Menton metabolism kinetics in rats at higher doses (Section 4.4.2.2).
     Prout et al.  (1984) also have made observations on the overall rate of
elimination of TCI from the blood of rats and of mice.  After single intra-
gastric doses of 1000 mg/kg in corn oil vehicle to rats and to mice, the blood
levels of TCI and principal metabolites were monitored by GC analysis.  Figure
4-5 shows the results of these experiments.  Peak blood TCI occurred at 1 hour
in the mouse, but rats exhibited a minor peak at 1.5 hours and a major peak at
3 hours.   This absorption behavior in the rat has been previously reported by
Withey et al. (1983) and is characteristic of corn oil vehicle but not of
water solutions of TCI (Section 4.1.2).  Following absorption and distribution
(post blood peak), blood TCI disappeared (pulmonary elimination plus metabolism)
in an exponential  first-order manner with a half-life of about 1.75 hours for
the mouse and 2.25 hours for the rat, i.e., with nearly the same rate constant
in the two species.   The primary hepatic metabolite, chloral (see Figure 4-5),
appears in mouse blood rapidly with a major peak at 2 hours, but in the rat
the peak was delayed to 10 hours; furthermore, the blood levels of chloral
were 2 to 4-fold higher in the mouse than rat.  These investigators suggest
that these results indicate the metabolism of TCI in the mouse in considerably
more rapid than that of the rat.  A further difference between the mouse and
rat was an apparent difference in the conversion of chloral to TCE and TCA.
Prout et al. (1984) stress that the more rapid conversion of chloral to TCA in
the mouse leads to an overall sustained elevation of blood levels of TCA in
the mouse (~7-fold; Figure 4-5, Panel D) than in the rat, although the relative
proportion of TCA to other metabolites (7 to 15 percent TCA) in urine, bile,
and feces was not greater than that of the rat (Table 4-9); indeed, the metab-
olite proportion appears independent of dose, species, or animal strain (Green
                                    4-19

-------
Figure 4-5. Whole blood levels (jug/ml) of TCI and metabolites after single gavage dose of TCI
(1000 mg/kg) in corn oil to male Osborne-Mendei rats (triangles) and B6C3F1 mice (circles).

Source: Prout et al. (1984).
                                      4-20

-------
           TABLE 4-9.   PROPORTION OF TCI METABOLITES (TCA AND TCE)
        IN URINE OF RATS AND MICE AFTER SINGLE ORAL DOSES OF 14C-TCI
Dose
mg/kg
10
500
1000
2000


TCA
5.8
6.3
8.3
7.0
% of urine
Rat
TCE
91.5
90.1
88.4
91.9
radioactivity
Mouse
TCA
12.2
7.2
7.0
7.2


TCE
84.5
91.1
88.5
90.1
 Urine collected for 24 hours from Osborne-Mendel  rats and B6C3F1 mice.
 Similar results from Alderly Park (Wistar derived) rats and Swiss-Webster
 mice.
Source:   Green and Prout (1984).

and Prout, 1984).   However, as seen in Figure 4-5, Panel D, the area under the
TCA blood curve (AUC), as a measure of TCA production in the mouse, is 10-fold
greater than the AUC for the rat, with the implication of a considerably
greater production of TCA per unit dose in the mouse than rat.
     Prout et al.  (1984) also evaluated the excretion of TCA and of CO,, (as
measures of TCI),  as a function of dose in both the rat and the mouse, respec-
                                               14
tively.   Animals were dosed with a solution of   C-TCI in corn oil given as a
single intragastric administration at levels of 10, 500, 1000,  and 2000 mg/kg.
14                                     14                             14
  C-radioactivity excreted in urine as   C-TCA, and in expired air as   CO^
(from TCA metabolism as one source) from 0 to 24 hours, was measured.  Figure
                                   14          14
4-6 shows the results.  The sum of   C-TCA and   COp radioactivity excreted by
mice (B6C3F1) over the dose range 10 to 2000 mg/kg was linear with dose.  In
contrast, the excretion of these metabolites by rats (Osborne-Mendel) was non-
linear and rate-limited, indicating in the rat (but not mouse) dose-dependent
kinetics for TCI metabolism to TCA.
     Green and Prout (1984) have investigated the kinetics of urinary excretion
of TCI metabolites in mice (B6C3F1) during chronic administration of TCI.
Single intragastric doses of TCI in corn oil (1000 mg/kg) were administered
                                         14
daily for 180 days.  On days 10 and 180,   C-TCI was incorporated into the
                                        14             14
dose.   The 24-hr excretion of unchanged   C-TCI and of   CO,, in the exhaled
                                    4-21

-------
  250r
  200
 i
O
u
Q
  150
CO

Q
  100
cc
u
X
CO
O
Q
   50
                O MOUS
-------
                     14                                      14
breath, and of total   C-radioactivity in urine expressed as   C-TCA and
14
  C-TCE, were measured.  Control animals (untreated) were given single doses
   14
of   C-TCI on days 10 and 180.   The results are shown in Table 4-10.   Chronic
dosing of TCI in mice for up to six months had no overall effect on the extent
of metabolism; a similar proportion of the TCI dose was excreted through the
                                                           14
lungs unchanged, and a similar proportion was metabolized (  C02 in exhaled
air plus total radioactivity in urine) after six months to that metabolized
after a single dose.  These observations indicate that at this relatively high
dose (1000 mg/kg), induction of TCI metabolism in response to chronic admini-
stration did not occur, although TCI was undergoing nearly maximal (80 percent)
metabolism in any case.  Hence, these observations might not pertain to "satur-
ating" doses of TCI (>2000 mg/kg; Figure 4-7).  However, although there was no
overall increase in the proportion of TCI metabolized, a change in the relative
amounts of the major metabolites (TCA and TCE) in urine was observed.  After
10 days of dosing and thereafter, the amount of TCA excreted in the urine was
double that found after a single dose (Table 4-5), while TCE (as the glucuro-
nide) decreased in proportion.   These observations suggest that the oxidative
metabolism of chloral to TCA is increased at the expense of the reductive path-
way of TCE with chronic dosing as suggested by Green and Prout, or alternatively,
the kinetics of urinary excretion of TCA and TCE do not achieve a steady-state
with daily dosing of TCI until  after 10 days, because of a relatively long half-
time for renal excretion of TCA.  A similar phenomenon has been observed in the
human (Section 4.2.2) and in the rat (Muller et al., 1972, 1974).
4.3  MEASURES OF EXPOSURE AND BODY BURDEN
     Trichloroethylene is principally absorbed by the lungs during industrial
or ambient air exposure.   However, as worker exposure periods and air concentra-
tions are likely to vary, air analysis does not necessarily provide a reliable
indication of true exposure.   An index of a worker's current exposure and
total body burden can be estimated in accordance with the amount of TCI absorbed
and retained as predicted by the known pharmacokinetics and metabolism of TCI.
Therefore, monitoring TCI levels in blood or expired air and measuring urinary
metabolites are the most common methods for determining the body burden (Stewart
et al., 1962, 1970, 1974; Morgan et al., 1970; Nomiyama, 1971; Ogata et al. ,
1971; Ertle et al., 1972; Ikeda et al., 1972; Muller et al., 1972, 1974;
                                    4-23

-------
         TABLE 4-10.   EFFECT OF  CHRONIC DAILY ORAL DOSING  OF  TCI  (1000  mg/kg)  ON  METABOLISM AND
                                   METABOLITE EXCRETION  IN  B6C3Fa  MICE*
Day
Unchanged
TPT rn~
i \* x *•**"' 2
Uri ne
Feces
As % of Dose (24-hr Collection)
1
10
Control**
180
Control
17.5
13.6
15.4
23.7
12.5
± 8
±8.6
± 8.6
± 6.2
± 2.0
6.0
6.3
5.3
2.3
4.2
±0.7
± 0.6
± 1.0
±1.0
± 0.6
41.8
52.3
46.9
48.3
50.7
± 6.5
± 7.8
± 10
± 6
± 5.1
13.1 ±9.5
16.4 ± 3.3
12.3 ± 1.9
1.9 ± 0.5
1.3 ± 0.5
Urine
TCA
As
i
7.
15.
8.
19.
10.
Urine
TCE
% of 14C-TCI Radioactivity
n 24-hr urine collection
0
6
4
9 ± 3.1
5 ± 2.4
88.5
81.9
88.1
77.4 ±
86.9 ±



2.8
1.8
 *N = 4 test, N = 3 control.

**Test and untreated control mice received single doses of 14C-TCI (1000 mg/kg:  10 uCi) on day 10 and day 180.

Source:  Green and Prout (1984).

-------
  700
  600
  500

  300
oc
D
  200
  100
          1     I     I     I     I
I     I     |     I     I    I    I    I    I    I
                                                             MEANS±SEM

                                                                 r2 . 0.995

                                                                  a • 0.318
                                    I
          I
              400        800      1200      1600       2000

                                     TCI DOSAGE, mg/kg
                        2400
2800
3200
Figure 4-7. Relationship between TCI dose and the amount of total urinary metabolite excreted
  Sr day by mice in each group. Values represent means ± SEM. N = 7-9 mice per group except
  r the 100 and 3200 TCI/kg groups where N = 3 and 4 respectively. The slope (0.318) and r2 value
coefficient (0.996) are for linear regression fit of the 100 -1600 mg/kg data points.

Source: Buben and O'Flahery (1984).
                                          4-25

-------
Pfaffli and Blackman,  1972;  Kimmerle and Eben,  1973a,b;  Lowry et a!.,  1974;
Vesterberg et al.,  1976).  These methods,  however,  are subject to high inter-
personal variations in the pharmacokinetics and metabolism of TCI that limit
their predictive value.   Some of the factors now known to contribute to these
variabilities are pulmonary  dysfunctional  states,  interindividual differences
in intrinsic liver metabolism capacity for TCI, modification of metabolism by
drugs and environmental  xenobiotics, age,  sex,  exercise or work load,  and body
anthropometry.   Stewart and  his associates (1974)  have been vigorous proponents
of the value of postexposure alveolar concentrations of TCI as measures of
recent exposure and body burden.  However, this method is highly dependent on
time of sampling and influenced by individual variations in metabolism of TCI.
Muller et al. (1974) have concluded that neither TCA excretion nor the sum of
urinary metabolites can be taken as representative parameters of previous
exposure, except possibly for 24-hr daily collections.  These workers favor
either alveolar TCI concentrations or blood analysis of TCA or TCI as the best
present indicators of exposure and body burden.  Monster et al. (1979) con-
cluded from their studies that the most "promising parameter for biological
monitoring in repeated exposure to TCI" is blood TCA concentration level,
because of the ease of sampling and GC analysis, the greater independence of
time after exposure, accumulation with chronic exposure and increased body
burden, and the smaller interindividual variation of this parameter.
     Greater and more precise information than is now available on the absorp-
tion, distribution and metabolism kinetics of TCI is required for better means
of estimating or predicting body burdens with exposure.   To this end, many in-
vestigators have developed mathematical models formulating the pharmacokinetics
of inhaled TCI and its metabolism (Fernandez et al., 1977; Droz and Fernandez,
1978; Feingold and Holaday,  1977; Gobbato and Mangivachhi, 1979; Sato et al.,
1977; Sato, 1979).
4.4  METABOLISM
4.4.1  Known Metabolites
     It has been known for 50 years that TCI is extensively metabolized in the
body to TCE, TCE-glucuronide ("urochloralic acid"), and TCA.  Following the
introduction in the 1930's of TCI as a general anesthetic, Barrett and co-
workers (1939) isolated TCA from the urine of dogs and human patients anesthe-
tized with TCI.  Powell (1945, 1947) demonstrated the presence of TCA in both

                                    4-26

-------
plasma and urine.   She also observed that TCA continued to be excreted for
more than 2 days after a single anesthetic exposure.   Butler (1949), who had
previously discovered that the hypnotic, chloral  hydrate, was metabolized to
TCE, TCE-glucuronide, and TCA (Butler, 1948), hypothesized that this widely
used hypnotic was  an intermediate metabolite of TCI.   His experiments with
dogs exposed to TCI clearly demonstrated that the principal  metabolites of TCI
and TCE were TCE-glucuronide and TCA.   However, Butler could not demonstrate
the presence of chloral hydrate in either plasma or urine.   Nonetheless, this
demonstration of biotransformation of TCI to TCE as well  as  to TCA was the
first proof that TCI anesthetic was converted to a metabolite with hypnotic
properties.  Trichloroethanol has a hypnotic potency comparable to that of
chloral hydrate (Marshall and Owens, 1954).   Since that time, the intermediate
metabolite, chloral hydrate, has been found in the plasma of rats (Kimmerle
and Eben, 1973a) and humans inhaling TCI (Cole et al., 1975), and has been
shown to be produced in the isolated rat liver upon perfusion with TCI (Bonse
et al., 1975) as well as with rat liver microsomal preparations (Byington and
Liebman, 1965; Ikeda et al., 1980; Miller and Guengerich, 1982) and with
isolated hepatocytes (Costa and Ivanetich, 1984).
     The principal site of metabolism of TCI is the liver, although Dalbey and
Bingham (1978) have demonstrated that the rat and guinea pig lung also metab-
olize TCI to TCE,  and TCA.  Other organ tissues,  such as kidney, spleen and
small intestine, may also metabolize TCI, as these have been shown to be sites
of cellular protein binding of metabolites from TCI (Bolt and Filser, 1977;
Banerjee and Van Duuren, 1978; Stott et al., 1982).
     In addition to the major end-product metabolites of TCI, chloral, TCE,
TCE-glucuronide and TCA, a number of other metabolites  have been found in
urine or exhaled breath (CO, C02, monochloro- and dichloroacetic acids, TCA-
glucuronide, oxalic acid, N-(hydroxyacetyl) aminoethanol  (HAEE), chloroform)
or ir\ vitro with microsomal systems (TCI-epoxide, glyoxylic  acid, CO).  These
metabolites are listed in Table 4-11.   Except for HAEE, oxalic acid, TCA-CoA
conjugate, and TCA-glucuronide, these metabolites have been established by
more than one investigator.  However,  chloroform, which may be an artifact of
analytical methodology, needs further confirmation.
     TCI-epoxide (1,1,2-trichloroethene oxide) is known to decompose in aqueous
physiological conditions (Table 4-11) to formic acid, dichloroacetic acid,
glyoxylic acid, and CO.  These metabolites may, therefore, be formed at cellular
extra-microsomal sites of TCI-epoxide formation under aqueous conditions.

                                    4-27

-------
    TABLE  4-11.   IDENTIFICATION  BY  GC/MS  OF  TCI  METABOLITES  IN  EXHALED  AIR
        AND  URINE OF  WISTAR  RATS AND  NMRI MICE AFTER  A  SINGLE ORAL  DOSE
                            OF 200  mg/kg  OF  14C-TCI
                                             Rat
                       Mouse
                                         % of  dose
                                        radioactivity
                     % of dose
                   radioactivity
Exhaled air
Unchanged 14C-TCI
14C02
Urine


52.0
1.9
41.2
% of

11.0
6.0
76.2
urine radioactivity
  Oxalic acid3
  Dichloroacetic acid
  N-(Hydroxyacetyl)-
    aminoethanol (HAAE)
  Trichloroacetic acid
  Trichloroethanol  (free)
  Trichloroethanol
    (conjugated)
 1.3                   0.7
 2.0                   0.1
 7.2
15.3
11.7
61.9
 4.1
 0.1
 0.1
94.2
 Metabolites identified by GC-MS;  female Wistar rats and NMRI  mice;  collec-
 tions for 72 hr.
Source:   DeKant et al.  (1984).

     Until recently, a thorough chemical isolation and identification of TCI
metabolites by rigorous methodologies (GS-MS, etc.), in one or more  species,
had not been attempted.  Dekant et al.  (1984) have now reported their results
                                           14
of a comparative study in rats and mice of   C-TCI metabolites in breath and
in urine, where metabolite isolation was accomplished by reverse-phase HPLC
and identification by radio-GC and GC-MS.  The metabolites identified and
their relative abundance are given in Table 4-11.   All metabolites found in
the rat were also found in the mouse.  The urine radioactivity represented 41
and 76 percent of the total radioactivity given to rats and mice, respectively,
                                    4-28

-------
              14
from a 200-mg   C-TCI/kg dose.   The radioactivity in feces represented less
than 5 percent of the dose with total  recovery in urine,  feces,  and breath of
93 to 98 percent.   Two new urinary metabolites were found in both species,
oxalic acid and HAAE.  Dekant et al.  also identified significant amounts of
HAAE in the urine of human volunteers  exposed to 200 ppm  (1096 mg/kg) TCI for
6 hours.   However, the occurrence of oxalic acid in urine and of COp in breath
have not as yet been identified in man as specific metabolites of TCI; to do
                                                            14
so would require administration of labeled TCI (radioactive   C-TCI or stable
        14
isotope   C-TCI) since both compounds  are also normal  endogenous compounds of
intermediary metabolism.
     The data of Dekant et al.  (Table  4-11) would indicate that TCA is not a
prominent urinary metabolite of TCI in NMRI mice, in comparison to Wistar rats
(or man, Section 4.2.2).  Green and associates (1982,  1984) have postulated a
TCA-CoA conjugate in the mouse, which  is slowly excreted  via the biliary
system with the feces.  Dekant et al.  (1984) have also suggested that an
intense enterohepatic recirculation of TCA (as an alkali-labile conjugate) may
occur in this species.  However, the small percentage  of  dose radioactivity
appearing in the feces of the mice (<5 percent), would not indicate this route
is important in the excretion of TCA.
     In contrast to the observations of Dekant et al., Green and Prout (1984)
found a much greater relative proportion of TCA to TCE in mice (Table 4-9),
and also they found no difference in the proportion (TCA/TCE) between mouse
and rat species, or between different  strains of these species.   The propor-
tions were unaffected after single oral doses between  10  and 2000 mg/kg.
Furthermore, the proportion of TCA/TCE increased with  chronic exposure to TCI
in the mouse (Green and Prout, 1984),  an observation which has been shown also
for the rat and for man (Section 4.2.2).  Green and Prout (1984) found that
not only were the relative proportions of the major metabolites TCA and TCE in
urine very similar in rats and mice but also minor metabolites (dichloroacetic
acid, monochloroacetic acid, free TCE) were also similar, and they therefore
concluded that there is not a major difference in pathways of TCI metabolism
in these two species.

4.4.2  Magnitude of TCI Metabolism:  Evidence of Dose-Dependent Metabolism
4.4.2.1 Man.  Estimates of the extent  of metabolism in man, as a percentage of
retained TCI, have been made from balance studies, i.e.,  accounting for a
                                    4-29

-------
retained dose after vapor exposure by measuring pulmonary elimination of TCI
and metabolites excreted into the urine.   Balance studies with isotopically
labeled TCI have not been reported in man.   Man appears to metabolize inhaled
TCI extensively (between 40 to 75 percent of the retained dose) as reported by
various investigators over the last three decades (Barrett et al., 1939;
Powell, 1945; Forssman and Holmqvist, 1953; Soucek and Vlachova, 1960; Bartonicek,
1963; Ogata et al., 1971; Ertle et al.,  1972; Muller et al., 1972, 1974, 1975;
Kimmerle and Eben,  1973a,b; Stewart et al., 1974; Fernandez et al., 1975,
1977; Vesterberg et al., 1976; Nomiyama and Nomiyama, 1971, 1974a,b, 1977;
Sato et al., 1977;  Monster et al., 1976,  1979).  These studies do not present
any evidence to indicate that a saturation of metabolism occurs in man.   The
data of Nomiyama and Nomiyama (1977) (Figure 4-8) and of Ikeda (1977), which
correlate urinary metabolite excretion with increasing levels of TCI exposure,
indicate that liver capacity for metabolism of TCI is nonlimiting, at least up
to an exposure of 315 ppm (1695 mg/m ) for 3 hours (equivalent to 25 mg TCI/kg
b.w.).  Furthermore, the alveolar to inspired air concentration ratio (25 per-
cent; retention 75 percent) at exposure equilibrium approximates the cardiac
output flow through the liver (25 percent) and suggests that TCI is completely
removed from blood by the liver in a single pass.  However, Feingold and
Holaday (1977) predicted from mathematical simulation models of TCI at anesthe-
tic concentration in man (2000 ppm; arterial concentration, 9.9 mg/dl) that
metabolism was rate-limited by saturation to 49 percent of uptake for an 8-hr
exposure.
     The problems encountered with balance studies in man are the difficulties
associated with accurate measurement of the retained dose of TCI from vapor
exposure, and the imprecision of the older methodologies for determinations of
urinary metabolites by extraction and colorimetry using modifications of the
Fujiwara reaction.   More recent careful studies, using GC methods, show that
after  single or repeated daily exposures from 50 to 380 ppm (274 to 2082
mg/m  ), an average of 11 percent of retained TCI is eliminated unchanged by
the  lungs (tj , 5 hr), 2 percent of the dose is eliminated as TCE by the lungs
            "i
(t^,  10 to 12 hr), and 58 percent is eliminated as urinary metabolites  (Fernandez
et  al., 1975, 1977; Muller et al., 1977; Monster et al., 1976, 1979).   Unac-
counted for  is nearly 30 percent of the TCI dose.  These investigators  suggest
additional pathways or  routes of elimination of one or more unknown metabolites.
This  suggestion  is particularly  apropos in  light of the recent  demonstration
                                    4-30

-------
                600
                                 100
200
300
                ENVIRONMENTAL TRICHLOROETHYLENE CONCENTRATION, ppm
Figure 4-8.   Relationship between environmental TCI concentration and urinary
             excretion of TCI metabolites in human urine.   Urine was collected
             for 6 days after exposure.  (Adapted from Nomiyama and Nomiyama,
             1977).

by Dekant et al.  (1984) of two new metabolites of TCI  excreted in the urine of
rats and mice, oxalic acid and HAEE.   Other reported but unconfirmed minor
urinary metabolites of TCI, such as chloroform (1 percent in rats, Muller
et al., 1974) and monochloroacetic acid (4 percent in  man, Soucek and Vlachova,
1960) do not account for the discrepancy.  On the other hand, early studies in
dogs by Owens and Marshall (1955) show that TCE-glucuronide is concentrated
and secreted in bile.  Several studies also show that TCE or TCA given orally
or intravenously is not fully recovered in the urine (Marshall and Owens,
1954; Owens and Marshall, 1955;  Paykoc and Powell, 1945; Muller et al.,  1974).
Bartonicek (1962) found that 8.4 percent of a TCI dose in humans is excreted
in the feces as TCE and TCA.  Part of the unaccounted TCI dose may also be
explained by gastrointestinal excretion; TCI can be expected to be rapidly
distributed into the GI lumen because of its high lipid solubility and ubiqui-
tous distribution.   In any case, these studies indicate at least 60 percent
and possibly as much as 90 percent of retained TCI from vapor exposures  of
less than 500 ppm (2690 mg/m ) is metabolized in the body.
                                    4-31

-------
4.4.2.2  Animals.   There is an increasing body of experimental evidence that
the magnitude of metabolism of TCI by rodents (rat and mouse) is dose-dependent
and possibly best described by Michaelis-Menten metabolism kinetics.   Thus, at
very high doses, TCI metabolism exhibits saturation kinetics in both the mouse
and the rat, but differently, in that evidence of approaching saturation
occurs in mice at about 2000 mg/kg (absolute dose—60 mg/animal), and in rats
at about 500 to 1000 mg/kg (absolute dose--125-250 mg/animal).  Hence, both
species have a considerable capacity to metabolize TCI.  Furthermore, the
maximal capacities of the rat versus the mouse appear to be more closely
related to relative body surface areas (on an absolute dose/animal basis) than
to body weight.
     The types of studies that have been done to investigate dose-metabolism
relationships are fourfold:  1) single-dose studies across species using
14
  C-TCI and urinary radioactivity and/or other methods as measures of metab-
olism, 2) kinetic models based on inhalation kinetics, 3) comparative balance
                                   14
studies between rat and mouse with   C-TCI at 1 or 2 dose levels, 4) multiple
dose levels studies in rat and/or mouse.
     Across species: Early studies by Butler (1948) demonstrated the extensive
metabolism of TCI in dogs for which he estimated the extent of metabolism (as
a percentage of the administered dose) to be 60 to 90 percent.  Muller et al.
(1982) recently compared the metabolism of TCI in chimpanzees, baboons, and
                 14
rhesus monkeys.    C-TCI was administered by intramuscular injection at a dose
level of 50 mg/kg.  Radioactivity excreted in urine and feces (as an index of
metabolism) ranged from 40 to 60 percent of the dose in chimpanzees, 11 to 28
percent in baboons, and 7 to 40 percent in rhesus monkeys.  Their results
showed a significant difference in excretion route between man/chimpanzee and
rhesus/baboon.  While man and chimpanzee excreted only from 2 to 5 percent of
the total eliminated TCI metabolites  in the feces, baboon and rhesus monkey
excreted 18 to 35 percent of the radioactivity in the  feces.  Rodents have
been reported to excrete in  feces 3 to 8 and 17 to 19  percent for the rat and
mouse, respectively, of a radioactive dose of TCI (Table 4-12, Prout et al.,
1984).  Green and Prout (1984) have directly demonstrated by  bile duct cannula-
tion that TCI metabolite(s)  are excreted in bile by both rats and mice.
                                    4-32

-------
            TABLE 4-12.  DEMONSTRATED METABOLITES OF TCI (OTHER THAN
        CHLORAL AND CHLORAL DERIVATIVES:  TCE, TCE-GLUCURONIDE, AND TCA)
Metabolite
TCI-epoxide
Glyoxylic acid
Carbon monoxide
System
rat, mouse
microsomes
rat, mouse
microsomes
rat, mouse
microsomes
Investigator
Miller and Guengerich,
Miller and Guengerich,
Miller and Guengerich,
Traylor et al. , 1977
Fetz et al . , 1978

1982
1982
1982
Carbon dioxide
Monochloroacetic acid
Dichloroacetic acid
Oxalic acid


N-(Hydroxyacetyl)-
  aminoethanol (HAAE)

TCA-glucuronide

TCA-CoA


Chloroform
exhaled air
rat and mouse
human and
rabbit urine
mouse and rat
urine

rat and mouse
urine
rat and mouse
urine

rat, mouse,
human urine

chimpanzee urine

mouse feces,
bile

rat adipose
tissue, serum,
urine
Parchman and Magee, 1982
Stott et al., 1982
Dekant et al., 1984
Green and Prout, 1984

Soucek and Vlachova, 1960
Ogata and Saeki, 1974

Green and Prout, 1984

Hathaway, 1980
Dekant et al., 1984
Green and Prout, 1984

Dekant et al., 1984
Dekant et al., 1984
Muller et al.,  1982

Green et al., 1982
Green and Prout, 1984

Pfaffenberger et al., 1980
Muller et al.,  1974
Products from TCI-epoxide decomposition in aqueous solution, at 37°C
Formic acid
Dichloroacetic acid
Carbon monoxide
Glyoxylic acid
                      Henschler et al., 1977, 1978
                      Kline and Van Duuren, 1977
                      Bonse et al., 1975
                      Miller and Guengerich, 1982
                                      4-33

-------
     Inhalation kinetics:   Filser and Bolt (1979) have exposed rats (Wistar)
to TCI in a closed system (desiccator jar chamber) and determined, from disap-
pearance of TCI from the chamber the inhalation pharmacokinetics, in accordance
with the following model:

       vi              k12     ,                       K
                                        V2          	    metabolism.
   Chamber                              rat
   atmosphere

They found that the metabolism of TCI was dose-dependent in the rat, with the
                                              o
saturation point occurring at 65 ppm (350 mg/m ).  Zero order V    was 210
                                                               fllCL/\
umol/hr/kg and first-order clearance (below 65 ppm) was 77 L/hr/kg b.w.
     Andersen et al.  (1980) also analyzed the pharmacokinetic behavior of TCI
in rats (Fischer 344) using a Michaelis-Menten approach to demonstrate dose-
dependent metabolism.  Exposures were conducted  in a 31-L battery jar chamber,
and the rate of depletion of atmosphere TCI was  determined for 30, 100, 1000,
3500, and 8000 ppm (116, 538, 5380, 18,830 and 43,040 mg/m3).  TCI uptake was
fitted to a four-compartment model as follows:
     TCI                               Blood,                           Fat,  bone,
  atmosphere     < _        liver, etc.    < _          etc.
                                              ffl            32
                                      Metabolites

 The  metabolism  of  TCI was  found  to  be dose-dependent,  with a  saturation point
 of about 1000 ppm  (5380  mg/m  ) (Km,  463  ppm).  The  kinetic parameters,  calcu-
 lated as molar  equivalents TCI,  were Vmav,  185 umoles/kg/hr;  K,  292 umoles/
                                       MlaX                     III
 hr/kg for uptake  from 31-L chamber.   A "whole animal "/air partition coefficient
 of 15.4 (as  compared to  9.0 for  blood/air)  was calculated and suggested to be
 a more accurate estimate of microsomal-to-gas distribution than a blood/gas
 constant.
      Balance studies:   For rodents  (rat  and mouse), the extent of metabolism
 after single doses of TCI  has been  estimated by  balance studies using isotopi-
 cally labeled TCI.  Daniel (1963) carried out the earliest study in rats.   The
                                     4-34

-------
                                     36
rats were dosed by stomach tube with   C£-1abe1ed TCI liquid (40 to 60 mg/kg
b.w.); 80 percent of the dose was eliminated via the lungs with a half-time of
elimination of about 5 hours; urinary metabolites TCE, TCE-glucuronide, and
TCA accounted for only 15 percent of the dose.   More recent studies show
considerably different results.
     Dekant et al. (1984) have compared the metabolism of TCI by balance
studies, with female rats (Wistar) and with female mice (NMRI).   These results
                         14
are shown in Table 4-9.    C-TCI was administered as a single dose by gavage
in corn oil at 200 mg/kg.  The absolute dose given to the rats,  based on their
experimental weights, was about 48 mg and to the mice about 5.1 mg.  Radioac-
tivity was determined in urine, feces, carcass, and exhaled air for 72 hours
post administration.   Virtually complete oral absorption occurred with a total
recovery of radioactivity of 93 to 98 percent.   For this single dose, rats
excreted through the lungs 52 percent of the dose unchanged (~25 mg), while
the mice excreted only 11 percent (~0.56 mg).  Hence, the rats metabolized an
average of 48 percent of their dose (23 mg) and the mice 89 percent (4.54 mg).
The ratio of the amounts metabolized by the rat versus the mouse (5.05) closely
approximates their relative surface areas, which as calculated from their
                                               2/3
experimental body weights (rat b.w./mouse b.w.)    is 4.46.   Therefore, the
metabolic capacity of rats and mice for TCI appears to be proportional to
their surface areas.   On a mg/kg basis the mice would appear to metabolize TCI
to an extent 1.9 times greater than rats.   Furthermore, the high proportion of
the dose (52 percent) excreted unchanged via the lungs would suggest that a
single, orally administered 200 mg/kg dose exceeded the overall  hepatic capacity
of the rat resulting in a significant first-pass effect.
     Stott et al.  (1982) exposed rats and mice to 10 and 600 ppm (54 and 3228
    3  14
mg/m )   C-TCI for 6 hours.   Their results are given in Tables 4-6 and 4-7.
For exposed mice,  virtually 100 percent of the body net uptake was metabolized
with no evidence of metabolic saturation.   The body burden (dose) from the
6-hr exposure was  estimated as 78.5 umol of TCI/kg body weight (10 mg/kg) and
3138 umol of TCI/kg (412 mg/kg), respectively.   In contrast, for rats 98
percent of the net uptake from 10 ppm exposure was metabolized but only 79
percent at 600 ppm (3228 mg/m ) exposure,  suggesting an incremental approach
to saturation of metabolism in this species.   For rats, the inhalation doses
were 35.8 umol/kg  (4.7 mg/kg) and 1075 umol/kg (141 mg/kg),  respectively.
These results indicate a saturation of metabolism in the rat.   Support for
                                    4-35

-------
this conclusion is found also in the observation that pulmonary elimination of
unchanged TCI was only 2 percent of the dose from 10 ppm (54 mg/m ) exposure
but 21 percent of the dose at 600 ppm (3228 mg/m ) (Table 4-7).  In absolute
amounts, mice metabolized 2.2 times more TCI per kg body weight than rats
after exposure to 10 ppm, and 3.6 times more after exposure to 600 ppm (3228
    o
mg/m ) (Tables 4-6 and 4-7).  However, on a relative body surface area basis,
the ratio for the amounts metabolized by the rat versus the mouse approximated
their relative body surface areas.
     Multiple dose studies:  Dose-dependent kinetics as opposed to first-order
kinetics requires multiple dose level studies to rigorously differentiate
first-order kinetics and to appropriately evaluate the associated kinetic
parameters.  Two investigative groups, Prout and Provan (1984) and Buben and
0'Flaherty (1984), recently have used this approach to examine the kinetics of
TCI metabolism in rats and in mice.
     Buben and 0'Flaherty administered doses of 0, 100, 200, 400, 800, 1600,
2400, and 3200 mg/kg of TCI in corn oil by intragastric intubation to male
Swiss-Cox mice in groups of 7 to 9 mice.  The control groups consisted of 24
mice.  These doses were given 5 days/week for 6 weeks.  The subchronic dosing
ensured that steady-state conditions  for metabolism of TCI and for metabolite(s)
excretion were achieved.  Metabolism  was measured by determining 24-hr urine
collections of TCA, TCE, and TCE-glucuronide, as determined by GC.  Of the two
other metabolites assayed,  ethylene glycol proved negative, and oxalic acid
was  not increased over excretion  by control animals.  Metabolites  in feces
represented  less than 5 percent of the amount in urine.  No significant  day-to-
day  variability  in  urinary  metabolites excreted or week-to-week trends was
found.
      The  use of  "total  urinary metabolites" (TCA, TCE,  and TCE-glucuronide)  as
an  index  of  metabolism  is  an approximation, as  stressed by Buben  and 0'Flaherty
themselves,  because there  is experimental  evidence  that HAEE  and  oxalic  acid
may be  significant  urinary metabolites  of  mice, 0.7  and 4.1 percent,  respec-
tively  (Table  4-12),  and COp in  exhaled  breath  may  amount  to  6 to 9 percent of
the TCI  dose (Tables  4-2,  4-6,  and 4-12).
      Thus,  "total  urinary metabolites"  may underestimate actual  metabolism by
 at least  20  percent or  more.   However,  urinary  TCA,  TCE,  and  TCE-glucuronide
 excretion,  as  the principal metabolites,  provides  an appropriate reflection of
 overall  TCI  metabolism.
                                     4-36

-------
     Figure 4-7 shows the relationship between dose and metabolism ("total
urinary metabolites") found by Buben and 0'Flaherty in mice.   Up to a dose of
about 1600 mg/kg, a linear relationship between dose and amount metabolized
obtained with a constant 27.5 percent of the dose metabolized to the urinary
metabolites TCA, TCE, and TCE-glucuronide.   This percentage of dose metabolized
is considerably lower than that observed above by Dekant et al.  (1984) and
Stott et al.  (1982) for mice.  Above the 1600 mg/kg dose, the metabolism of
TCI shows a rapid decrease of the percent of dose metabolized and approached a
saturation of metabolism at about 2400 mg/kg.   These observations would indicate
that TCI follows first-order metabolism (proportion to dose)  in the mouse up
to a relatively high dose of 1600 mg/kg (~48 mg/mouse) and thereafter dose-
dependent and then zero-order kinetics follow at higher doses.   Buben and
0'Flaherty did not fit their data to the Michaelis-Menten equation, although
it would have been instructive.
     Prout et al. (1984) have carried out a similar investigation in rats
(Osborne-Mendel and Alderly Park Wistar) and in mice (B6C3F1  and Swiss Webster).
However, in this case the dosage series (10, 500, 1000, and 2000 mg/kg) was
given as single doses (i.e., not subchronically) in corn oil  by stomach tube
   14
as   C-TCI.  The metabolism can be expected therefore to reflect that of
single doses and not of steady-state metabolic conditions.   But the use of
14
  C-TCI allows a complete estimate of total  metabolism.  Thus,  Prout et al.
                14
(1984) measured   C-radioactivity (expressed as percent of dose) in urine,
feces, and expired air over a collection period of 72 hours,  with total
recovery of administered dose-radioactivity of 92 to 98 percent.  Their results
are tabulated in Table 4-2 for Osborne-Mendel  rats and for B6C3F1 mice;
entirely similar results were observed for Wistar rats and Swiss-Webster mice,
i.e., the disposition of TCI was independent of rat or mouse  strain.   The data
of Table 4-2 indicate that for both the rat and mouse, the percent of the dose
metabolized decreased with increase of dose size, and in tandem, the percent
of the dose excreted via the lungs as unchanged TCI increased with increase of
dose size.  However, these results were more evident and occurred to a much
greater extent in the rat than in the mouse.  Thus, a plot of dose against the
portion of the dose metabolized, Figure 4-9, indicates that for the rat the
metabolism of TCI approaches saturation at a dose of about 1000 mg/kg (~200
mg/animal), while for the mouse the metabolism of TCI is still  linear up to a
dose of 2000 mg/kg (~60 mg/animal).  These results with the mouse are in
                                    4-37

-------
  2000
  1600
  l 200
Q
iU
N



8
   800
   400
                -O— B6C3F1 MICE

                -D	SWISS WEBSTER MICE

                -A—-A.P. RATS

                •O~OSBORNE-MENDEL RATS
                       500
 1000

TCI DOSAGE, mg/kg
                  1500
2000
 Figure 4-9.  Relationship between administered single oral doses of   c-TCI to rats and mice and
 amount of dose metabolized in 24 hr, expressed as mg/kg b.w., as calculated from 14c- radio-
 activity excreted in urine, feces, and expired air (other than unchanged 14C-TCI). Each data point
 represents 4 rats or mice.
 Source: Prout et al. (1984).
4-38

-------
accord with those of Buben and 0'Flaherty, who found that metabolism was not
saturated in mice up to doses of 2400 mg/kg (Figure 4-7).  It can be concluded,
therefore, that metabolism of TCI  is dose-dependent and saturable at doses
above 500 to 1000 mg/kg in the rat and 1600 to 2400 mg/kg in the mouse.   Both
species, therefore, exhibit a very high hepatic capacity to metabolize TCI.
At lower doses in these species, metabolism is nearly linearly related to TCI
dose, i.e., "first-order" metabolism prevails.  However, it is likely that in
both species TCI metabolism can be best described overall by Michaelis-Menten
kinetics, particularly at high dose levels.

4.4.3  Enzyme Pathways of Biotransformation
     Current knowledge of the details of the enzymatic biotransformation of
TCI derives mainly from jm vitro studies, principally with liver cell fractions,
and from iji vivo studies with end metabolite identification in exhaled breath,
urine, feces, and bile.  Figures 4-10, 4-11, and 4-12 summarize the probable
pathways involved in the metabolism of TCI as determined from those studies.
Among the halogenated hydrocarbons, TCI biotransformation appears to be unusual,
since neither dehalogenation nor removal of carbon atoms occurs in the first
step.  The initial biotransformation of TCI involves metabolism to chloral
hydrate and to formation of trichloroethylene epoxide (Figure 4-10).  The
formation of chloral/epoxide is the rate-limiting step in the metabolism of
TCI (Ikeda et al., 1980; Nomiyama and Nomiyama, 1979).  Intramolecular rear-
rangement with migration of a chlorine atom is presumed to occur in the forma-
tion of chloral hydrate, in order to explain the genesis of TCE and TCA.
Conversion of chloral hydrate to TCE and TCA is shown in Figure 4-11.
4.4.3.1 Formation of ch1ora1--Byington and Leibman (1965) conclusively identi-
fied chloral hydrate as the product of TCI when incubated with rat liver
microsomes fortified with NADPH and Oy (^450 niono-oxygenase system), thereby
verifying the hypothetical chloral hydrate formation proposed by Butler nearly
20 years earlier.  These workers also postulated an intermediate formation of
an epoxide (as did Powell in 1945), analogous to the known microsomal oxidation
to epoxides of dihydronaphthalene, aldrin, heptachlor and isodrin.  Costa and
Ivanetich (1984) found that chloral hydrate was the major metabolite in isolated
TCI-treated hepatocytes from phenobarbital-pretreated or untreated rats.
Daniel (1963) provided evidence that chloride migration must occur in chloral
                               oc
formation since no dilution of   Cl atoms with chlorine atoms from other body
                                    4-39

-------
CI


H
/CI    NADPH, 02
CI
                         \
                                                       H

                                           CHLORIDE    1
                                           	^ C
                                          MIGRATION ff
                                                      O
                                                                             — CI


—
ci>

EPOXIDE

1
OH
H2O
OH ""
-i_ckCI
1 XQ
H _
                                                                          CHLORAL
                                                     2HCI
                 o=c'
                    CI
     o

     o
               H
                         H
                          H
                        + CO
O


O
                                               C	C'
                                          CK         ^H
                                                                H20
                                       H
                                                        H
                                                                               CO2
             FORMIC ACID AND
            CARBON MONOXIDE
                                         GLYOXYLIC ACID
          Figure 4-10.  Postulated scheme for the metabolism of trichloroethylene to chloral
          (see Figure 4-11 for chloral metabolism) and to trichloroethylene epoxide and its
          metabolites.

          Source:  Miller and Guengerich (1982).
                                                4-40

-------
 MITOCHONDRIA
    CYTOSOL

     NAD +
   ALDEHYDE
DEHYDROGENASE
      C CI3 COOH
        [TCA]
                          CHLORAL
                          C CI3 CHO
   CYTOSOL
     NADH
   ALCOHOL
DEHYDROGENASE

    NADPH
   ALDEHYDE
  REDUCTASE
MICROSOMES
 NADPH, 02
                  C CI3 CH2OH
                     [TCE]
                                                  GLUCURONYL
                                                  TRANSFERASE
                                         C CI3 CH2O C6H9O6

                                         [TCEGLUCURONIDE]
          Figure 4-11. Metabolism of chloral hydrate.

          Source: Modified from Ikeda et al. (1980).
                           4-41

-------
Cl           Cl

 \      /      NADPH,  02
    C = C        	
 /      \      P-450
H           Cl
                                  Cl    . i     Cl      Lewis      H        Cl
                                   \/\/                .    \
                                     C —C
                                   /      \
                                  H          Cl      Cl-shift    0        C1N,
                                                                             TCA
                 —	*"    c—c-ci
                 Catalyzed    //     \
                                                                              TCE
                                                       Cl    0
                                                        \
                                                                  CHLORAL
                                      EPOXIDE
                                          I
                                          I
                                          I   H20
                                          4-
                                      OH   OH           rl    n    H20
                                Cl    I     \ ^- Cl

                                H"^^*^              LI
                                                         C-C-Cl 	> C12CHCOOH
                                                           \
                                                           H
                                                                     DICHLOROACETIC
                                                                     ACID

      C — C
    Cl
           \
              H
     GLYOXALIC
     ACID CHLORIDE
         oxidation
         hydrolysis
     HO
            \
               OH
       OXALIC
        ACID

Figure 4-12.
                                             phosphatidyl-
                                             ethanolamine
                                             alkylation
                                     r — r
                                            \
           N   -   CH2  -  CH2  -  0  -  phosphatidyl
                                          1   -  phosphatidyl
                                          I
                                              reduction
HO — CH2— C '
              NH — CH2 — CH2—OH
N-(HYDROXYACETYL)-AMINOETHANOL
               Postulated scheme for the metabolism of TCI based on urinary
               metabolite profiles of rats and mice.   From Dekant et al.,
               1984.   Suggested pathways are indicated by dashed arrows.
                                      4-42

-------
                                        36
pools was observed in the conversion of   Cl-TCI to its end products TCE and
TCA.  Microsomal P.5g-mediated chloral formation from TCI has been repeatedly
confirmed (Ikeda et al., 1980, Miller and Guengerich, 1982) and the formation
of the epoxide (1,1,2-trichloroethene oxide) has also been directly demonstrated
in liver microsomal preparations by trapping with p-nitrobenzyl pyridine
(Miller and Guengerich, 1982).  In addition, a specific P.™ binding spectrum
(Type I) in liver microsomes of rats and rabbits has been observed after
addition of TCI or TCI-epoxide (Kelley and Brown, 1974; Uehleke et al., 1977b;
Pelkonen and Vaino, 1975).   TCI also competitively inhibits the metabolism of
hexobarbital and aniline by microsomal fractions (Kelley and Brown, 1974), and
the magnitude of the binding and chloral formation is enhanced with microsomes
from animals pretreated with phenobarbital but not 3-methylcholanthrene (Pelkonen
and Vaino, 1975; Leibman and McAllister, 1967).   However, the question arises
whether the epoxide is an obligatory intermediate in the formation of chloral
(Miller and Guengerich, 1982).
     Attempts have been made to demonstrate chloral as a metabolite of TCI-
oxide.  TCI-oxide has been synthesized and its reactions studied under a
variety of conditions (ti  ~ 1.3 minutes in water at pH 7.4, 37°C) Bonse et al.,
1975; Kline and Van Duuren, 1977; Henschler et al., 1979; Miller and Guengerich,
1982).  Henschler et al. (1979) reported that, under physiological conditions
(Tris buffer, pH 7.4), TCI-oxide rearranged to dichloroacetic acid, CO, formic
acid, and glyoxylic acid, but not to chloral.   Entirely similar results were
observed by Miller and Guengerich (1982).   However, Bonse and Henschler (1976)
demonstrated that the TCI oxide can be forced to rearrange to chloral by the
catalytic action of Lewis acids (FeCl_, A1C1-, BF_), and Henschler and co-
workers proposed that the formation of chloral iji vivo could therefore be due
to catalytic action of the iron of P45Q in the trivalent form at the hydrophobic
microsomal site of formation of the oxide (Figure 4-2).  Although Miller and
Guengerich (1982) found that excess ferric iron salts catalyzed the rearrange-
ment of TCI-oxide to chloral in nonaqueous solutions (CH^Clp or CH3CN), chloral
was not formed in aqueous media even when iron salts, ferriprotoporphyrin IX
or purified cytochrome P4c0» were present.
     In studies with rat or mouse liver microsome fortified with NADPH and Op
(or a reconstituted P.™ enzyme system), Miller and Guengerich (1982) found
that TCI was metabolized to chloral, TCI-oxide, glyoxylic acid and CO.  Traylor
et al.  (1977) also found TCI converted to CO in microsomal incubations.
                                    4-43

-------
Addition of TCI-oxide to these systems did not produce chloral, but the oxide
disappeared with a tj  -17 seconds or, in the presence of rat liver epoxide
hydrolase, 9 seconds.   On the basis of the reaction kinetics of formation of
chloral and disappearance of TCI-oxide in microsomal preparations, Miller and
Guengerich (1982) proposed that TCI-oxide was not an obligatory intermediate
step in the formation of chloral.   Alternatively, these investigators postu-
lated an oxygenated TCI-P.-n transition state leading to two products:   chloral
(with chloride migration) and a separate epoxide formation, as shown in Figure
4-11.  The working scheme for the metabolism of TCI as presented in Figure 4-11
accounts for all the microsomal metabolites of TCI.  Miller and Guengerich
(1982) excluded the likelihood of a rearrangement of TCI-oxide or of TCI
glycol to chloral, because while rearrangement to chloral occurred in non-
aqueous media in the presence of excess ferric iron salt, the incubation of
TCI-oxide with purified P.50 did not result in rearrangement to chloral.
Since j_n vivo chloral is the primary metabolite of TCI metabolism, then in
view of Miller and Guengerich's (1982) postulate, the P.5Q-mediated reaction
to chloral may be preferential.  The possibility of TCI-epoxide being a secon-
dary, minor product in the metabolism of TCI may account for the lesser muta-
genic and oncogenic potential of TCI in comparison to other halogenated ethy-
lenes (Bolt et al., 1982; Dekant et a!., 1984).
     Van Dyke (1977) and Van Dyke and Chenoweth (1965) also proposed a direct
microsomal oxidation to chloral.  TCI conversion to chloral was postulated to
involve a chloronium ion transition intermediate state in concert with oxida-
tion, presumably by cytochrome P.™ system.  As illustrated, the mechanism
does not require an epoxide  intermediate.  The oxidative attack forces chloride
migration to the adjacent carbon if possible, as in the case of TCI; if not,
then the chlorine bonded to  the carbon, which is oxidatively attacked,  is
released (dechlorination).

                                           Cl+   0"
                          nn       n      '  \   I                     n
                        /Cl       C\  s   \ 4                   ^ 0
                 C  =  C       ' ->       C  	 C     -»•   CKC  —  C '
           A»T ^        ^ M        n S          U         ^         >• U
           LI            ^ n        \i i            n                   n

This mechanism  also results  in "reactive"  intermediates which  can  covalently
bind to cell constituents.
                                    4-44

-------
4.4.3.2 Metabolism of chloral hydrate—The enzymatic pathways for the biotrans-
formation of chloral hydrate formed from TCI exposure, to the plasma and
urinary metabolites TCE, TCE-glucuronide, and TCA, have received less attention
in recent years than the mechanism of chloral formation.   Current knowledge of
the enzymatic pathways of chloral hydrate biotransformation is outlined in
Figure 4-12.  Chloral hydrate is very rapidly metabolized i_n vivo with a
half-life of only a few minutes (Marshall and Owens, 1954; Breimer et al.,
1974; Cole et al., 1975).
     A number of investigators have concluded from i_n vitro studies that
chloral hydrate is a substrate for alcohol dehydrogenase purified from horse
liver and rat liver cytosol fractions (Butler, 1948, 1949; Marshall and Owens,
1954; Friedman and Cooper, 1960; Sellers et al., 1972b).   By this reaction,
with NADH as the required coenzyme, chloral hydrate is reduced to TCE (K ,
        -3
2.7 x 10  M, horse enzyme).  The reaction is essentially unidirectional, since
conversion back to chloral hydrate has not been observed with this enzyme
(Sellers et al., 1972b; Owens and Marshall, 1955; Friedman and Cooper, 1960).
Tabakoff et al.  (1974) studied the reduction of chloral to TCE and postulated
the presence of enzymes in rat liver cytosol and the presence of enzymes other
than alcohol dehydrogenase that are capable of reducing chloral.  Ikeda et al.
(1980) demonstrated at least two NADPH-dependent enzymes to be present in rat
liver cytosol (in addition to NADH-dependent alcohol dehydrogenase) with a K
           -3
of 6.0 x 10  M.   These investigators also found that rat liver microsomes in
the presence of NADPH and 02 oxidized TCE to chloral similarly to the oxidation
of methanol and ethanol to acetaldehyde (Liebler and DiCarli, 1969).
                                                                     -4
     Human red cells also reduce chloral hydrate to TCE (K , 4.0 x 10  M)
(Sellers et al., 1972b).  The efficiency of red cells and their relatively
large mass may explain the nearly undetectable levels of chloral hydrate in
plasma and urine after exposure to TCI, while TCE is prominent in plasma and
urine.  TCE-glucuronide present in plasma and urine is accepted as resulting
from conjugation by hepatic microsomal glucuronyl transferase.
     The origin of TCA as a plasma and urinary metabolite of chloral hydrate
is not defined as clearly.  The most likely reaction for the oxidation of
chloral hydrate to TCA would involve the well-known enzyme acetaldehyde dehy-
drogenase.  However, chloral hydrate has been reported not to be a substrate
for human acetaldehyde dehydrogenase (Kraemer and Deitrich, 1968; Blair and
Bodley, 1969; Sellers et al., 1972b).  Cooper and Friedman (1958) obtained
                                    4-45

-------
from rabbit liver acetone powder a partially purified form of an NAD-dependent
"chloral  hydrate dehydrogenase"  which was substrate-specific for chloral
hydrate.   Acetaldehyde was not a substrate but rather a markedly effective
inhibitor.   Chloral  dehydrogenase has not been identified in any other species
even though the rabbit, in comparison with the rat or man, poorly converts TCI
to TCA.   Furthermore, Cooper and Friedman (1958) reported that their enzyme
was not inhibited by disulfiram.  However, orally administered disulfiram is
known to markedly decrease excretion of TCA and TCE after TCI exposure to man
(Bartonicek and Teisinger, 1962) and to dogs and rats (Forssman, 1953, cited
in Soucek and Vlachova, 1960).
     In contrast to earlier studies on aldehyde dehydrogenase that found most
of this enzyme activity in the cytosol (Buttner, 1965), more recent studies
show that about 80 percent of total activity resides in the mitochondrial and
microsomal cellular fractions.  Tottmar et al. (1973) described at least two
different aldehyde dehydrogenases (NAD and NADP-dependent) in rat liver,
localized respectively in mitochondria and microsomes.  Grunett (1973) found
the NAD-dependent mitochondrial enzyme to have a broad substrate specificity,
but chloral hydrate was not a substrate.  However, Ikeda et al. (1980) found
that an aldehyde dehydrogenase prepared from rat liver mitochondria (NAD-
                                                  -3
dependent) converted chloral  to TCA  (K  , 62.5 x 10  M).  Furthermore, rat
liver mitochondria had the highest specific activity for TCA formation, cytosol
less, but the microsomal  fraction contributed little to the formation of TCA.
     Present evidence  indicates that  the  liver is the principal organ contribut-
ing to the metabolism  of  TCI  to TCE  and TCA.  For example, Bonse et al. (1975)
showed that TCI  perfused  through  the  isolated rat liver led to  the appearance
of TCA, TCE, and chloral  in the post  perfusate.  However,  other tissues may
also contribute  significantly.  Dal bey  and  Bingham (1978)  have  demonstrated
the metabolism  of TCI  by  the  isolated perfused  lungs  of rats and guinea pigs
to TCE and TCA,  and  pretreatment  of  rats  with phenobarbital  increased the  lung
metabolism to TCE.   Tabakoff  et al.  (1974)  found that  rat  brain cytosol was
capable of reducing  chloral to  TCE.   The  conversion  of chloral  in  the presence
of  NADPH  was 5  times faster than  with NADH,  the co-factor for  alcohol dehydro-
genase.   As  noted  above,  human  red blood  cells  actively  reduce chloral  to TCE.
4.4.3.3   Minor  Metabolites  and  Pathways—For rats  and mice exposed to TCI,
 several metabolites  (C02, oxalic  acid,  dichloroacetic acid,  HAAE;  Tables  4-11
 and 4-12) other than TCA, TCE,  and TCE-glucuronide  have  been identified in
                                     4-46

-------
exhaled air, urine, feces, or bile.  Some of these minor metabolites have also
been identified in human urine (dichloroacetic acid, HAAE).  Together, these
metabolites represent less than 15 percent of the TCI dose in rodents (Table
4-12).
     Dekant et al. (1984) have suggested that the appearance of small amounts
of dichloroacetic acid in the urine of rats and mice (Table 4-12), and possibly
of man (Table 4-11), stems from a limited rate of conversion of TCI-oxide to
chloral with escape of the epoxide from the hydrophobic environment of P^n,
and subsequent decomposition in hydrophilic conditions to dichloroacetic acid
and other compounds.
     COp from TCI metabolism has been postulated by Miller and Guengerich
(1982) to originate from hydrolytic rearrangement or cleavage of the epoxide
to formyl chloride and glyoxalic acid chloride with hydrolysis to formic acid
and glyoxalic acid, respectively (Figure 4-11).   These two metabolites, entering
the endogenous one- and two-carbon pools, are then metabolized further to CCL.
Green and Prout (1984) have shown that CCU may also be the product of decarbox-
ylation of TCA and dichloroacetic acid.   These investigators have administered
14
  C-TCI to rats and mice by gavage or intravenously and demonstrated that 7 to
                                                          14
15 percent of the dose was recovered in exhaled breath as   CCL within 24
hours.   Green and Prout suggest the formation of a coenzyme A ester of TCA
followed by degradation to oxalyl CoA and subsequent release of two moles of
co2.
     Dekant et al. (1984) have recently identified two new metabolites of TCI,
oxalic acid and HAAE, in the urine of rats and mice.  These investigators
suggest that oxalic acid may be formed as the end-product of enzymatic or
non-enzymatic cleavage of TCI-oxide as shown in Figure 4-13.   Hydrolytic ring
opening produces the glycol which is expected to undergo spontaneous elimina-
tion of two molecules of HC1 to form glyoxalic acid chloride with subsequent
hydrolysis to glyoxalic acid, which then is oxidized further to oxalic acid.
The spontaneous formation of glyoxalic acid from TCI-oxide decomposition in
aqueous solution has been amply demonstrated (Table 4-12).   HAAE also had not
been identified previously as a metabolite of TCI.   Dekant et al.  suggest that
TCI-oxide may interact with ethanolamine or phosphatidyl-ethanolamine in cell
membranes to produce HAAE as indicated in Figure 4-13.   The minor metabolites
of TCI may occur as follows:  if it is assumed that the conversion of TCI-oxide
to chloral is a rate-limiting step in the P.50-mediated metabolism of TCI,
                                    4-47

-------
then at high substrate concentrations of TCI an accumulation of TCI-oxide may
occur and activate alternative pathways leading to the formation of metabolites
other than chloral.

4.4.4  Metabolism and Covalent Binding
     The extent of irreversible or covalent binding of reactive intermediates
from the metabolism of TCI to cellular macromolecules (protein, RNA, DMA) has
been extensively investigated, both i_n vitro and i_n vivo, as a theoretical
basis for evaluating the mutagenic and carcinogenic potential of TCI (Bolt
et al., 1982).  The microsomal oxidative metabolism of TC! to chloral, as
outlined in Figures 4-11 and 4-13, suggest the possible reactive metabolites
that may result in covalent binding include chloral, TCI-epoxide, dichloroacetyl
chloride and formyl chloride.  Chloroform, a putative metabolite as yet of
indeterminate importance in TCI metabolism, is known to generate covalent
binding (Davidson et al., 1982).  The derivative metabolites of chloral (TCE,
TCE-glucuronide, TCA) have not been implicated in covalent binding.  Primary
concern accrues to the highly reactive epoxide formed in TCI metabolism.
     Henschler and coworkers (Bonse et al., 1975; Greim et al., 1975; Bonse
and Henschler, 1976; Henschler and Bonse, 1979) suggested that all chlorinated
ethylenes are initially transformed by microsomal P.™ systems to epoxide
intermediates which are subsequently "detoxified" by molecular rearrangement
(in the case of TCI to chloral).  Further, reactivities, and hence, toxicities
of individual epoxide intermediates depend on the type of chlorine substitu-
tions, in that symmetric substitution renders the epoxide relatively stable
and not mutagenic, but asymmetric substitution causes unstable and therefore
mutagenic epoxides.  To explain the lower-than-predicted mutagenic and carcino-
genic potential of TCI (Chapter 7), these investigators suggest that TCI  is an
exception to this general rule, because though its  epoxide is  asymmetrically
substituted and reactive, it  immediately  upon formation rearranges within the
hydrophobic premises of the  P.J-Q  site to  the less reactive chloral; thus  the
epoxide is  isolated from reaction with cellular macromolecules and protected
from aqueous  spontaneous decomposition to other reactive metabolites  (dichloro-
acetyl chloride,  formyl chloride) (Henschler et al.,  1979).  Alternatively,
Miller and  Guengerich  (1982)  have provided  evidence that the epoxide  is  not  an
obligatory  intermediate for  chloral  formation, and  that,  indeed,  the  two  are
separate  P.50 oxidative products  with  chloral production  four- to five-fold
                                     4-48

-------
greater than the epoxide.   Furthermore, they observed a four- to five-fold
greater production of chloral plus epoxide from mouse microsomes than from rat
microsomes.
4.4.4.1 In Vitro Binding.   Table 4-13 lists investigators who have demonstrated
covalent binding of TCI to microsomal protein nucleic acids and lipids i_n
vitro.   In many of these experiments, j_n vitro covalent binding from TCI
microsomal metabolism was decreased by animal pretreatment with SKF-525A
(which inhibits P.5Q metabolism) or enhanced by phenobarbital and other in-
ducers of the P450 system.  Binding was also decreased by addition of an
inhibitor of epoxide hydratase (trichloropropane epoxide) and by the addition
of competing nucleophiles such as reduced glutathione, imidazole or mercapto-
ethanol, indicating the importance of cellular sulfhydryl groups and glutathione
specifically for providing a "detoxification mechanism" for reactive inter-
mediates from TCI metabolism.  These experiments provide ample evidence that
TCI is metabolized jji vitro to reactive products by P.™ systems.  Metabolism
and covalent binding of reaction products occurs with liver microsomes, the
major site of metabolism, but also with microsomes from kidney, stomach, and
lung (Banerjee and Van Duuren, 1978).  Mouse microsomes are two- to four-fold
more active in metabolizing TCI to reactive intermediates than rat microsomes,
and males greater than females (Uehleke and Poplawski-Tabarelli, 1977; Banerjee
and Van Duuren, 1978).  DiRenzo et al. (1982a) found that aging rats (27 months
old) have less capacity for microsomal metabolism, as reflected by covalent
binding of TCI, than either adult (11 months old) or young (4 months) rats.
These investigators have also compared the rat hepatic microsomal covalent
binding to added DNA for a series of aliphatic ha!ides, including TCI.  The
degree to which TCI and other aliphatic halides form a DNA-adduct is summa-
rized in Table 4-14.  TCI was bioactivated and covalently bound to DNA at a
level of 0.36 nmol/mg DNA, or fifth in the series of 10 compounds.  Binding
was confirmed by sedimentation of the DNA-adduct in a cesium gradient and
Sephadex LH-20 chromatography of the nucleotides, although structural identifi-
cation of modified DNA bases was not made.
4.4.4.2 In Vivo Binding.  A number of investigators have determined the extent
of irreversible binding to cellular macromolecules following the administration
of TCI to rats and mice.  Bolt and Filser (1977) exposed rats (male Wistar)
for 5 hours in a closed system to initial concentrations of 100 and 1000 ppm
                  3  14
(538 and 5380 mg/m )   C-TCI and determined protein-bound metabolites in
                                    4-49

-------
                                  TABLE 4-13.   IN VITRO COVALENT BINDING OF TCI
      Investigator
        System
 Macromolecules
covalently bound
Additions
Van Duuren and Banerjee rat liver microsomes   microsomal  protein
(1976)
                                             SKF-525A or 7,8 benzoflavone- * binding; GSH,
                                             methylmercaptoimidazole, mercaptoethanol, urea-^
                                             binding; epoxide hydratase inhibitor (trichloro-
                                             propane epoxide) t binding; pretreatment with
                                             phenobarbital- t binding
Bolt and Filser (1977)  rat liver microsomes
Bolt et al.  (1977)

Uehleke and Poplawski-  rat and mice liver
Tabarelli (1977)        microsomes
                      albumin, polylysine    GSH, -J. binding
                      microsomal protein
                      lipid
Allemand et al.  (1978)  rat liver microsomes   microsomal  protein
Banerjee and Van Duuren Mice (B6C3F)  and  rat
(1978)                  (Osborne-Mendel)
                        microsomes
                      microsomal protein
                                              Salmon sperm DNA
Laib et al.  (1979)
rat liver microsomes   yeast RNA micro-
                      somal protein
DiRenzo et al.  (1982a)  rat liver microsomes   calf thymus DNA

                        rat liver microsomes   microsomal  protein
                        young (4 mo)  vs  adult
                        (11 mo) vs aged
                        (27 mo)
                       Mouse > rat; pretreatment with phenobarbital,
                       t binding

                       GSH or piperonyl butoxide - * binding; pretreat-
                       ment of animals with phenobarbital - t- binding

                       Mice > rats, males > females; microsomes from all
                       organs bind equally

                       Male > female enhanced by phenobarbital and 3-
                       methylcholanthrene admin; enhanced by addition
                       of epoxide hydratase inhibitor (trichloropropane
                       epoxide)

                       VC > TCI
                       TCI > VC

                       Rats pretreated with phenobarbital

                       * ^450 system in aged; 4- binding; pretreatment
                       with phenobarbital abolishes difference

-------
    TABLE 4-14.  MICROSOMAL BIOACTIVATION AND COVALENT BINDING OF ALIPHATIC
                          HALIDES TO CALF THYMUS DNA

                                                         Binding to DNA
  Aliphatic halides                                    (nmol/mg/h ± SD)

1,2-Dibromoethane                                       0.52 ± 0.14 (6)
Bromotrichloromethane                                   0.51 ± 0.18 (6)
Chloroform                                              0.46 ± 0.13 (6)
Carbon tetrachloride                                    0.39 ± 0.08 (6)
Trichloroethylene                                       0.36 ± 0.14 (7)
1,1,2-Trichloroethane                                   0.35 ± 0.07 (7)
Dichloromethane                                         0.11 ± 0.05 (5)
Halothane                                               0.08 ± 0.01 (6)
1,2-Dichloroethane                                      0.06 ± 0.02 (6)
1,1,1-Trichloroethane                                   0.05 ± 0.01 (3)

14C-labeled aliphatic halides (1 mM) were incubated with hepatic microsomes.
Carbon tetrachloride, bromotrichloromethane and halothane were incubated
under an N2 atmosphere, and all other incubations were under an 02 atmos-
phere.

Source:   DiRenzo et al. (1982b).


various organs.  The irreversibly bound metabolites were calculated as a

percentage of that amount of TCI that was taken up by the animals in 5 hours.

Radioactivity following exposure was mainly concentrated in the liver (0.77 -

0.88 percent dose), the primary organ of metabolism, and in kidneys (0.37 -

0.39 percent dose), the organ of primary excretion.  Lesser amounts were found

in lung, spleen, small  intestine, and muscle.   No major differences in percent-

age of the dose bound versus the different initial atmospheric concentrations

of TCI were observed.  However, the extent of binding in liver and kidney for

TCI was less than for vinyl chloride and approximated that observed with
carbon tetrachloride.

     Uehleke and Poplawski-Tabarelli (1977) determined in vivo binding for
                                         14
mice (NMRI) following i.p.  injections of   C-TCI (0.2 uCi/10 umole in 5 (jl of

peanut oil/g b.w.).  Binding to liver microsomes, mitochondria, and cytosol

protein was measured at 2,  6, 12, and 24 hours after injection.  Peak binding

occurred after 6 hours  (8,  4 and 1.8 nmol/mg protein) respectively, and there-
                                              14
after declined slowly.   When the same dose of   C-TCI was given by gavage at

the 6-hours time, binding to microsomal protein was found to be only 5.4 nmol/
             14
mg protein.    C-TCI was also bound to hepatic lipids; similarly to protein

binding, 6 hours after  i.p. dosage, the highest concentrations were 9.4 nmol/mg
                                    4-51

-------
of endoplasmic fractions of livers and 5.6 nmol/mg in mitochondria!  lipids of
the liver.
     Allemand et al.  (1978) investigated the metabolic activation of TCI and
                                                  14
irreversible binding in Sprague-Dawley rats after   C-TCI (100 umol, 100 uCi)
was administered i.p.  in 200 ul  of methanol to nonpretreated and to phenobar-
bital-pretreated rats; the rats  were sacrificed 4 hours later.  Trace amounts
of irreversibly bound radioactivity were found on muscle proteins (3.1 mmol
equiv./g tissue), while 40 times more was found on liver proteins (117 nmol
equiv./g tissue).  Pretreatment of the animals with phenobarbital did not
significantly increase muscle binding but increased liver binding by 46 percent.
Furthermore, TCI reduced hepatic glutathione levels 61 percent 4 hours after
administration, confirming similar observations by Moslen et al. (1977b).
     Parchman and Magee (1982) compared i_n vivo covalent binding of TCI metabo-
lites to liver protein and DNA in B6C3F1 male mice and Sprague-Dawley male
rats.  Rats were injected i.p. with 10, 100, 500, and 1000 mg/kg b.w. of
  C-TCI in corn oil (each animal received 160 uCi/kg of radioactivity).  After
6 hours, liver "cytoplasm" and "nucleus" fractions contained 2.5 to 6.1 and
0.3 to 3.9 percent of dose radioactivity, respectively.  Radioactivity was
detected in DNA isolated from the nucleus fraction and purified by cesium
chloride centrifugation and extraction with phenol, isoamyl alcohol, and
chloroform.  However, the level of radioactivity was extremely  low (35 cpm/mg
DNA) and contamination with labeled protein could not be entirely ruled out.
Identification of DNA adducts by hydrolysis followed by HPLC  of nucleotides
was  not successful.   Similar  results were obtained with mice  injected  i.p.
with 10 and 250 mg/kg 14C-TCI (160 uCi/kg b.w.) and sacrificed  6 hours later.
Hepatic DNA contained only 51 cpm/mg for the  lower dose and 283 cpm/mg for  the
higher dose.   Compared  to DNA binding  by dimethylnitrosamine  (DMN)  under
identical  conditions, the carcinogen-binding  index (Lutz,  1979) calculated  for
TCI  was only  2,  as compared to 2045 for DMN.  These  investigators concluded
that the ability of TCI to  interact with DNA  w vivo was very slight.
     Stott et al.  (1982)  investigated  covalent binding  of  TCI in male  B6C3F1
mice and male Osborne-Mendel  rats, the rodent strains  used for carcinogenicity
testing of TCI  by  the National Cancer  Institute  (NCI)  (NCI, 1976).   The  animals
were exposed  to  10 or 600  ppm (54  or  3228  mg/m  )  of    C-TCI (11.3 mCi/nmol)
for  6  hours.   The  mice  metabolized more  inhaled TCI  than rats at low and high
exposure  levels  (123  and  260  percent,  respectively).   The  amount of hepatic
                                     4-52

-------
and renal covalent binding was three-fold higher in mice than for the high
dose (Table 4-15) and four-fold greater for liver than kidney.   Mice (4) were
                              14
also dosed with 1200 mg/kg of   C-TCI with relatively high specific activity
(2.4 mCi/nmol) and a maximum estimate of a liver DMA alkylation level of
0.62 ± 0.42 alkylations/10  nucleotides was observed after isolation and
separation by HPLC (Table 4-16).   However, no chemical identification of
possible nucleotide adducts was made.  Maximum covalent binding index observed
was 0.12, or one-twentieth of that estimated by Parchman and Magee (1982), and
many orders of magnitude less than that for known carcinogens such as DMN,
methylnitrosurea, etc.  (footnote, Table 4-15).  These results are in accord
with the conclusion of Parchman and Magee (1982) that the ability of TCI to
interact with DMA i_n vivo is very slight.
     Stott et al. (1982) also chronically administered to male B6C3F1 mice (by
gavage) 2400 mg TCI/kg/day for 3 days.   Localized cell necrosis, enhanced DMA
synthesis, and centrilobular hepatocellular swelling were observed.   With pro-
longed exposure (3 weeks), the primary response observed was dose-related cen-
trilobular hepatocellular swelling and the occurrence of mineralized cells.
In Osborne-Mendel male rats, a maximum tolerated dose of 1100 mg/kg/day led to
increased liver weight and elevated DMA synthesis not coupled to histopathology,
but consistent with a simple increase in the number of normal hepatocytes.  In
view of the weak genotoxic potential of TCI, as indicated by only slight DNA
covalent binding and weak hepatotoxicity in mice, these investigators suggest
an epigenetic mechanism for the tumorigenie action of TCI observed in this
species only in the NCI carcinogenicity bioassay (NCI, 1976).  They implicate
also the greater covalent protein binding (three-fold) that occurs in the
mouse versus the rat from the three-fold greater total metabolism of TCI per
kilogram body weight.

4.4.5  TCI Metabolism:   Drug and Other Interactions
     A number of commonly used drugs might be expected to modify the extent of
metabolism of TCI during human exposure.  The converse may also occur, result-
ing in important modifications of the therapeutic action of drugs.  Although
largely undocumented in man, the induction of hepatic microsomal mixed-function
oxidate system by drugs taken for therapeutic reasons or by exposure to certain
environmental chemicals (e.g., barbiturates, PCBs, etc.), may bring about an
increased rate of TCI metabolism.  Numerous jji vivo and ijn vitro experimental
                                    4-53

-------
     TABLE 4-15.   HEPATIC  AND  RENAL  MACROMOLECULAR BINDING OF 1,1,2 (UL-14C)
             METABOLITE  IN MALE  B6C3F1 MICE AND OSBORNE-MENDEL RATS
                     EXPOSED TO  10 OR 600 PPM FOR 6 HOURS
Exposure Test
(ppm) Tissue animal
10 Liver Mouse
Rat
Kidney Mouse
Rat
600 Liver Mouse
Rat
Kidney Mouse
Rat
Binding (pmol-
eq(14C) TCI/ug
protein) ± SD
0.318
0.268
0.168
0.155
20.4
4.72
5.06
1.77
(0.004)
(0.013)
(0.014)
(0.007)
(2.49)
(0.415)
(0.667)
(0.200)
Ratio
(mouse/rat)
1.2
1.1
4.3
3.0
Source:   Stott et al., 1982.
       TABLE 4-16.   IN VIVO ALKYLATION OF HEPATIC DNA BY 1,1,2(UL-14C) TCI
         IN MALE B6C3F1 MICE DOSED WITH 1200 MG/KG (14C) TCI BY GAVAGE

Animal
1
2
3
4

Detection limit
(alk/104)
0.12
0.52
0.15
0.18

Non-Ci dpma
> 2 Sigma error
18.0
NDd
7.1
26.6
Maximum
alkylations 104
nucleotides
0.50
0.27
1.1

Maximum
CBIC
0.055
0.030
0.120
adpm not associated on HPLC analysis with normal DNA bases from normal metabolic
 incorporation and not attributable to counting error (i.e., mean background
 dpm + 2 SD).
 "Maximum" alkylations possible due to the relatively large amount of dpm
 associated with protein binding and metabolic incorporation into normal bases.

GCovalent binding index (Lutz, 1979).  Comparative hepatic CBI for dimethyl-
 nitrosamine, diethylnitrosamine, methylnitrosourea, and aflatoxin B-l are
 approximately 5500, 125, 640, 17,000, respectively.  CBI = (umol adduct/mol dN)
 (nmol/kg dose).
 None detected.
Source:  Stott et al.  (1982).
                                    4-54

-------
studies with animals pretreated with microsomal inducers have demonstrated an
enhanced metabolism of TCI (Leibman and McAllister, 1967; Carlson, 1974;
Moslen et a!., 1977a,b; Pessayre et al., 1979).  Conversely, TCI competitively
inhibits the metabolism of barbiturates, producing exaggerated effects of
these drugs (Kelley and Brown, 1974).
4.4.5.1  Disulfiram.  A drug interaction has been shown to occur in persons ex-
posed to TCI who are also on a therapeutic regimen of disulfiram for alcoholism.
Disulfiram inhibits the oxidation of acetaldehyde by competing with NAD for
active sites on aldehyde dehydrogenase.  Bartonicek and Teisinger (1962) showed
that disulfiram markedly inhibits the terminal enzymatic steps of TCI metabolism
in man.  Since these steps, which involve the conversion of chloral hydrate to
TCE and TCA (Figure 4-12), can be considered detoxification mechanisms, disul-
firam has the potential of enhancing TCI and chloral hydrate toxicities.
4.4.5.2  Alcohol.   Sellers et al. (1972b) investigated the interaction of
ethanol and chloral hydrate metabolisms.  These investigators observed that
concomitant administration of alcohol  and chloral hydrate in man exacerbated
the side effects of chloral hydrate exposure (vasodilatation, tachycardia,
facial flushing, headache and hypotension).   Their patients also exhibited
impairment of auditory function and of performance of complex motor tasks.
Similar effects have been observed during acute and chronic TCI exposure and
concomitant ingestion of alcohol (Muller et al., 1975).   Stewart et al. (1974)
have described the phenomenon of "degreasers" flush (facial and upper extremity
skin vasodilatation) and determined that this effect of alcohol interaction
with TCI persists after cessation of exposure.  This intolerance to alcohol
syndrome is ascribed to mutual inhibition of TCI and alcohol metabolism
producing increased plasma ethanol and TCE concentrations from competitive
inhibition of aldehyde dehydrogenase (Sellers et al., 1972a,b; Stewart et al.,
1974).
     Muller et al. (1975) specifically investigated the interaction of TCI and
alcohol in human volunteers.   After the volunteers inhaled TCI (50 ppm) for
6 hours per day for 5 days, ethanol ingestion inhibited TCI metabolism to TCE
and TCA by 40 percent.   Blood TCI concentration increased more than two-fold.
This latter observation suggested that alcohol also inhibited, by competition,
the microsomal oxidation of TCI to chloral hydrate.  These investigators
proposed that the intolerance syndrome from combined exposure to TCI and
ethanol was due to increased accumulation of TCI in the CNS, resulting from
                                    4-55

-------
depression of TCI oxidation.   From experiments in rats, Nakanishi  et al.
(1978) found that ethanol  ingested after TCI exposure increased blood acetal-
dehyde levels three- to four-fold.   They suggest that TCI inhibits acetaldehyde
dehydrogenase, and the resulting increase of blood levels of acetaldehyde is
responsible for the intolerance syndrome.
     White and Carlson (1981), working with rabbits, observed that acute
ethanol administration (1 g/kg i.v.  or p.o. 30 minutes prior to TCI exposure,
                    3
6000 ppm; 32280 mg/tn ) decreased TCI metabolism, as indicated by increased
peak blood levels of TCI (22 percent), decreased peak blood levels of TCE (50
percent) and decreased blood levels of TCA (99 percent).  Also, the zero-order
metabolism of alcohol itself was decreased 25 percent by TCI exposure (blood
ethanol disappearance rate, -0.707 control versus -0.539 exposed,  mg/dl/min).
Furthermore, rabbits treated with ethanol exhibited markedly increased suscep-
tibility to the development of TCI-induced cardiac arrhythmias, provoked by
epinephrine.  The results of this ethanol study in rabbits confirm those of
the human studies by Muller et al. (1975).
     In contrast to these studies, Sato et al. (1981) found that an acute
ethanol dose to rats (4 g/kg, 40 percent solution intragastric) significantly
lowered the blood concentration decay curve of TCI resulting from a subsequent
                                   3
(16 hours later) 400 ppm (2152 mg/m ) for 4-hr exposure to TCI.  Furthermore,
ethanol-treated rats excreted in urine much greater amounts (three-fold) of
total trichloro compounds (TCE and TCA) than untreated rats.  These findings
would indicate that a single large dose of ethanol can accelerate the in vivo
metabolism of TCI.  When ethanol (4 g/kg) was given to rats and liver microsomes
were prepared and tested for their rate of metabolism of TCI, an enhancement
of TCI metabolic rate was found, with the  greatest increase (two-fold) occurring
16 hours after ethanol dosing.  At this time, almost no blood ethanol was de-
tected and no increase occurred in microsomal protein and P.rQ liver contents.
Ethanol addition directly to the TCI microsomal incubations inhibited TCI
metabolism, thus demonstrating that ethanol was a competitor of TCI metabolism.
These  investigators  suggest that ethanol,  when taken i_n vivo, can exert  a dual
effect on drug-metabolizing enzymes,  i.e., both inhibition  and stimulation  of
TCI metabolism.  When present  in an early  period at  high blood levels, ethanol
inhibits metabolism  of TCI; but ethanol  also  increases  metabolic potential  of
the enzymes  so that  16 to 18  hours later,  when ethanol  blood level  has dissi-
pated, the microsomal enzymes  have an  increased capacity to metabolize TCI.
                                     4-56

-------
4.4.5.3 Warfarin.   Sellers and Koch-Weser (1970) observed a potentiation of
the anticoagulant effect of Warfarin in patients after chloral hydrate inges-
tion.   This increased potential for bleeding appeared to result from displace-
ment of Warfarin from plasma protein binding sites by the chloral hydrate
metabolite, TCA.  Considering the significant amounts of TCA formed after TCI
exposure and the long half-time of excretion of TCA, it is likely that TCI may
potentiate the effects of many other drugs which normally bind to the same
plasma protein sites as TCA.
4.4.5.4 Diet and Age.  Nakajima et al.  (1982) have explored the effect of diet
on the hepatic microsomal metabolism of TCI and other chlorinated hydrocarbons.
These investigators found that hepatic microsomes from rats fed for 3 weeks on
an isocaloric diet deficient in carbohydrate (sucrose) had an increased capacity
(two and one half-fold) to metabolize TCI and other chlorinated hydrocarbons.
Varying protein content of the diet produced only a negligible influence on
enzyme activity.  Microsomal  protein and P450 hepatic content (per g liver)
were not significantly changed, although liver weight was slightly decreased.
They suggest from their findings that a low-carbohydrate diet produces enhanced
P.™ system activity.  This concept, according to these investigators, is not
in conflict with the widely held belief that a high-protein diet enhances this
activity, because high-protein diets are usually low in carbohydrate to maintain
isocaloric conditions.  Thus, a diet rich in carbohydrate may offer significant
protection against the hepatic toxicity of chlorinated compounds by decelerating
their metabolism.
     DiRenzo et al. (1982a) have recently shown that the hepatic microsomal
capacity to metabolize TCI and other chlorinated hydrocarbons is influenced by
chronological age.  These investigators isolated hepatic microsomes from young
(4 months), adult (11 months) and aging (27 months) Fischer 344 rats.  Bio-
transformation and subsequent covalent binding of labeled TCI and other com-
pounds was studied.  Levels of microsomal protein and lipid binding increased
slightly between 4 and 11 months, but decreased 50 to 75 percent in the 27-
month-old animals, as compared to young adults.  Senescent rats were also
found to have significantly less hepatic P^Q (-35 percent), NADPH cytochrome
C reductase (-31 percent), and ethylmorphine N-demethylase activity (-43
percent).  Pretreatment with phenobarbital, however, abolished these defi-
ciencies of aging rats.  These investigators suggest a reduced metabolic
capacity as a function of age, which results in a decrease in bioactivation of
                                    4-57

-------
the aliphatic halides, and hence,  possible differences in toxicity in the
elderly from xenobiotics requiring bioactivation.

4.4.6  Metabolism and Cellular Toxicity
     The metabolism of TCI has been suggested as a possible vector for the
hepatotoxicity and nephrotoxicity associated with excess human exposure to
this haloalkene (James, 1963; Defalque, 1961; Smith, 1966; Litt and Cohen,
1969; Baerg and Kimberg, 1970).  Experimentally, TCI has been shown to be
hepatotoxic in dogs and rats (Orth and Gillespie, 1945; Klaassen and Plaa,
1967; Cornish and Adefuin, 1966; Carlson, 1974; Cornish et al., 1977; Reynolds
and Moslen, 1977).  Reynolds and Moslen (1977) and Henschler and Bonse (1979),
on physicochemical and biological correlative bases, have proposed that unsym-
metric chlorinated ethylenes like TCI are more hepatotoxic than symmetric
ethylenes as perch!oroethylene and more so than vinyl chloride.  The mechanism
for TCI hepatotoxicity is not definitely established, but many investigators
have proposed that the observed centrilobular cellular necrosis results from
the microsomal mixed-function oxidase system formation of chemically "reactive"
metabolites which react with and bind to tissue macromolecules and cause
intracellular damage  (Pessayre et al., 1979; Allemand et al., 1978; Reynolds
and Moslen, 1977).  Figures 4-11 and 4-13 indicate that these reactive metabo-
lites may  include TCI  epoxide, dichloroacetyl chloride, and  formyl chloride
(Section 4.4.4).  Chloral, chloral hydrate,  and TCE,  as ordinarily used  hyp-
notics, are  not  associated with significant  hepatotoxicity (Goodman and  Gilman,
1975).
     Buben  and 0'Flaherty (1984) have  recently  provided convincing evidence
that the hepatotoxicity  of TCI  is  directly proportional to the  extent  to which
TCI  is  metabolized.   These  investigators  administered to  male  Swiss-Cox  mice
by gavage  TCI  in corn oil vehicle  for  6 weeks  (5  days/week), at dose  levels of
100  to  3200 mg/kg.   Metabolism was measured  by  the  amount of urinary  metabo-
 lites  (TCA,  TCE, and TCE glucuronide)  excreted  per  day (24-hr urine  collec-
tions).   Indices of hepatotoxicity measured  in  the  dosage groups were increases
 in liver weight, decreases in liver  glucose-6-phosphatase (G6P) activity,
 increases  in liver triglycerides,  and  increases in  SGPT activity.   TCI signifi-
 cantly affected two of the four hepatotoxicity parameters,  liver weight, and
 G6P activity,  in a dose-related manner.   Increases  of liver triglycerides and
 increases  of serum SGPT occurred only  at the highest dosage levels.   However,
                                     4-58

-------
dose-related percentage increases of liver weight and of percentage inhibition
of G6P activity exhibited biphasic curves (Figures 4-13 and 4-14) exactly
analogous to the dose-related metabolism curve (Figure 4-8).   In each case,
the relationships between dose and toxic effect, and dose and metabolism were
linear up through 1600 mg of TCI/kg/day, but reaching a plateau at higher
doses.  Furthermore, when these same hepatotoxicity data points were plotted
against the amount of metabolism (i.e., total urinary metabolites) at each
respective dose, linear relationships were observed throughout the entire dose
range (Figures 4-13 and 4-14).  These observations indicate that the non-
linearity seen in the dose-toxic effect relationships can be fully explained
by the kinetics of TCI metabolism.  Chemical and histopathologic examination
confirmed that the dose-related toxic responses, liver weight increase, and
liver GBP activity decrease were paralleled by cellular change.  The increases
in liver size were attributable to hypertrophy of the liver cells, since a
dose-related decrease in the DNA concentration (mg/g liver) was found.   Micro-
scopic examination revealed evidence of dose-related cellular damage with
cellular degeneration, swollen hepatocytes, karyorrhexis, and central lobular
necrosis.  Hepatic cells with two or more nuclei or enlarged nuclei containing
increased amounts of chromatin were observed, suggesting that a regenerative
process was ongoing.
     As TCI hepatotoxicity in mice parallels TCI metabolism,  then covalent
binding of reactive metabolites to cellular macromolecules should also be
directly proportional to the extent to which TCI is metabolized.  Presently,
there is less information about a possible dose-related binding of "reactive"
metabolites to macromolecules that results in an incremental  increase in a
toxic liver response; nor have the critical target binding sites been identi-
fied.  However, numerous factors, such as species, age, sex,  and diet,  all of
which modify TCI microsomal metabolism, also markedly influence the toxic
response.  Drugs such as phenobarbital (Carlson, 1974) and ethanol (Cornish
and Adefuin, 1966) greatly influence TCI metabolism and hepatotoxicity.
Cornish et al. (1977) found that TCI-induced liver toxicity in rats, as
measured by plasma SGOT, is potentiated by both ethanol and phenobarbital.
Ethanol acted by competitive substrate inhibition, and also by stimulation of
microsomal mixed-function oxidase system; phenobarbital potentiated toxicity
by induction of mixed function oxidase system and thereby increased TCI metab-
olism.  Allemand et al. (1978) observed a similar potentiation of toxicity in
                                    4-59

-------
    80
  <70
  cc
  H

  >. 60
  Q
  O
  m
  ul 50
  HI

  o40
  z
    30
    20
  cc 10
                                                                       r2 = 0.974

                                                                       a = 0.036
                                                                       b = 6.27
           J	L
           J	I      I
                400


                1
800       1200       1600      2000
              TCI DOSAGE, mg/kg
                      T
                                  2400
2800
3200
                                                                      r2 = 0.974

                                                                      a = 0.105

                                                                      b = 6.64
                100
200
   300        400       500        600
TOTAL URINARY METABOLITE, mg/kg
 700
 800
Figure 4-13. Top: Dose-effect relationship between chronic daily oral TCI dose and increase in liver
weight of mice after six weeks.
Bottom:  Relationship between liver weight increase with chronic dosing at increasing levels (top) and
total urinary metabolite (TCA and TCE) excreted per day by mice at various dose levels.
Source: Buben and O'Flaherty (1984).
                  4-60

-------
   35
   30
   25
 +-«
 c
 8
 I
 zT 20
55
I  15
Q.
(O
O
   10
               400        800      1200       1600      2000
                                      TCI DOSAGE, mg/kg
                                    T
                       2400
2800
3200
                                                                r2 = 0.988
                                                                 a = 0.051
                                                                 b = 1.89
                        200
                                              400
                               TOTAL URINARY METABOLITES, mg/kg
                         600
            800
Figure 4-14. Top: Dose-effect relationship between chronic daily oral TCI dose and inhibition of
liver glucose-6-phosphatase activity (G6P) of mice after six weeks.
Bottom:  Relationship between inhibition of liver G6P activity with chronic dosing at increas-
ing levels (top) and total urinary metabolite (TCA and TCE) excreted per day by mice at the
various dosages.
Source: Buben and O'Flaherty (1984).
4-61

-------
Reduced glutathione had a protection effect against TCI liver toxicity.   TCI
acutely administered to rats initially decreased hepatic glutathione by 61
percent, but increased glutathione levels (106 percent) 16 hours later (Moslen
et a!., 1977b).   Glutathione added to microsomal preparations decreased binding
                        14
of reactive products of   C-TCI,  suggesting a glutathione conjugation of
metabolites.  Pessayre et al.  (1979) found that single or acute TCI administra-
tion to rats decreased cytochrome PA,-n of phenobarbital-treated rats, but not
                                                     14
untreated animals, and increased  covalent binding of   C-TCI metabolites to
microsomal protein.  Chronic administration to normal rats simultaneously
                                                14
decreased P.™ cytochrome content and increased   C-TCI metabolite binding,
but also induced other microsomal enzyme systems.
     Since acute TCI exposure rarely produces functional liver impairment in
man (Smith, 1966), whereas chronic exposure does (Litt and Cohen, 1969; Baerg
and Kimberg, 1970), Pessayre and  co-workers (1979) suggest that the explanation
may lie in the ability of TCI to  enhance its own metabolism and its own toxi-
city.   However, Green and Prout (1984), after daily gavage dosing of B6C3F1
mice for 180 days (1000 mg/kg), could find no evidence of an increase of the
amount of metabolism.  The proportion of TCA/TCE in the urine increased and
thus suggests a stimulation of the TCA pathway at the expense of TCE (Section
4.2.3).
     Poplawski-Tabarelli and Uehleke (1982) have investigated other mechanisms
that contribute to the toxic manifestations of halogenated aliphatic hydrocar-
bons,  including TCI.  These investigators attempted a correlation of the
inhibition  by these compounds of various microsomal oxidative reactions.  Of
the several possible factors influencing these mono-oxygenase activities, such
as interaction with lipid environment of microsomal cytochromes; ligand forma-
tion with  reduced  P.5Q cytochrome;  covalent interactions of metabolic  interme-
diates with cytochrome P450; destruction of cytochromes by  lipid peroxidation;
and production and  interaction of CO with cytochromes, the  overriding  correla-
tion appeared with  the simple physicochemical  factors  of high  lipid  solubility
and vapor  pressure.  The  ability of chlorinated  ethylene to  inhibit  mitochon-
drial  respiration  and  hence ATP production as  a  mechanism of cellular  toxicity
has also  been  investigated.  Takano and  Miyazaki  (1982)  found  that  all  of  a
series of  14 chlorinated  ethanes and ethylenes,  including TCI,  inhibited
oxygen consumption of  isolated rat  mitochondria;  their inhibitory  potency
increased  in the  order of the  number of  contained chlorines.
                                     4-62

-------
4.5 REFERENCES
Allemand,  H. ,  D.  Pessayre, V. Descatoire,  C.  Degott, G. Feldman,  and  J-P.
     Benhamou.  1978. Metaboli'- activation of trichloroethylene into a chemically
     reactive  metabolite  toxic to the  liver.   J.  Pharmacol.  Exptl. Therap.
     204:714-723.

Andersen, M.E., M. L. Gorgas,  R.A.  Jones and L.J.  Jenkins.  1980.  Determination
     of  the  kinetic constants for metabolism  of inhaled toxicants ijn  vivo
     using gas uptake  measurements.   Toxicol.  Appl.  Pharmacol.  54:100-116.

Astrand,  I.  and F. Gamberale.  1978.  Effects on  humans  of solvents in the
     inspiratory air:   A method of estimation of  uptake.  Environ. Res. 15:1-4.

Astrand,  I., and  P.  Ovrum. 1976.  Exposure  to  trichloroethylene.   I.   Uptake
     and  distribution  in  man.   Scand.  J. Work Environ,  and Health 4:199-211.

Baerg, R.D., and D.V.  Kimberg. 1970. Centrilobular  hepatic necrosis and acute
     renal failure  in  "solvent  sniffers.:   Ann.  Intern.  Med.  73:713-720.

Banerjee, S. ,  and  B.L.  Van Duuren. 1978. Covalent  binding of the  carcinogen
     trichloroethylene to  hepatic  microsomal proteins and  to  exogenous  DMA in
     vitro.  Cancer Res.  38:776-780.

Berkley,  J. , J. Bunch,  J.T.  and Bursey, et al. 1980. Gas chromatography mass
     spectrometry computer analysis  of volatile  halogenated  hydrocarbons  in
     man  and his environment—a  multimedia  environmental study.   Biomed.  Mass
     Spectrom.  7:139-147.

Barrett,  H.M.,  J.C.  Cunningham,  and J.H. Johnston.  1939.  Study of fate  in
     organism  of  some  chlorinated  hydrocarbons.   J.  Indust.  Hyg.  Toxicol.
     21:470-490.

Bartonicek,  V.  1962.  Metabolism  and excretion  of  trichloroethylene after
     inhalation by human subjects.  Brit. J. Ind. Med. 19:134-141.

Bartonicek, V., and J.  Teisinger. 1962. Effect of tetraethyl thiram disulphide
     (disulfiram)  on metabolism  of trichloroethylene in man.  Brit.  J.  Ind.
     Med. 19:216-221.

Blair, A.M.,  and  F.H.  Bodley.  1969.  Human liver  aldehyde dehydrogenase.
     Partial  purification and properties.  Can. J. Biochem. 47:265-271.

Bolt, H.M., and J.G. Filser.  1977. Irreversible binding  of chlorinated ethylenes
     to macromolecules.  Environ. Health Persp. 21:107-112.

Bolt, H.M.,  A.  Buchter,  L. Wolowski, D.L.  Gil, and W. Bolt. 1977. Incubation
     of  14C-trichloroethylene  vapor with rat  liver microsomes.    Uptake  of
     radioactivity and covalent  protein binding of metabolites.   Int.  Arch.
     Occup. Environ. Health 39:103-111.
                                    4-63

-------
Bolt, H.M.,  R.J.  Laib, and J.G. Filser. 1982. Reactive metabolites and carcino-
     genicity of halogenated ethylenes.  Biochem. Pharmacol. 31:1-4.

Bonse, G., and D. Henschler. 1976. Trichloroethylene.  CRC Crit. Rev. Toxicol.
     5:395.

Bonse, G., T. Urban, D. Reichart, and D. Henschler. 1975. Chemical reactivity,
     metabolic oxirane  formation and  biological reactivity of  chlorinated
     ethylenes in  the isolated  perfused rat  liver preparation.  Biochem.
     Pharmacol.  24:1829-1834.

Briemer,  D. ,  H.C.  Ketclaars, and I.M.  von Rossum.  1974.  Gas chromatographic
     determination  of  chloral  hydrate, trichloroethanol  and trichloroacetic
     acid in blood and in urine employing head-space analysis.    J. Chromatogr.
     88:55-63.

Buben, J. A.  and E. J. 0'Flaherty. 1984. Delineation of the role of metabolism
     in  the  hepatotoxicity  of  trichloroethylene and perchloroethylene:   a
     dose-effect study.  Toxicol. Appl. Pharmacol., in press.

Butler, T.C.   1948.  Metabolic fate of chloral  hydrate.   J.  Pharmacol.  Exptl.
     Therap.  92:49-58.

Butler, T.C.  1949.  Metabolic transformation of trichloroethylene.  J. Pharmacol.
     Exptl.  Therap. 97:84-92.

Buttner,  H.   1965.  Aldehyde and alcohol dehydrogenase activity in liver and
     kidney of the rat.  Biochem. Z. 341:300-314.

Byington, K.H. ,  and K.C.  Leibman.  1965. Metabolism of  trichloroethylene  in
     liver microsomes.  II.  Identification of the  reaction product as chloral
     hydrate.  Mol. Pharmacol.  1:147-154.

Carlson,  G.P. 1974.  Enhancement  of  the  hepatotoxicity of trichloroethylene by
     inducers of drug metabolism.  Res. Commun.  Chem. Path. Pharmacol. 7:637-640.

Cole, W.J.,  R.G.  Mitchell, and R.F.  Salamonsen.  1975. Isolation, characteriza-
     tion and quantitation of chloral  hydrate as  a transient  metabolite  of
     trichloroethylene in  man  using electron capture gas chromatography and
     mass fragmentography.  J. Pharm. Pharmacol.  27:167.

Conkle,  J.P., B.J.  Camp,  and  B.E.  Welch.  1975.  Trace composition  of  human
     respiratory gas.  Arch. Environ. Hlth. 30:290-295.

Cooper, J.R., and P.J. Friedman. 1958.  The enzymic  oxidation of  chloral  hydrate
     to trichloroacetic acid.  Biochem. Pharmacol.  1:76-82.

Cornish,  H.H.,  and  J.  Adefuin.  1966.  Ethanol  potentiation of  halogenated
     aliphatic solvent toxicity.  Am. Ind.  Hyg.  Ass. J.  27:57-61.

Cornish,  H.H., M.L.  Barth,  and B. Ling. 1977.  Influence of aliphatic alcohols
     on  the  hepatic response to halogenated olefins.   Environ.  Hlth.  Persp.
     21:149-152.
                                    4-64

-------
Costa, A.  K.  and K.  M.  Ivanetich.  1984.  Chlorinated ethylenes:   their metabo-
     lism  and  effect on DNA repair in rat hepatocytes.   Carcinogenesis 5(12):
     1629-1636.

Dalbey,  W. ,  and  E.  Bingham.  1978. Metabolism  of trichloroehtylene by  the
     isolated perfused  lung.  Toxicol. Appl. Pharmacol.  43:267-277.

Daniel,  J.W.  1963.  The metabolism of  36Cl-labelled trichloroethylene  and
     tetrachloroethylene in the rat.  Biochem.  Pharmacol. 12:795-802.

Davidson,  I.W.F., D.D.  Sumner, and J.C. Parker. 1982. Chloroform:   a review  of
     its metabolism, teratogenic, mutagenic, and  carcinogenic potential.   Drug
     and Chem. Toxicol. 5:1-87.

Dekant, W., M. Metzler, and D. Henschler. 1984. Novel metabolites of trichloro-
     ethylene  through  dechlorination reactions in  rats,  mice  and humans.
     Biochem. Pharmacol. 33:2021-2027.

DiRenzo, A.B.,  A.J.  Gandolfi,  I.G.  Sipes,  and  J.N.  McDougal.  1982a.  Effect of
     senescence  on  the  bioactivation  of aliphatic  halides.  Res.  Comm.  Chem.
     Path. Pharmacol. 36:493-502.

DiRenzo, A.B.,  A.J.  Gandolfi,  and  I.G.  Sipes.  1982b.  Microsomal  bioactivation
     and covalent  binding  of aliphatic halides to  DNA.   Toxicol.  Lett. 11:
     243-252.

Droz, P.O.,  and J.G.  Fernandez.  1978. Trichloroethylene  exposure.   Biological
     monitoring  by  breath  and urine analysis.   Brit. J.  Ind. Med.  35:35-42.

Eger, E.I., and C.P. Larson. 1964.  Anesthetic solubility in blood and tissues;
     values and significance.  Brit. J.  Anaesth.  36:140-149.

Ertle, T. , D.  Henschler,  G. Muller, and  M.  Spassowski.  1972. Metabolism  of
     trichloroethylene  in  man.   I.   The significance of trichloroethanol  in
     long-term exposure conditions.  Arch. Toxikol. 29:171-188.

Feingold, A.,  and  D.A.  Holaday.  1977. The  pharmacokinetics  of  metabolism  of
     inhalation anesthetics.  Brit. J. Anaesth. 49:155-162.

Fernandez, J.G.,  B.E.  Humbert,  P.O.  Droz, and J.R. Caperos. 1975. Exposition
     au trichloroethylene,  Bilan de Tabsorption, de Texcretion et du  metab-
     olisme sur des sujects humains.  Arch. Mai.  Prof. 36:(7-8)397-407.

Fernandez, J.G.,  P.O.  Droz, B.E.  Humbert, and J.R. Caperos. 1977. Trichloro-
     ethylene exposure.   Simulation of uptake,  excretion, and metabolism using
     a mathematical model.   Brit. J. Ind. Med.  34:43-55.

Fetz, H. ,  W.R.  Hoos,  and D. Henschler.  1978.  On  the metabolic formation  of
     carbon  monoxide from  trichloroethylene.   Naunyn-Schmiedeberg's Arch.
     Pharmacol. 302:Suppl.  Abstr. 88.

Filser, J.G., and  H.M.  Bolt.  1979. Pharmacokinetics of halogenated ethylenes
     in rats.  Arch. Toxicol.  42:123-136.
                                    4-65

-------
Forssman, S. ,  and C.E.  Holmqvist.  1953.  The  relation  between inhaled and
     exhaled trichloroethylene and trichloroacetic  acid excreted  in  the  urine
     of  rats  exposed to trichloroethylene.   Acta Pharmcol.  et Toxicol.  9:
     235-244.

Friedman, P.J., and  J.R.  Cooper.  1960. The  role  of alcohol  dehydrogenase in
     the metabolism  of chloral  hydrate.  J.  Pharmacol.  Exptl.  Therap. 129:
     373-376.

Fujiwara, K.  1914.  Sitzungsker  u.  Abhandl naturforsch.   Ges.  Postock.  6:33.

Gobbato, F.   and C.  Mangiavacchi.  1979. Mathematical model for the simulation
     of the  turnover of an industrial toxicant (solvent)  subject to metabolism
     transformation.  G. Ital.  Med.  Lav. 1:53-60.

Goodman, L.S.  and A. Gilman. 1975. The Pharmacological  Basis of Therapeutics.
     5th ed. Macmillan, New York.

Green, T. ,  M.  S.  Prout, and W.  N. Provan.  1984.  Abstr.  8th  Eur.  Workshop on
     Drug Metabolism, Leige, p.  130.

Green, T. and M.  S.  Prout.  1984.  Species differences in  response to trichloro-
     ethylene.  II.  Biotransformation  in  rats and mice.   Toxicol.  Appl.
     Pharmacol.   In press.

Greim, H., B. Bonze, Z. Radwan,  D.  Reichart, and  D. Henschler. 1977. Mutageni-
     city and chromosomal  aberrations  as an  analytical  tool  for i_n  vitro
     detection of mammalian enzyme-mediated  formation of  reactive metabolites.
     Arch.  Toxicol.  39:159-169.

Grunett, N.  1973. Oxidation  of  acetaldehyde  by  rat liver  mitochondria  in
     relation  to  ethanol  oxidation  and the  transport of  reducing equivalents
     across the mitochondrial membrane.  Eur. J.  Biochem.  35:236-243.

Hathaway, D.E.  1980. Consideration  of the  evidence for  mechanisms  of tri-
     chloroethylene  metabolism including  new identification of its  dichloro-
     acetic  and trichloroacetic metabolites in mice.   Cancer Lett.  8:262-269.

Helliwell,  P.O.,  and  A.J.  Mutton.  1950.  Trichloroethylene  anesthesia.   I.
     Distribution in the fetal and maternal  circulation of pregnant sheep and
     goats.   Anaes.   5:4-15.

Henschler,  D.  1977.  Metabolism  and  mutagenicity  of halogenated olefins  -  a
     comparison of  structure and activity.   Environ.  Health Persp.  21:61-64.

Henschler,  D.,  and  G.  Bonse.  1979.  In: Advances in Pharmacology and Therapeu-
     tics,  Proc.  7th Int.  Congr.  Pharmacol.  Paris, 1978.  Vol. 9, Toxicology,
     pp. 123-130.    Pergamon  Press, Oxford.

Henschler,  D. , W.R.  Hoos,  H. Fetz, E.  Dallmeier and M.  Metzler. 1979.  Reactions
     of  trichloroethylene  epoxide in aqueous systems.   Biochem.  Pharmacol.
     29:543-548.
                                     4-66

-------
Ikeda, M.  1977.  Metabolism of  trichloroethylene and tetrachloroethylene  in
     human subjects.  Environ. Health Persp. 21:239-245.

Ikeda, M. , Y. Miyake, M.  Ogata,  and  S.  Ohmori.  1980.  Metabolism of trichloro-
     ethylene.  Biochem. Pharmacol. 29:2983-2992.

Ikeda, M. , H. Ohtsuji,  T.  Imamura, and  Y.  Komoike.  1972.  Urinary excretion of
     total trichloro-compounds,  trichloroethanol,  and trichloroacetic acid as
     a measure  of  exposure  to  trichloroethylene  and tetrachloroethylene.
     Brit. J. Ind. Med.  29:328-333.

Jakobson, I., J.E.  Wahlberg,  B.  Holmberg,  and  G. Johannsson.  1982.  Uptake via
     the  blood and  elimination of 10 organic solvents following epicutaneous
     exposure of  anesthetized guinea  pigs.   Toxicol. Appl.  Pharmacol.  63:
     181-187.

James, W.R.L. 1963.  Fatal  addiction  to  trichloroethylene.   Brit.  J.  Ind.  Med.
     20:47-49.

Kelley, J.M., and B.R. Brown, Jr. 1974. Biotransformation of  trichloroethylene.
     Intern.  Anesthesiol. Clin. 12:85-92.

Kimmerle, G. ,  and A. Eben. 1973a. Metabolism,  excretion and toxicology  of
     trichloroethylene  after  inhalation.   I.   Experimental exposure  on rats.
     Arch. Toxikol.  30:115-126.

Kimmerle, G. , and A. Eben. 1973b. Metabolism  studies of trichloroethylene.
     II.   Experimental human exposure.  Arch. Toxikol. 30:127-138.

Klaassen, C.D., and G.L.  Plaa.  1967. Relative effects of various chlorinated
     hydrocarbons on liver and  kidney function in dogs.  Toxicol.  Appl.
     Pharmacol.  10:119-131.

Kline, S.A.,  and B.L. Van Duuren. 1977. Reaction of epoxy-l,l,2-trichloroethane
     with nucleophiles.   J. Heterocycl. Chem. 14:455-458.

Kraemer,   R.J.,  and  R.A.  Deitrich. 1968.  Isolation and characterization  of
     human liver aldehyde dehydrogenase.  J. Biol.  Chem. 243:6402-6408.

Laib,  R.J.,  G.  Stockle,  H.M.  Bolt,  and W. Kung.  1979.  Vinyl chloride and
     trichloroethylene:   comparison  of  alkylating  effects of metabolites  and
     induction or preneoplastic  enzyme deficiencies  in rat liver.  J. Cancer
     Res. Clin.  Oncol. 94:139-147.

Leibman,   K.C., and  W.J.  McAllister,  Jr.  1967.  Metabolism  of trichloroethylene
     in  liver microsomes.   III.   Induction of the  enzymic  activity and  its
     effect  on excretion  of  metabolites.   J. Pharmacol.  Exptl.  Therap.  157:
     574-580.

Liebler,   C.S., and  L.M. DiCarli. 1969.  Hepatic  microsomes,   a new  site  for
     ethanol  oxidation.   J. Clin. Invest. 47:62  (Abst.).

Litt, K.F.,  and M.I. Cohen. 1969. Dangei—vapor  harmful.  Spot-removing sniffing.
     N. Eng.  J.  Med. 281:543-544.
                                    4-67

-------
Lowry, L.K. , R. Vandervort, and P. L. Polakoff. 1974. Biological  indicators  of
     occupational  exposure to  trichloroethylene.   J.  Occup. Med.  16:98-101.

Lutz, W.K. 1979.  In  vivo covalent binding of  organic  chemicals to DNA as a
     qualitative indicator in  the process of chemical  carcinogenesis.  Mutat.
     Res.  65:289-356.

Marshall,  E.K., and A.M. Owens. 1954. Absorption, excretion and metabolic fate
     of chloral hydrate.  Bull. Johns Hopkins Hosp.  95:1-18.

McConnell, G.,  D.M. Ferguson, and C.R.  Pearson. 1975. Chlorinated  hydrocarbons
     and the environment.  Endeavor 34:13-18.

Migdal, A.,  E.  Graczyk, Z.  Obodecka, and K.  Piesiak.  1971. Biochemical and
     neurological   studies  in  acute  oral  poisoning with  trichloroethylene.
     Wiad. Lek. 24:1669-2673.

Miller, R.E.,  and P.P.  Guengerich.   1982. Oxidation of trichloroethylene by
     liver microsomal cytochrome  P-450:   evidence for  chlorine migration in a
     transition state  not involving  trichloroethylene oxide.    Biochemistry
     21:1090-1097.

Monster, A.C., G.  Boersman, and W.C. Duba. 1976.  Pharmacokinetics  of trichloro-
     ethylene in volunteers:    Influence of workload and exposure concentration.
     Int.  Arch. Occup.  Environ. Health 38:87-102.

Monster,  A.C.,  G.  Boersma, and W.C.  Duba.  1979.  Kinetics  of trichloroethylene
     in repeated  exposure  of volunteers.  Int.  Arch.  Occup.  Environ.  Health
     42:283-292.

Morgan, A.,  A.  Black,  and  D.R. Belcher.  1970.  The excretion in breath of some
     aliphatic  halogenated hydrocarbons  following administration by  inhalation.
     Ann. Occup. Hyg. 13:219-233.

Moslen, M.T.,  E.S. Reynolds, and  S.  Szabo. 1977a.  Enhancement  of the metabolism
     and  hepatotoxicity of trichloroethylene and perchloroethylene.   Biochem.
     Pharmacol. 26:369.

Moslen, M.T.,  E.S. Reynolds,  P.O.  Boor,  K.  Bailye,  and S.  Szabo.  1977b.  Tri-
     chloroethylene- induced  deactivation  of  P-450 and  loss  of  liver  glutathione
      in vivo.   Res.  Comm.  Chem. Path.  Pharmacol.  16:109-120.

Muller, G.,  M.  Spassovski, and D. Henschler.  1972.  Trichloroethylene exposure,
      and  trichloroethylene metabolites in urine  and blood.  Arch.  Toxicol.
      29:335-340.

Muller, G. ,  M. Spassovski,  and D.  Henschler.  1974. Metabolism of trichloro-
      ethylene  in  man.   II.   Pharmacokinetics of metabolites.   Arch.  Toxic.
      32:283-295.

Muller,  G. ,  M. Spassovski,  and D.  Henschler.  1975. Metabolism  of trichloro-
      ethylene  in  man.   III.   Interaction of  trichloroethylene  and  ethanol.
      Arch.  Toxicol.  33:173-189.
                                     4-68

-------
Muller, W.F.,  F.  Coulston,  and F. Korte.  1982.  Comparative metabolism  of
     14C-trich1oroethylene  in  chimpanzees, baboons  and  rhesus monkeys.
     Chemosphere 11:215-218.

Nakanishi, S. ,  E.  Shiohara, M.  Tsukada, Y.  Yamazaki,  and K. Okumura. 1978.
     Acetaldehyde level in the blood and liver aldehyde dehydrogenase activities
     in trichloroethylene-treated rats.  Arch.  Toxicol. 41:207-214.

Nakajima,  T., Y. Koyama, and A.  Sato.  1982. Dietary modification of metabolism
     and toxicity of  chemical  substances with special  reference to  carbohy-
     drates.   Biochem. Pharmacol.  31:1005-1011.

National Cancer Institute. 1976. Carcinogenesis bioassays of trichloroethylene.
     CAS No.  79-01-6 DEW Publ. No.  (NIH) 76-802.

Nomiyama,   K.  1971.  Estimation  of  trichloroethylene exposure by biological
     materials.  Intl. Arch. Arbeitsmed. 27:281-292.

Nomiyama,  K., and H. Nomiyama. 1971. Metabolism of trichloroethylene in humans.
     Sex difference in urinary excretion of trichloroacetic acid and trichloro-
     ethanol.  Intl.  Arch.  Arbeitsmed.  28:37-48.

Nomiyama,  K., and H. Nomiyama. 1974a.  Respiratory retention, uptake and excre-
     tion of organic solvents in man.   Int. Arch. Arbeitsmed. 32:75-83.

Nomiyama,   K. ,  and H.  Nomiyama.  1974b.  Respiratory elimination of organic
     solvents in man.   Int. Arch.  Arbeitsmed. 32:85-91.

Nomiyama,  K. ,  and  H.  Nomiyama.  1977.  Trichloroethylene metabolism in man and
     animals, with  special  reference  to host and agent factors modifying the
     trichloroethylene metabolism.   Ind. Environ. Xenobiot. 2:173-176.

Nomiyama,  H. ,  and  K.  Nomiyama.  1979.  Pathway  and  rate of metabolism  of  tri-
     chloroethylene in rats and rabbits.  Ind.  Hlth. 17:29-37.

Ogata, M. , and T.  Saeki. 1974. Measurement of chloral  hydrate, trichloroacetic
     acid, and  monoacetic  acid  in  the serum and urine by gas chromatography.
     Int.  Arch. Arbeitsmed. 33:49-56.

Ogata, M.  , Y.  Takatsuka,  and  K. Tomokuni.  1971.  Excretion of organic  chlorine
     compounds in the urine of persons exposed to vapours of trichloroethylene
     and tetrachloroethylene.  Brit. J. Ind. Med. 28:386-392.

Orth, O.S., and N.A.  Gillespie.  1945. A further study of trichloroethylene
     anesthesia.  Brit. J. Anaesth. 19:161-173.

Owens, A.M.,  and E.K.  Marshall.  1955.  Further studies  on the metabolic rate  of
     chloral   hydrate  and  trichloroethanol.  Bull.  Johns Hopkins Hosp.  97:
     320-326.

Parchman,  L.G.  and  P.N. Magee.  1982.  Metabolism of 14C-trichloroethylene to
     14C02 and  interaction  of a metabolite with liver DNA in rats and mice.
     J. Toxicol. Environ.  Hlth.  9:797-813.
                                    4-69

-------
Paykoc, Z.V.,  and J.F.  Powell.  1945.  The excretion of sodium trichloroacetate.
     J. Pharmacol.  Exptl.  Therap.  85:289-293.

Pelkonen, 0.,  and H.  Vainio.  1975.  Spectral interactions of a series of chlori-
     nated hydrocarbons with cytochrome P-450 of liver mircosomes from variously
     treated rats.   FEBS Lett.  51:11.

Pessayre, D. , H. Allemand,  J.C.  Wandscheer, V.  Descatoire, J-Y  Artigou,  and
     J-P Benhamou.  1979.  Inhibition, activation, destruction  and induction  of
     drug-metabolizing enzymes by trichloroethylene.  Toxicol. Appl. Pharmacol.
     49:355-363.

Pfaffenberger, C.D.,  A.J.  Peoples and H.F. Enos. 1980. Distribution of volatile
     halogenated organic compounds between rat blood serum and adipose tissue.
     Int. J. Environ. Anal.  Chem.  8:55-65.

Pfaffli, P., and A-L. Blackman. 1972. Trichloroethylene concentration in  blood
     and expired air as  indicators of occupational  exposure.  A preliminary
     report.  Work Environ.  Health 9:140-144.

Poplawski-Tabarelli,  S. ,  and  H.  Uehleke.  1982.  Inhibition of microsomal  drug
     oxidations by aliphatic  halohydrocarbons:   correlation with vapour pres-
     sure.  Xenobiot. 12:53-61.

Powell, J.F. 1945.  Trichloroethylene:  Absorption, elimination and metabolism.
     Brit. J.  Ind.  Med. 2:142-145.

Powell, J.F. 1947.  Solubility or distribution coefficient of  trichloroethylene
     in water, whole blood, and plasma.   Brit. J. Ind. Med. 4:233-236.

Prout,  M.S.;  W.M.  Provan; and T.  Green.  1984.  Species differences in response
     to  trichloroethylene.   I.  Pharmacokinetics in  rats  and  mice.   Toxicol.
     Appl. Pharmacol.  In press.

Reynolds, E.S., and  M.T.  Moslen. 1977. Damage to hepatic  cellular membranes by
     chlorinated olefins  with emphasis on synergism and antagonism.   Environ.
     Health Persp. 21:137-147.

Savolainen, H. ,  P.  Pfaffli,  M.  Tengen, and H.  Vainio.  1977.  Trichloroethylene
     and  1,1,1-trichloroethane:   Effects  on  brain  and  liver after 5  days
     intermittent  inhalation.  Arch. Toxicol. 38:229-237.

Sato,  A.  1979. Movements of  trichloroethylene within the body  in repeated
     exposure:   a  theoretical approach.  Sangyo Igaku (Jap.  J.   Ind. Health)
     21:361-365.

Sato,  A., and  T.  Nakajima.  1978.  Differences  following  skin or inhalation
     exposure  in the absorption and excretion  kinetics of trichloroethylene
     and toluene.  Brit.  J.  Ind. Med.  35:43-49.

Sato,  A., T.  Nakajima, and Y.  Koyama.  1981.  Dose-related effects of a single
     dose of ethanol  on  the metabolism  in  rat liver of some aromatic and
     chlorinated hydrocarbons.  Toxocol.  Appl.  Pharmacol.  60:8-15.
                                     4-70

-------
Sato, A. , T.  Nakajima,  Y.  Fujiwara, and N. Murayama. 1977. A pharmacokinetic
     model  to  study the excretion of  trichloroethylene and its metabolites
     after an inhalation exposure.  Brit. J. Ind. Med.  34:55-63.

Sellers,  E.M.,  and J.  Koch-Weser.  1970.  Potentiation  of Warfarin-induced
     hypoprothrombinemia by  chloral  hydrate.   N. Engl. J.  Med.  283:827-831.

Sellers,  E.M. , G.  Carr,  J.G.  Bernstein,  S.  Sellers,  and J.  Koch-Weser.  1972a.
     Interaction of chloral hydrate and ethanol  in man.  II.  Hemodynamics and
     performance.  Clin. Pharmacol.  Ther. 13:50-58.

Sellers,  E.M.,  M.   Lang, J.  Koch-Weser, E. LeBlanc,  and  H.  Kalant. 1972b.
     Interaction of chloral  hydrate  and  ethanol  in man.   I.  Metabolism.  Clin.
     Pharmacol. Ther.  13:37-49.

Smith,  G.F.  1966.  Trichloroethylene.   A Review.  Brit.  J. Ind.  Med.  23:
     249-262.

Soucek, B.,  and  D.  Vlachova.  1960.  Excretion of  trichloroethylene  metabolites
     in human urine.  Brit. J. Ind.  Med. 17:6064.

Stewart,  R.D., and H.C.  Dodd. 1964. Absorption  of carbon tetrachloride, tri-
     chloroethylene, tetrachloroethylene,  methylene  chloride,  and  1,1,1-tri-
     chloroethane through the human skin.  Am.  Ind. Hyg. Assoc. J.  25:439-446.

Stewart,  R.D., C. L. Hake,  and J.E.  Peterson.  1974. Use of breath  analysis to
     monitor trichloroethylene exposures.  Arch  Environ. Hlth.  29:6-13.

Stewart,  R.D., H.C. Dodd,  H.H.  Gay, and D.S.  Erley.  1970. Experimental  human
     exposure to trichloroehtylene.   Arch. Environ. Hlth.  20:64-71.

Stewart,  R.D.,  H.H. Gay,  D.S.  Erley,  C.L. Hake,  and J.E.  Peterson. 1962.
     Observations  on  the concentrations  of trichloroethylene in  blood  and
     expired air following  exposure of humans.  Am.  Ind.  Hyg.  Assoc.  J. 23:
     167-172.

Stewart,  R. ,  C.  Hake,  and J. Peterson.  1974.  "Degreasers flush":  dermal
     response to trichloroethylene  and ethanol.  Arch.  Environ.  Hlth.  29:1-5.

Stott, W.T.,  J.F.  Quast and  P.G. Watanabe. 1982. Pharmacokinetics  and macro-
     molecular interactions  of  trichloroethylene in mice and rats.  Toxicol.
     Appl. Pharmacol.  62:137-151.

Tabakoff, B. , C.  Vugrincic, R. Anderson, and S.G.A. Alivisatos. 1974. Reduction
     of  chloral  hydrate to  trichloroethanol  in brain  extracts.   Biochem.
     Pharmacol. 23:455-460.

Takano, T., and Y.  Miyazaki.  1982. Effect of chlorinated ethanes and ethylenes
     on  electron transport in rat  liver mitochondria.  J.  Toxicol.  Sci.  7:
     143-149.

Tottmar,  S.O.C.,  H. Petterson, and  K.  H.  Kiessling.  1973. The  subcellular
     distribution  and  properties of aldehyde  dehydrogenase in  rat liver.
     Biochem. J.  135:577-586.
                                    4-71

-------
Traylor, P.S., W. Nastainczyk, and V. Ullrich.  1977.  Conversion  of trichloro-
     ethylene to carbon monoxide by microsomal cytochrome P450-  In:  Microsomes
     and drug  oxidations.   Proceedings of  the  3rd International  Symposium,
     Berlin, 1976.   Suppl.  Biocheml Pharmacol., Pergamon, Oxford,  pp. 615-621.

Tsuruta, H.  1978. Percutaneous absorption of trichloroethylene in  mice.   Ind.
     Hlth.  16:145.

Uehleke, H. , and S.  Poplawski-Tabarel 1 i.  1977. Irreversible  binding of 14C-
     labelled  trichloroethylene  to mice  liver  constituents i_n  vivo  and  jm
     vitro.  Arch.  Toxicol. 37:289-294.

Uehleke, H. , and S.  Poplawski-Tabarel1i,  G.  Bonse,  and D.  Henschler.  1977.
     Spectral evidence for oxirane formation  during microsomal trichloroethylene
     oxidation.  Arch. Toxicol. 37:95-105.

Van Duuren,  B.L.,  and S. Banerjee. 1976.  Covalent interaction of  metabolites
     of the  carcinogen trichloroethylene  in  rat  hepatic microsomes.   Cancer
     Res.  36:2419-2422.

Van Dyke, R. 1977.  Dechlorination mechanisms  of chlorinated olefins.  Environ.
     Health  Persp.  21:121-124.

Van Dyke,  R. ,  and  M.B.  Chenoweth. 1965.  Metabolism of volatile anesthetics.
     Anesth. 26:348.

Vesterberg,  0., J.  Gorczak, and M. Krasts.  1976.  Exposure to  trichloroethylene.
     II. Metabolites  in blood and urine.   Scand.  J.  Work Environ,  and  Health
     4:219-221.

Vignoli, L. , J.  Jouglard,  P. Vignoli,  and  P. Terrasson de Fourgeres.  1970.
     Acute  intoxication by trichloroethylene ingestion.  Med. Leg. Assicur.
     18:789-798.

Vyskocil,  J. ,  and  B.   Polak.  1963. Acute  trichloroethylene  poisoning.   Vnitr.
     Lek.  9:860-863.

White,  J.F. , and G.P.  Carlson. 1981. Epinephrine-induced cardiac  arrhythmias
     in rabbits exposed to  trichloroethylene:   potentiation by   ethanol.
     Toxicol.  Appl. Pharmacol. 60:466-471.

Wiecko, W.  1966.  Oral  poisoning  with  trichloroethylene.  Wiad.   Lek.  19:
     1117-1118.

Withey, J.R.,  and B.T. Collins.  1980. Chlorinated aliphatic hydrocarbons used
     in the foods  industry:   the comparative pharmacokinetics  of methylene
     chloride, 1,2-dichloroethane, chloroform,  and  trichloroethylene  after
     i.v.  administration in the rat.   J.  Environ. Pathol. Toxicol. 3:313-332.

Withey, J.R.,  B.T.  Collins,  and  P.G.  Collins.  1983.  Effect of vehicle on the
     pharmacokinetics and  uptake  of  four  halogenated  hydrocarbons from the
     gastrointestinal tract  of  the  rat.   J. Appl.  Toxicol.  3(5):249-253.
                                     4-72

-------
           5.   TOXICOLOGICAL EFFECTS IN MAN AND EXPERIMENTAL ANIMALS
5.1  INTRODUCTION
     Trichloroethylene (TCI) has  been  an important industrial chemical since
its commercial application  in  1906.   It has been the focus of wide research
interest and the  subject  of numerous case reports.   The available literature
concerning the potential  effects of TCI is reviewed in this chapter.
5.2  NEURAL AND BEHAVIORAL EFFECTS
     Ostlera (1953), in a historical discussion of TCI, cited descriptions of
its anesthetic, analgesic,  and  neurotoxic effects from reports  as  early as
1911.   The  symptoms  described  are remarkably similar to those found in  more
modern literature.   In 1931, Stuber reviewed a total  of 283 cases of presumable
overexposure to TCI,  including  26 fatalities.   Principal  findings  indicated
involved neural and  behavioral effects.   No cases of injury to the  liver were
reported.   The above  findings  were further supported by a study of 384 cases
of industrial  gassing  poisonings  by TCI  and other halocarbons  from 1961 to
1980 (McCarthy and Jones, 1983).   TCI exposure was attributed to 288 reports.
The majority of symptoms  were  attributed to effects on  the central nervous
system (CNS).  However, gastrointestinal  and respiratory  symptoms  also  were
quite common.  The  observation  that TCI  caused paralysis  of  the trigeminal
nerve led  to its early use  in the  treatment of trigeminal  neuralgia.   Boulton
and Sweet  (1960) suggested  that trigeminal palsies that occurred in patients
under TCI  anesthesia may  have  resulted actually from inhalation of dichloro-
acetylene or phosgene, formed when exhaled TCI was passed through soda lime to
remove carbon  dioxide (C0?).  The  formation  of  these  impurities and  their
subsequent toxicity  resulted in a  decline in the  use of TCI as an anesthetic.

5.2.1  Human Studies
5.2.1.1  Short-Term Exposures.
5.2.1.1.1  Case reports.   Numerous case histories have described the effects
of short-term  exposure to high levels  of TCI (Feldman et  al., 1970; Longley
and Jones, 1963; Harenko,  1967; Masoero and Lavarmo,  1955; Vallee and Leclercq,
1935; Kleinfeld and  Tabershaw,  1954; Tomasini and Sartorelli, 1971; Mitchell
                                    5-1

-------
and Parsons-Smith, 1969; Maloof, 1949; Steinberg, 1981; Lawrence and Partyka,
1981).   The earlier studies  reported  on the toxicity of TCI following anes-
thesia, in  particular the occurrence  of trigeminal palsies  (Humphrey  and
McClelland, 1944).   Effects  reported included dizziness,  headache,  nausea,
confusion,  and,  at  very high levels, unconsciousness.  Facial  numbness  and
blurred vision also were  reported  occasionally.   Psychotic symptoms resulted
from ingestion of TCI and alcohol  (Harenko,  1967).
5.2.1.1.2   Experimental exposures.   Stewart  et al.  (1970)  exposed six humans
                     3
to 200 ppm  (1076 mg/m ) TCI for 7  hours.  The subjects reported rapid adaptation
to the odor of  TCI.   Some complaints of  dryness  of the throat and mild eye
irritation occurred after the first 30 minutes of exposure.
     Experimental exposures  to  95 ppm (511 mg/m ) (Konietzko et al., 1975) or
up to  300 ppm  (1644 mg/m ) for 2.5 hours (Ettema et al.,  1975) had no effect
on such behavioral  task performances as simple or choice reaction time, hand
steadiness, tapping tests, or  pursuit tracking.   While Ettema et  al. (1975)
reported  slight  effects,  repeated significance tests  never produced  a p  value
below 0.10  and effects were not dose-related.
     Vernon and Ferguson (1969) exposed eight males repeatedly to 0, 100, 300,
or  1000 ppm (0,  538,  1644,  or 5380 mg/m )  at 2-hr intervals,  with 3 days
separating  sequential  exposures.  Statistical tests showed significant  decre-
ments  in  the performance of groove-type hand steadiness, depth perception, and
                                      3
pegboard  tests  at 1000 ppm (5380 mg/m ) only.  It  is  noteworthy that plots  of
trends of the scores  demonstrated nonsignificant increases in errors at  levels
                               3
lower  than 1000 ppm (5380 mg/m ).   Trend tests or  studies  with  a  larger group
of  subjects perhaps would have been more informative.  No effects were  noted
on Muller-Lyer  illusions,  flicker fusion, and code  substitution tests.
     Ferguson  and Vernon (1970) repeatedly  exposed individuals to 0, 300, or
                                3
1000 ppm  (0,  1644,  or 5380  mg/m )  TCI  for  2 hours.  In one  group of  eight
males,  individuals  were given  also  either meprobamate (800 mg)  or an antihis-
tamine,  thonzylamine  hydrochloride  (50 mg).   In  a second  group of six exposed
to  TCI as  above, an  alcoholic  drink sufficient  to produce a  blood alcohol
level  of  20 to  30 mg/100  ml blood  (2 mg/kg) was  also given.   By itself, TCI
exposure  produced decrements in depth perception, hand  steadiness,  and  pegboard
performance that were statistically significant only  at 1000  ppm  (5380  mg/m  ).
The  two  drugs in addition to  TCI  exposure produced  no effects  beyond  that
 induced by TCI exposure alone.  Alcohol, however, substantially augmented the
                                     5-2

-------
effects of TCI  such  that effects were  significant  at 300 ppm (1644 mg/m ).
Flicker fusion and Muller-Lyer perception were unaffected by any of the condi-
tions.
     Windemuller and Ettema  (1978)  exposed four groups of six males each for
2.5 hours to either (a) room air only, (b) 200 ppm TCI, (c) 1 ml/kg alcohol  in
a mixed drink,  or  (d)  both  alcohol and  TCI.  Performance was measured  in a
pursuit rotor task or  by a binary  choice reaction  time test.  Heart  rate,
sinus arrhythmia, and  respiration  rate also were measured and interpreted as
correlates of "mental load."  The only variable affected was sinus arrhythmia,
which was  significantly reduced only by the  combination  of  200  ppm (1076
    3
mg/m ) TCI and  alcohol.  According  to  the authors,  behavioral performance was
at too high a level  to  show any decrement due  to the  experimental conditions.
A separate group of  15  subjects was studied  in a design in which all partici-
pated under all  conditions.  The same results obtained.
     In a demonstration project, Stopps and  Mclaughlin (1967) showed that TCI
exposure involving one  individual  caused  slight effects.   During a 1.5-hour
exposure,  TCI was administered  sequentially in first  an  ascending  and then
descending order (100,  200, 300, or 500 ppm; 538, 1076, 1644, or 2690  mg/m ).
Of a  battery  of tests, the Necker  cube  reversal  illusion was considerably
affected.   Because only one individual was exposed,  such findings  are difficult
to interpret.   Kyi in  et al. (1967)  purported to show that optokinetic nystagmus
                                                3
in 12 humans was affected  by 1000 ppm  (5380  mg/m )  TCI  for 2  hours.  However,
data were graphically  displayed  only  and thus trends  are  difficult to see.
5.2.1.2  Long-Term Exposure
5.2.1.2.1  Case reports.   Evaluation  of  case reports indicates  that symptoms
involved in  short-term exposure situations  also are  present in  long-term
exposures  but  in more   extreme and  persistent  forms (Harenko, 1967; James,
1962; Kleinfeld  and  Tabershaw,  1954;  Mitchell, 1969;  and  Steinberg,  1981).
Dizziness,  nausea,  and  headache persist even after cessation of  daily exposure.
Longer exposures may produce  ataxia,  decreased appetite,  sleep  disturbances,
and perhaps even psychotic episodes.   Trigeminal  neuropathy  was reported in
two cases (Mitchell, 1969; Feldman et  al. , 1970).   In  some instances recovery
was reported, but in others,  permanent damage seems  to have  occurred.   Data
are  lacking  so  that it is not  possible  to  characterize  the  parameters  of
exposure beyond which  irreversible  damage occurs.   In  some instances, death
occurred (Kleinfeld  and Tabershaw, 1954;  James, 1963).  In four of  six cases,
death occurred  from  1  to 17 hours  after cessation of exposure.   In a case of
                                    5-3

-------
TCI ingestion, death occurred several days after the incident.  Kleinfeld and
Tabershaw (1954) conjectured that  death was due to ventricular fibrillation.
It is possible that in cases where repeated TCI exposures were not sufficient
to produce anesthesia, a  day-to-day  accumulation took place  in body tissues
(see Chapter 4).
5.2.1.2.2  Workplace exposure.   Formal  symptom  surveys of workers exposed to
various levels of  TCI  on  a daily basis revealed that many symptoms of short-
term exposures  became chronic complaints.  Grandjean  et  al.  (1955)  studied
73 workers, Bardodej and  Vyskocic  (1956) studied 75, and Lilis et al.  (1969)
studied  70 workers.   Complaints included  vertigo, fatigue,  headache,  and
nausea.  As  in  case  reports,  tremors,  ataxia,  and impaired vision also were
found.   Reported alcohol  intolerance in TCI-exposed workers  could be due to
day-to-day accumulation of  TCI  in  tissue (see Chapter 4) and the interaction
of alcohol  with TCI.   Bardodej and  Vyskocic  (1956) also reported frequent
cases of addictive behavior in which insomnia occurred unless TCI was sniffed.
In the study of Grandjean et al. (1955), complaint frequency was higher in the
85 ppm (457 mg/m ) exposure group than  in the 14 and 34 ppm (75 and 183 mg/m )
exposure  categories.   Grandjean et  al.  (1955)  also administered tests  for
short-term memory  and  for measuring understanding of  instructions along with
psychiatric  interviews to 73 workers exposed  to average  levels of 14, 34,  and
85 ppm  (75,  183,  and  457 mg/m3),  as part of the  study.  In  the 85 ppm (457
mg/m )  group  there were  higher frequencies of  short-term memory  loss,  fewer
word associations,  increased  perseverance, and increased  rates  of misunder-
standing.
     Workers  have  also been formally  tested  for trigeminal  nerve function.
Barret  et al.  (1982)  studied  11  unexposed and 20 TCI-exposed  (undocumented
levels) workers.   Compared  to the  unexposed group,  8 of the 20 exposed  indivi-
duals  had  elevated thresholds for  trigeminal  nerve evoked potentials.   Triebig
et al.  (1982)  studied trigeminal  motor and sensory nerve  conduction velocity
 in 26  individuals  occupationally exposed  to  an average of  40 ppm (215  mg/m  )
TCI  and in 24 workers having  no TCI exposure.   No differences between  groups
were found,  but  variances were  large and exposure  levels  small.
     Konietzko  et  al.  (1974)  found EEC changes  in  workers  exposed to TCI.   The
 EEC  for  six  workers  was collected  during  work on TCI-exposed days and  on
 nonexposure  days.   During  exposure  days (50  to 100 ppm TCI), there were more
 bursts of alpha waves and the  alpha  waves had higher amplitude than on  unexposed
 days.   No clinically abnormal EEC observations occurred during unexposed work
                                     5-4

-------
or  in  a clinical  setting.   Increased alpha activity  frequently  occurs in
individuals who are in a relaxed state.
5.2.1.2.3  Experimental exposure.   Stewart et al.  (1970) exposed six individuals
to 200 ppm  (1076 mg/m  ) TCI  for 7 hr/day, for 5 sequential days.   Subjective
reports were collected  in addition to blood and physiological measures.  Odor
perception decreased over the  course of repeated exposures as did complaints
of eye  irritation, dryness of throat, and other mild symptoms.  By the  fourth
and fifth  days, however, fatigue and sleepiness during  exposure were reported
by five of the six individuals.
5.2.1.2.4   Summary of  human  neural and  behavioral  effects.   The  effects  of
short-term TCI  exposure range  from  mild eye  irritation,  nausea, vertigo,
headache,  confusion,  unconsciousness,  and  death,  depending  upon exposure
                                                                         3
concentration.   Mild irritation occurs  at levels near  200 ppm (1076 mg/m  ).
Hand steadiness, coordination,  and possibly depth perception are  affected at
1000 ppm (5380 mg/m ) and perhaps  below (see Section 5.2.1).   If combined with
alcohol ingestion, TCI  can  produce  these effects at levels of 200 to 300 ppm
(1076 to 1614 mg/m3).
     Extended exposure  can  increase the  duration  and  intensity of nausea,
vertigo, and headache,  but eye irritation and olfactory sensation  are reduced.
Confusion, reduced cognitive performance,  impaired short-term memory,  sleep
disturbances,  loss of  appetite,  addiction,  alcohol intolerance,  tremors,
ataxia, and  trigeminal  neuropathy also  are reported.   The threshold for such
complaints is difficult to estimate  since such data are gathered from workplace
surveys with all of the attendant problems in quantification  and  control.   It
appears that effects,  however,  are  absent below 85 to  100 ppm (457 to 538
mg/m ).

5.2.2  Laboratory Animals
5.2.2.1  Short-Term Animal Exposures.  Exposure to 30,000  ppm (161,000  mg/m )
TCI for  about  20 minutes  was lethal  to 15 dogs  (Baker,  1958).   After about
5 minutes, dogs began to salivate  heavily and to become ataxic.  Between 5 and
10 minutes, they exhibited  foreleg tonicity and became semicomatose.   During
the last 10 minutes,  dogs were unconscious and frequently exhibited convulsive
(particularly hindlimb)  movements.   Not surprisingly,   no  CMS pathology was
found during autopsy.
     Grandjean (1963)  exposed 29 rats to 400 ppm (2152  mg/m  ) TCI and a group
of 11  to 800 ppm (4304  mg/m  ) TCI, for 6 hours.  Immediately  upon termination
                                    5-5

-------
of exposure, rats were  tested  in a swimming escape  task.   Rats were tested
with and without a 27 gram weight attached to their tails.   In each test,  rats
were given ten  sequential  trials.   At 400 ppm (2152 mg/m ) no effects of  TCI
were observed without the added load and only slight slowing effects were  seen
with the added load.   At 800 ppm (4304 mg/m ) effects were slight and inconsis-
tent without the  added  load; definite and consistent slowing  of  swimming time
occurred with the added load.   During the sequential trials, animals exposed
                      3
to 800 ppm  (4304  mg/m ) TCI under load showed  the greatest effects on the
eighth, ninth,  and especially  the tenth trial.   The author interpreted these
data to mean that TCI exposure produced greater fatigue since the effects  were
dependent upon load and time.
     In the  above report,  Grandjean  (1963) also presented  data  on  jiggle cage
activity levels from 12 rats per group exposed to 0, 400, 800, or 1200 ppm (0,
2152,  4304,  or  6456  mg/m  )  TCI  for  5 hours.   Activity was measured  during
exposure and again for  a period from 20 to 80 minutes after exposure.  Activity
levels  were significantly  decreased  during  and after exposure  to  1200 ppm
(6456  mg/m  ).   Means of the 400  and  800 ppm  (2152  and  4304 mg/m )  groups also
were lower but  not significantly so.
     Grandjean  (1960) exposed rats for 8 hours to  either 200, 600, or 1600 ppm
                         o
(1076,  3228, or 8608 mg/m  )  TCI which were then subsequently  tested for explora-
tory behavior  in  a maze.   Exploratory behavior  was reported to  have decreased
at  200 and 1600 ppm, but  increased at 600 ppm  (3228 mg/m  ).  Since no actual
data were presented,  the information  is difficult  to interpret.
     Baetjer et al.  (1970)  used  hypothalamic electrical  stimulation to reinforce
lever  pressing  behavior in  rats.  Groups  ranging  in size from 14 to 36 were
exposed to either 0, 2500,  or 3500 ppm (0,  13, 450, or  18,830  mg/m  ) TCI  for
30  minutes  on  three  sequential days.  At  each  exposure level,  one group  was
placed on water deprivation for  3 days while the  other  group was  not.  Thus,
there  were six groups of rats.   TCI at both levels reduced the lever pressing
 rate  in a  dose-dependent  fashion.   Especially  at 2,500 ppm  (13,450  mg/m  ),
 response  rate reduction was less noticeable on day 2 than on day 1 and on day
 3 than on  day 2.  Thus,  it appears  that short-term adaptation  can  occur.
Water-deprived  rats  were  affected less and  had delayed effects in comparison
 to  normals.  The authors  speculate  that  in brain cells  of dehydrated rats,
 either TCI levels accumulated more  slowly or the  brain cells were more resistant
 to  TCI effects.  Another  possible  interpretation is that the  interoceptive
 stimuli from water  deprivation  plus  the  stimuli  associated with TCI exposure
                                     5-6

-------
simply increased the activity and, as a result, the  lever pressing  rate, with
respect to rats that were subjected to only one of the above.   Hence, dehydra-
tion may have been unimportant except as another source of stimulation.
5.2.2.2   Repeated  Exposures  (4  days  to  5 weeks).   Savolainen  et  al.  (1977)
recorded open-field activity  in rats after the last of 4 days of exposure to
either 0 or 200 ppm (0 or 1076 mg/m ) TCI  for 6 hr/day.  During the first hour
after exposure, the activity  level  was higher  in  the  exposed group, but by
17 hours postexposure the effect was no longer present.
     Silverman and Williams  (1975)  exposed independent groups of 16 rats to
either 100,  200, 500,  or 1000 ppm (538, 1076,  2690,  or 5380 mg/m3) TCI for
6 hr/day,  5  days/week,  for 5 weeks.   For  each of  the abovementioned  four
groups, a separate contemporaneous group of 16 unexposed controls was studied.
The activity of pairs of male rats was studied in their home cages for 5 minutes
after termination of exposure on days 1, 3, 10, 17, and 24.   Where significant
TCI effects  occurred,  the effect was always a reduction of  activity.   At
                   3
1000 ppm (5380 mg/m ), the effect was evident on the first day.  Other  levels
of exposure  showed  a  gradual  decline in activity with the decline reaching a
                                                                             3
criterion of significance in  a dose-dependent fashion.   The  100 ppm (538 mg/m )
group was  not  affected  significantly at the end of  5  weeks  and was further
exposed  for  a total of  8  weeks.   This  group was eventually  affected  in a
significant manner.
     Goldberg  et al.  (1964)  exposed groups of  rats  to 200,  560, 1568,  and
4380 ppm (1076, 3013, 8436,  and 23,564 mg/m ) TCI  for 4 hr/day,  5 days/week,
for 2 weeks.   Signaled electric shock escape-avoidance behavior was studied in
previously trained  rats.  The highest exposure  level produced  a significantly
greater  rate of response inhibition  than did any of the lower  concentrations.
                                                        3
Concentrations from  200  to  1568  ppm (1076 to  8436 mg/m ) did not differ.
Another group of rats was trained on the same task during exposure.   No signif-
icant differences were found although asymptotic performance was reduced in a
                                             3
dose-related manner above  200 ppm (1076 mg/m ).  The 200 ppm group seemed to
reach the  asymptotic  level faster than controls.   It must be emphasized that
none of  the  acquisition  curves  were significantly different from each  other
and the reported effects  should, therefore, be viewed with caution.
     Battig et al.   (1960) exposed four groups of six rats each to either 0 or
                       3
600 ppm  (0 or  3228 mg/m  ) TCI for 3  to 4 hr/day.  The  daily exposure plan was
complex, with  each  group receiving TCI at  least once  during acquisition and
                                    5-7

-------
extinction of a shuttle box escape-avoidance task.   In no case was acquisition
affected by TCI exposure,  but reaction times were increased and extinction was
slowed.
     Three rats were trained in positively  reinforced, rope-climbing behavior
under the  control  of a positive discriminative  stimulus  (Grandjean,  1960).
They were  then tested  immediately after exposure to either 0,  200, or 800 ppm
(0,  1076,  or 4304 mg/m )  TCI.   Apparently  there  were 7 sequential days at
                  3                                                        3
200 ppm (1076 mg/m ) for 3  hr/day, followed by 5 days of 800  ppm  (4304 mg/m )
for  3 hours.   Controls were interspersed.   It is not clear  when the  data
presented  were collected;  they  are  probably a summary over all days.  There
was  no  effect  on  latency  or failures to respond, but the number of responses
in  the  absence of a discriminitive  stimulus was  higher in the TCI-exposed
groups.   There was  no  difference  between the  200 and 800 ppm (1076 and 4304
    3
mg/m ) groups.   Little can be concluded from only three rats.
     Goldberg et al. (1964) reported data on the performance of electric shock
                                                                   3
escape-avoidance responses  in  10  rats exposed  to 125 ppm (672 mg/m ) TCI for
4  hr/day,  5  days/week,  for a total  of 14 days.   Rats were tested during  the
last hour  of exposure.  At the onset  of  exposure,  rats  immediately began  to
receive more electric  shocks (avoidance response failures) and continued  to
perform more poorly than  on pre-exposure days throughout  the 14  days.  Some
evidence of  adaptation to TCI was observed.   After the 14-day  exposure period,
rats were  tested  for 7 more unexposed days  during which performance returned
to and  below pre-exposure levels.   The evidence for adaptation is poor since
(a)  no  contemporaneous controls were  studied and (b) the "adaptation" trend of
improvement  continued  for  the 7 postexposure days.
5.2.2.3   Long-Term  Exposures.   Battig and  Grandjean  (1963) exposed groups of
10 rats to either 0  or 400  ppm  (0 or  2152 mg/m3) TCI  for 8 hr/day, 5 days/week,
for  44  weeks.   Several behaviors  were tested  either  immediately  after daily
exposure or  16 hours after exposure.   TCI increased times  in  a swimming escape-
avoidance  test.   The magnitude of the effect  became  greater  as the number of
weeks  of  exposure increased.  When  tested  16  hours after  exposure, no effect
was  seen  on the swimming  test.   Acquisition of shuttle box avoidance behavior
was  not affected either immediately  after  exposure or 16 hours later.   Hebb
maze performance  was not  affected when tested  16 hours after exposure.   Tests
of exploratory behavior which were conducted immediately  after exposure during
the last  2 weeks  of the study  showed an  increased  level  of activity.   This may
be due to the lack of adaptation  to novel surroundings  (Daschiell maze).
                                     5-8

-------
     Kjellstrand  et  al.  (1980) studied  the effects  of continuous 320 ppm
(1721 mg/m ) exposure for 9 months on radial-arm maze performance in 24 gerbils.
An  unexposed  control  group of 24 also was tested.  Gerbils were removed from
exposure chambers  for the  test.   An  eight-arm  maze  test was  started  on  day  59
and terminated  on  day 87.   Performance was  nearly perfect  in both  groups.   At
day 237, a  16-arm maze  test was  begun with  16 gerbils  per group.  Again, no
significant differences were  observed even  though performance was  not as good
as  in  the  eight-arm  maze.   Eleven days  after  the termination of the 9-month
exposure, 32 gerbils  were  split into four groups:   two  groups of 8 previously
exposed to 9 months of TCI and two groups of 8 previously unexposed.   All four
groups were exposed  on  alternate  days to 2,300 ppm 1,1,1-trichloroethane for
6 hours and then tested in the 16-arm maze.   Those gerbils that had previously
been exposed  to TCI  showed worse performance after 1,1,1-trichloroethane
exposure than after  air exposure.  Similar  but somewhat less  dramatic results
were obtained from tests 75 days  after TCI  exposure.  The  authors  interpreted
these findings  to  mean  that permanent damage  to  the  CMS resulted during the
9 months of  exposure to TCI.   This  interpretation  is difficult to support
since exposure  to  1,1,1-trichloroethane  improved performance in gerbils pre-
viously exposed to TCI.   It is more  likely that the findings  can be explained
in terms of state dependent learning.
     Baker (1958)  exposed  dogs repeatedly to  concentrations  of TCI  ranging
from 500 to 3000  ppm (2690 to 16,140 mg/m3) for  2  to 8 hours.  He reported
anecdotally that animals exhibited such  chronic symptoms as  ataxia, excessive
salivation, head tremor, teeth gnashing, and extension of extremities.   Upon
repeated exposure, dogs progressively became unconscious more quickly.  Post-
mortem inspection  of  brain  tissue in three  dogs that had survived between  60
and 162 hours  of TCI  (unspecified  concentration) revealed selective destruction
of cells in the Purkinje  layer of the cerebellum and,  to a  lesser extent,
unselective cellular  damage to the cerebral  cortex.   Changes  to Purkinje cells
in rabbits injected  intramuscularly  with a total  of 53  g of TCI were reported
by Bartonicek and Brun (1970).
     Continuous  exposure of gerbils  to  0,  60,  or 320 ppm  (0, 323, or 1722
mg/m )  TCI for  3 months was followed by  a recovery period of  4 months (Haglid
et al., 1981).  Gerbils were sacrificed  (6 to  7 per group) and the percentage
of S100 protein  and DNA  in  various brain  regions was determined.   Although  the
authors purport to find a  consistent  change  in  S100 protein percentages,
                                    5-9

-------
inspection of the  data  reveals  changes which are  not  dose-related and are
inconsistent across  areas.   DNA seems  to be increased in  the  cerebellum,
sensory cortex,  and motor cortex,  but not in brainstem  or  hippocampus.   However,
if one makes Bonferroni corrections to the significance tests, no significant
increases remain.   This  report should be viewed  as  presenting pilot data only.
5.2.2.4     Summary of Neural  and  Behavioral  Effects in Laboratory Animals.
Effects of TCI  upon  behavior  have been demonstrated at exposures as low as
400 ppm (2154 mg/m )for  6 hours or  as low as 200 ppm  (1076 mg/m  ) for four,
6-hr days.  If  exposure  is  extended to 6 hr/day for up  to 8 weeks,  100 ppm
(538 mg/m ) will produce behavioral effects.  These effects include  increased
fatigue, decreased activity  level, and slowed extinction  of learned behaviors.
     These behavioral  changes might  well  be regarded as  analogous  to the
effects of  TCI  on humans.  Fatigue  is a common complaint  at  low exposure
levels.   Activity  level  reduction  is another possible analog  of fatigue.
Slowed extinction  of  a learned response might be related to short-term memory
loss.  In general,  the  laboratory animal as an analog of human behavior is
speculative.
     At very high repeated exposure levels, TCI  apparently selectively destroys
the Purkinje cell layer in the cerebellum.  Data regarding the neuropathological
effects of  lower  level exposures  for  long periods  are  inconclusive or  absent.
Neuropathological  investigations  could provide a model of  TCI-induced  ataxia.
5.3  EFFECTS ON THE CARDIOVASCULAR AND RESPIRATORY SYSTEMS
5.3.1  Cardiovascular Effects
     The use of TCI as an anesthetic has been associated with cardiac arrhyth-
mias,  including  bradycardia,  atrial  and ventricular premature  contractions,
and  ventricular  extrasystoles.   The  dose-response relationships for  these
effects, however,  has  not been established  in  man or experimental  animals.
High concentrations  of  TCI have been reported to  sensitize the myocardium to
circulatory catecholamines, resulting in cardiac arrhythmias including ventri-
cular  fibrillation or even cardiac arrest (Dhuner  et al., 1951, 1957;  Reinhardt
et al., 1973).  Metabolites of TCI, such as  chloral hydrate, may also  be asso-
ciated with  arrhythmias involving adrenergic stimulation (DiGiovanni, 1969).
Lilis  et al.  (1969)  observed in their  study that 60 percent of the workers
exposed  to inhaled TCI exhibited  an  increase in  cardiac output, an  effect
                                     5-10

-------
attributed  to  increases  of epinephrine  release and circulating plasma levels.
These  direct cardiac effects of TCI, with  sensitization to catecholamines,
have been  observed after accidental  ingestion  and have,  in  some cases,  proved
fatal  (Defalque, 1961).
     Several case  histories have been reported involving fatalities  following
acute  exposure to  large concentrations of TCI.   The deaths in these cases were
attributed  to  either cardiac arrest  (James,  1963) or  ventricular fibrillation
(Kleinfield  and  Tabershaw,  1954).  Bell  (1951),  Bardodej and Vyskocil (1956),
Ogata  et al.  (1971),  Andersson  (1957),  and  others have noted that exposure  to
TCI may either speed up or  slow down the heart rate, depending on the degree
of  exposure.   The most direct  evidence  that TCI causes  ventricular  extra-
systoles,  fibrillation, and  cardiac arrest comes from changes that are observed
in  the electrocardiograms  (EKG) of  subjects  accidentally  exposed to high
concentrations  of  TCI (Kledecki and Bura,  1963;  Mroczek and  Fedyk, 1971;
Yacoub etal.,  1973).  Bernstine  (1954) recorded ventricular  fibrillation
confirmed by EKG in  a worker who sustained  cardiac arrest after inhalation  of
TCI in analgesic concentrations.   Barnes and Ives (1944) found that 33 of 40
normal  patients  in deep  anesthetic planes showed signs of ventricular extra-
systoles and multifocal ventricular tachycardia.
     Few systematic animal studies  have  been done on the effects of TCI on the
cardiovascular  system.   Mazza and  Brancaccio  (1967)  studied the effects on
rabbits of chronic exposures to 2790  ppm (15,010 mg/m ) TCI, 4 hr/day, 6 days/
week,  for 45 days.   They found that severe  blood dyscrasia was produced in
these  animals.  The characteristics of  the hemochromocytometric tests and the
description of the bone marrow led  the  authors  to conclude that chronic intoxi-
cation with TCI has a direct effect on  the bone marrow, which causes myelotoxic
anemia.  Mikiskova and Mikiska  (1966)  studied  electrophysiological responses
of guinea pigs  injected  i.p.  with either with  6.7 mM/kg  TCI or  with 2.2 mM/kg
trichloroethanol.  They recorded EKG's  and EEG's and found that trichloroethanol
was at least three times as effective as TCI in inducing effects on the nervous
system and  in  slowing the heart, and suggested that TCI  effects were due, in
part, to its metabolite,  trichloroethanol.
     White and Carlson (1979) demonstrated that epinephrine-induced arrhythmias
in rabbits exposed to TCI could be  decreased by phenobarbital  treatment.   This
increases  metabolism  of TCI.   This would support the view that TCI,  and  not
its metabolites, is  responsible for  arrhythmias.   Caffeine  has been shown to
                                    5-11

-------
potentiate the arrhythmogenicity of TCI in rabbits upon challenge with exogen-
ously administered epinephrine (White and Carlson, 1982).   Rabbits were treated
with a vehicle control (1 mg/kg, i.p.) or caffeine (10 mg/kg, i.p. 30 minutes
prior to exposure) and exposed for 1 hour to 6000 ppm (32,280 mg/m )  TCI under
dynamic airflow conditions.   Epinephrine was infused until  arrhythmias occurred
after 7.5, 15, 30,  45 and 60 minutes of exposure and 15 and 30 minutes post-
exposure.  The  authors reported a pronounced  increase  in  the incidence of
epinephrine-induced  arrhythmias  in  TCI-exposed  rabbits when treated with
caffeine and challenged with doses of epinephrine as low as 0.5 ug/kg.
     White and Carlson (1981) also investigated  the  effect  of ethanol on epi-
nephrine-induced cardiac arrhythmias in rabbits exposed to 6000 ppm (32,280 mg/
m ) TCI.   Rabbits were treated with  saline  (1  ml/kg) or ethanol  (1 g/kg, i.v.
or p.o.) 30 minutes prior to exposure to 6000 ppm TCI for 1 hour.  Epinephrine
was infused as described above.   Rabbits treated with ethanol developed epine-
phrine-induced cardiac arrhythmias  sooner and at lower doses of epinephrine
than control  rabbits.  More  arrhythmias  developed  in  rabbits treated  with
ethanol i.v.  than p.o.  Ethanol  administration decreased the conversion of TCI
to  trichloroethanol  by 50 percent,  and  trichloroacetic  acid production was
completely blocked by ethanol treatment.
     Most recently, Carlson and White (1984) evaluated the effect of benzo(a)-
pyrene on  the  production of arrhythmias  in  epinephrine-induced,  TCI-exposed
rabbits.   Benzo(a)pyrene  increased  TCI metabolism but also  led  to increased
arrhythmias.   Thus,  it would appear that benzo(a)pyrene has intrinsic  arrhy-
                                                                 3
thmogenic  action.   Rabbits were  exposed  to  8100  ppm  (43,578 mg/m ) TCI  48  and
72 hours subsequent to administration of 40 mg benzo(a)pyrene/kg.
     Fossa et al. (1982) demonstrated that disulfiram exhibits a  cardioprotec-
tive activity not related to its effects on TCI metabolism  in rabbits.  Disul-
firam  (1.35  mmoles/kg,  p.o.) was administered 24 and 6 hours prior to  a 1-hr
exposure  to  6000 ppm (32,280 mg/m ) TCI.  Disulfiram or its metabolites, C$2
and diethyldethiocarbamate (1.35 mmoles/kg,  i.p.), were administered  at similar
                                                                           3
time intervals  also to rabbits exposed  for  1  hour to 9000  ppm (48,420  mg/m )
TCI.  Blood  levels  of TCI and its metabolites, trichloroethanol  and trichloro-
acetic  acid,  were measured  at periods  ranging from  7.5 minutes  of exposure to
30 minutes post-exposure.  The metabolism of TCI was inhibited in all treatment
groups.   When  challenged with 0.5 to  3.0 ug/kg  epinephrine, disulfiram pre-
vented the epinephrine-induced arrhythmias  resulting from exposure to 6000 ppm
                                     5-12

-------
(32,280 mg/m ) TCI.  All treatment groups  in the first 30 minutes of exposure
                         3
to 9000 ppm  (48,420  mg/m )  TCI exhibited  an increased percentage of animals
responding with arrhythmias compared to controls.  Thirty to sixty minutes of
exposure  resulted  in a  significant  decrease in the  percentage  of  animals
responding with  arrhythmias in the disulfiram-  and  diethyldithiocarbamate-
treated groups.
5.3.2  Respiratory Effects
     In anesthetic concentrations, TCI  causes  little or no irritation to the
respiratory tract  (Dobkin and  Byles,  1963).   A striking effect  is observed,
however, on the rate and depth of respiration.  TCI characteristically causes
increased  respiratory  rate  (tachypnea) but  decreased alveolar  ventilatory
amplitude, which is  associated with  decreased blood  oxygen tension  and  in-
creased carbon dioxide  tension (Dobkin  and Byles,  1963).   These effects  were
noted  in  early animal  experiments (Lehmann, 1911) and  in  early reports on
human patients under TCI anesthesia (Striker et al.,  1935).  The incidence of
tachypnea  increases as  the  concentration  of TCI increases (Atkinson, 1960).
     To investigate the mechanism responsible for the increase in respiratory
rate, Whitteridge  and Bui bring (1944) studied the effects of TCI on pulmonary
afferent nerve endings  in cats.  They found that 0.5  to 3.5 percent (5,000 to
                                  3
35,000 ppm; 26,900 to 188,300 mg/m )  TCI produced a dose-dependent increase in
the frequency  of vagal  discharge  and an increased  sensitivity of the stretch
endings.  Coleridge et al.  (1968)  observed similar  results  in  isolated pulmonary
receptor endings.
     Coleridge et al. (1968) attempted  to correlate  injury to rat pulmonary
alveolar parenchyma with biochemical  alterations following  repeated  inhalation
of TCI.  Mild  changes  in morphology  were  observed,  including  condensed mito-
chondria and  vacuolation  of the  cells.    Pulmonary surfactant  secretion  was
significantly  decreased.  The  authors  suggested that modification  of sur-
factant secretion was a consequence  of nonspecific injury  following exposure
of a susceptible cell population  to TCI.
     Lewis et al.  (1984) attempted to determine if  TCI alone (without previous
induction) can affect the activity of pulmonary mixed-function oxidase in mice
and if  inhalation  of  this  compound can injure the lungs  directly.   Animals
were exposed  to  10,000 ppm TCI (53,800  mg/m ) for 5 days, 1  or 4  hr/day.
Pathologic changes such as thrombus formation and vacuolization of bronchiolar
epithelial cells were also noted in the animals with  reduced microsomal NADPH
                                    5-13

-------
cytochrome c reductase.  The authors concluded that the decrease in pulmonary
enzyme activity may reflect direct damage to the lungs.
5.4  HEPATIC AND RENAL TOXICITY
5.4.1  Human Studies
     The similarity between TCI  and chloroform and other related halocarbons
known to cause  liver damage suggests that TCI might possess hepatotoxic acti-
vity.  Stuber (1931) reviewed 284 cases of industrial  poisoning due to TCI and
found no evidence of liver damage.  Brittain (1948) found no evidence of liver
or  kidney  damage in 250  neurosurgery  patients  who underwent prolonged TCI
anesthesia.  However,  fatal acute  hepatic failure  has been observed following
the  use  of TCI  as  an anesthetic.   Such  failure  has generally been  in  patients
with other complications such as malnutrition, toxemias and burns, or patients
who  had  received transfusions  (Ayre, 1945;  Berleb, 1953; Dodds,  1945;  El am,
1949; Heizer, 1951; and Werch et al. , 1955).
     Secchi et al. (1968) reported  acute liver disease in three of seven cases
of  poisoning  following ingestion of TCI.   They  suggested,  however, that  the
results  could be  attributed  to contamination with 1,2-dichloropropane and
1,2-dichloroethane since no liver  damage was associated with ingestion of pure
TCI.  Cotter  (1960) described 10 patients acutely exposed to TCI in a confined
space.   Four patients showed  hyperglobulinemia  and  six exhibited hypercal-
cemia.   No other liver tests were  abnormal  and,  in  all but one person, the
clinical  tests  returned  to normal  within two months  of  exposure.   Lachnit and
Brichta  (1958)  reported that of  22  workers  occupationally exposed  to  TCI, only
three  had  abnormal reactions  on two liver function tests (sulfobromophthalein
clearance  tests and  colloid  stability).  Albahary et al.  (1959)  conducted
liver  function  tests—measuring  serum  glutamic  oxaloacetic transaminase (SCOT)
and serum glutamic pyruvic transaminase  (SGPT)--on workers regularly exposed
to  TCI.    No  evidence  of  liver  disorders was  found.   Guyotjeannin and Van
Steenkiste (1958) found  some  abnormalities in cephalic flocculation,  total
lipids,  and plasma unsaturated  fatty  acids,  and an  increase of plasma gamma
globulins  in 18 workers regularly  exposed  to  TCI.  Tolot et al.  (1964) found
no  evidence of  liver  injury,  even though workers were  routinely  exposed to TCI
at  concentrations higher than the allowable tolerance limits,  that is, 0.9 to
3 mg of TCI per liter of air (167 to  558 ppm).  Joron  et al.  (1955)  reported
massive  liver necrosis (particularly  the left lobe)  followed by  death of a
                                     5-14

-------
chemist exposed intermittently to TCI over a 1.5-year period.   Subsequent ana-
lyses of room air indicated TCI concentrations as high as 300 ppm (1614 mg/m ).
     Phoon et al. (1984) reported recently on five individuals exposed to TCI
occupationally who had liver damage in conjunction with severe erythema (Stevens-
Johnson syndrome).  It is unlikely that TCI was  a causative factor in producing
the liver injury  since  other workers were unaffected and since the exposure
                                                                  3
levels were estimated to be between 9 and 169 ppm (50 and 912 mg/m ).
     In the case  reports  of occupational exposure to TCI described by Suciu
and Olinici  (1983),  factors other than TCI probably were more causally related
to the hepatorenal disturbances noted.
     From the  above  reports, it is apparent  that  observations  of liver or
renal dysfunction in workers exposed to TCI have been infrequent,  and relatively
little surveillance has been conducted to ascertain  latent effects.  Thus, it
is unlikely that  the potential magnitude of any  kidney or liver abnormalities
in humans exposed for  long periods to TCI  is  fully  known.   Cases of severe
liver and  kidney damage  resulting  from  ingestion of TCI may  provide  some
evidence for latent toxicity.  However, it is difficult to predict whether the
same effects would be observed if ingestion of TCI was spread over a period of
time similar to that found in inhalation exposure.
     TCI has also been reported to have some abuse potential.   Cases of delib-
erate and  repetitive  inhalation have been reported  by  James  (1963) and by
Ikeda and Imamura (1973).   Baerg and Kimberg (1970)  documented centrilobular
hepatic necrosis and acute hepatic and renal damage following repeated sniffing
of cleaning  fluid containing  TCI  and petroleum  distillates  by  teenagers.
However, as the  vapor  concentration was high and  inhalation  was  only for a
short period, the effect cannot  be  directly extrapolated to inhalation of the
same  total  dose  over a  long period of time, as  in  occupational  exposure.

5.4.2  Animal  Studies
     Hepatotoxic  potential  has  been evaluated in a  number of animal species
under various exposure  conditions  and schedules.  Findings are tabulated in
Table 5-1.
5.4.2.1  Inhalation Studies—In  a  series  of experiments, Kjellstrand et al.
(1981; 1983a,b) evaluated the effects of TCI exposure on body and organ weights,
and plasma butyrylcholinesterase in Sprague-Dawley rats, various  mouse strains,
and mongolian gerbils.   Kjellstrand et al.  (1981) found suggestive evidence
                                    5-15

-------
that rats, NMRI mice,  and  gerbils (consisting of pooled  males  and females)
showed relative and absolute  liver  weight gain after continuous exposure to
                 3
150 ppm (807 mg/m ) for  up to 30 days.   The  effect  seemed to be more pro-
nounced in  mice.   During  postexposure,  relative liver weights  decreased.
Study design limitations hampered evaluation  of the  statistical  significance
of results;  body weight gains  were determined  by linear regression.
     In a  follow-up  experiment,  Kjellstrand  et al.   (1983a)  obtained  more
convincing evidence that exposure to 150 ppm  (807 mg/m )  for 30 days caused
liver weight gain in  various mouse strains (wild, C57BL,  DBA,  B6CBA, A/sn,  and
NZB).  Plasma butyrylcholinesterase, a liver-produced enzyme of unknown func-
tion, also showed statistically significant increased activity in males of  all
strains and in females of strains A/sn and NZB.
     In a  further  extension  of their studies,  Kjellstrand et  al.  (1983b)
exposed male and female NMRI mice continuously (up to 120 days)  to levels from
37 to  300 ppm (200 to  1614 mg/m3) and  intermittently to 225 to 3600 ppm (1233
              3
to 19,728 mg/m  ).  Regardless of the exposure schedule,  enzyme activity was
highest in male mice under all  exposure  schedules.  Peak  activity was reached
about day 30.   Activity in female mice was largely unaffected by TCI exposure.
Liver weights in both  sexes increased  under either continuous or intermittent
exposure.   Peak increase  in liver weight during 150  ppm exposure occurred at
about day 30.
     Light microscopy revealed distinct changes in liver morphology at exposure
levels as  low as  37  ppm  (200  mg/m3).   Rehabilitation  during postexposure
revealed that the liver became histologically similar to air-exposed controls.
TCI exposure had little effect on body weight and spleen weight.   While kidney
weight increased during  exposure,  weight gain decreased  during postexposure
rehabilitation.   It was concluded that TCI in this species induced a transient
abnormal  state  but  no liver  damage.   It  was  noted  that Shimada and Yoshida
(1980) proposed that elevations  in butyrylcholinesterase activity may represent
fatty metamorphosis or toxic  damage  to the liver.  The finding that butyrylcho-
linesterase activity is largely  confined  to male mice and  that tumor induction
appears  similarly  sex-specific  to  males (NCI, 1976) suggests that  further
investigation of  the  role of butyrylcholinesterase  in metabolism  warrants
attention.
     Other inhalation  studies  (see Table  5-1)  have been more limited in scope,
and  findings  were  generally unremarkable (Gehring, 1958;  Newell  et al. ,  1954.;
                                    5-16

-------
                                           TABLE  5-1.   EFFECTS OF  TCI  EXPOSURE  ON  EXPERIMENTAL  ANIMALS
Mode of Exposure
    Exposure
      Dose
      Schedule
                                                                                      Effects
                                                                                                                         Reference
 I
I—»
-^1
Inhalation
  Mouse (NMRI)
  Mongolian
  gerbil
  Rat (S-D)
  Mouse (wild,
  C57BL,  DBA,
  B6CBA,  A/sn,  NZB,
  NMRI)
  Mouse (NMRI)
150 ppn




150 ppm





150 ppm



150 ppm
37, 75, 150, and
  300 ppm
Continuous
up to 30 days
for most groups
Continuous
for 30 days
Continuous
for 30 days
Continuous
for 30 days
Continuous
for 30 days
(also for 120 days
at 150 ppm)
Analysis of b.w. gain limited in mice
by poor study design.  Relative liver
weight increased and then decreased
during rehabilitation.

No strong evidence for b.w. gain or
loss.  Small but significant relative
liver weight Increases.  Kidney weight
increases significant for males and
females.

Only increased b.w. for some female
rats.  Small but significant in-
creases in relative liver weight.

All strains showed large increases
in liver weight.  Small changes in
kidney and spleen weight.  Plasma
butyrylcholinesterase activity in-
creased in males of all strains.

Males exposed to 300 ppm had 3.5
increase in butyrylcholinesterase;
females had slight changes at any
level.   Relative liver weight was
linear function of concentration
ranged.  Increase dropped when body
weight dropped.   Body weight gain was
less than controls.  Kidney weight in-
creased more in males (significant at
75 ppm).   Spleen weight unchanged.
Butyrylcholinesterase activity peaked
at day 30; returned to normal  during
postexposure.   Liver weight increase
also peaked near day 30.   Liver weight
returned to near normal during post-
exposure.
                                                                                                                    Kjell strand et  al.,  1981
Kjellstrand et al., 1983a
                                                                                                                   Kjellstrand et al., 1983b

-------
                                                                TABLE 5-1.   (Continued)
    Mode of Exposure
                                Exposure
                                  Dose
                               Schedule
                                                              Effects
                                                                                                 Reference
      Mouse (NMRI)
225, 450, 900,
1800, and 3600
ppm
 I
I—'
oo
      Mouse
5500 ppm
      Dog, rabbit,  rat      500 to 1000 ppm
      Rat
15,000 ppm
      Rat, guinea pig,      35 and 700 ppm
      dog, rabbit, monkey
  30 days
(hours/day
    x
 days/week)
  16x7
   8x7
   4x7
   2x7
   1x7
respectively
                                                     Acute
                         18 hr/day, 90 days
                                                     Intermittent
                         8 hr/day, 5 days/
                         week for 30 exposures
                         over a 6-week period.
After 30 days at 150 ppm, liver cells
larger and had fine vacuolization of
cytoplasm.  Kuppfer cells increased in
cellular and nuclear size; connective
tissue infiltrated with inflammatory
cells.  No bile stasis.  During
postexposure, morphology similar to
controls.

Butyrylcholinesterase significantly
increased in males by average expo-
sure of 150 ppm irrespective of
pulse.  Females unaffected.  Liver,
body, and kidney weights increase un-
affected by pulse.  Motor activity
increased at 3600 ppm but decreased
at 900 ppm.  Liver morphological dif-
ferences more pronounced than with
continuous exposure.  Mitoses more
frequent in groups exposed longer
intervals.

No increase in SGPT; histopathology
not performed.

No gross changes in liver, renal,
or blood functions.

No liver or kidney damage.
Histopathology not performed.

No signs of hepatotoxicity.  No
mortality.  15 rats and guinea
pigs, 3 rabbits and monkeys, and
2 dogs exposed at each dose.
No changes in liver enzymes.
                                                                                                                        Kjellstrand  et  al.,  1983b
                                                                                                                        Gehring,  1958
                                                                                                                        Nowill  et al.,  1954
                                                                                                                        Utesch et al.,  1981
                                                                                                                        Pendergast et al.,  1967

-------
                                                                     TABLE 5-1.   (Continued)
        Mode of Exposure
    Exposure
      Dose
                                                               Schedule
                                                                                              Effects
                                                                                                                                 Reference
en
t—*
10
          Albino rat
          Dog
          Rat
        Gavage
          Male mice
          Rat
          Swiss mice
          (mixed sexes)
Anesthetic





750 ppm


500 to 600 ppm


1% TCI
                                0, 250, 500,
                                1200, or 2400
                                mg/kg
                                0 and 1100 mg/kg
360 to 4300
mg/kg
                                                         2 hr
8 hr/day, 6 days/week
3 weeks

6 hr/day, 5 days/week
8 weeks

2 hr
                         5 days/week
                         3 weeks
Ix
 No  signs  of  parenchymal  liver  damage.
 Distribution of  acid phosphatase-
 positive  granules  irregular and
 diminished in number, 24 hr after
 exposure.

 Glycogen  depletion and hydropic paren-
 chymatous liver  degeneration observed.

 Test mixture may have contained con-
 taminants.

 Glutathione  levels decreased during
 TCI exposure of  phenobarbital-treated
 rats.  Rebounded during  postexposure.
Dose-related changes characteristic
of hepatocellular hypertrophy at all
dose  levels.  Slight changes at lowest
dose,  increased  liver weight at 500
'Ag/kg, severe centrilobular hepatocyte
swelling, giant  cell inflammation,
and mineralized  cells at highest dose.
No effect on kidney or body weight.

Slightly increased liver weight but
no histopathological changes.  No
effect on kidney or body weight.

Fatty  infiltration.  Minimum lethal
dose:  2920 mg/kg.  Minimum hepato-
toxic  dose: 720  mg/kg.
                                                                                                                            Ramadan and Ramadan,  1969
                                                                                                                            Seifter, 1944
                                                                                                                            Moslen et al.,  1977
                                                                   Stott et al., 1982
                                                                                                                            Stott et al.,  1982
                                                                   Jones et al., 1958

-------
                                                                       TABLE 5-1.   (Continued)
Mode of Exposure
Rat
Weanling CD-I
ICR outbred albino
mice
(mixed sexes)
Exposure
Dose Schedule
85 ing/kg and ix
340 mg/kg
24 and 240 rag/kg 14 days
5 days
Effects
No effect on liver weight or trigly-
ceHdes. Effects only if pretreated
with phenobarbital.
Apparent dose-related increase in
liver weight.
Aniline hydroxylase significantly
Reference
Danni et al . , 1981
Tucker et al . , 1982
ro
o
         Subcutaneous
           Rat
        Drinking Water
nnklng
 Weanli
               ing CD-I
           ICR outbred albino
           mice (mixed sexes)
                                730 mg/kg  (day 0)
                                1460 mg/kg (day 1)
                                2910 mg/kg (day 3)
                                1460 mg/kg (days 4,
                                sacrifice  (day 6)
                       500 mg/kg
0.1, 1.0, 2.5, and
5.0 mg/ml in IX
Emulphor
                                           5)
                             2x
                             2x
                             2x
                             Ix
                                                         Ix
4 and 6 mo
                         elevated and aminopyrene N-demethylase
                         significantly decreased.  Oral LD50 in
                         females: 1839 to 3779 mg/kg; in males:
                         2065 to 2771 mg/kg.
                                                  SCOT elevated within 12 to 16 hr.
                                                  Correlated with hepatocellular
                                                  damage seen in histopathology.
                                                                         Decreased water  consumption  in  females
                                                                         at  highest  dose  and  in males at two
                                                                         highest  doses.   Gross pathology un-
                                                                         remarkable.   Significant  increase  in
                                                                         kidney weight (males) at  highest dose
                                                                         at  6  mo  and in females at 4  and 6  mo.
                                                                         Elevated protein and ketone  levels in
                                                                         urine in high-dose females (6 mo)  and
                                                                         two highest doses for males  (6  mo).
                                                                         Decreased RBC count  in high-dose males
                                                                         (4  and 6 mo),  decreased leukocyte
                                                                         counts,  altered  coagulation  values in
                                                                         males (4 and 6 mo),  and shortened  pro-
                                                                         thrombin time (females, 6 mo).
                                                                   Wirtschafter and
                                                                   Cronyn, 1964
                                                                   Tucker et al., 1982

-------
Utesch, 1981; Pendergast et al., 1967; Ramadan and Ramadan, 1969).  In many of
these studies, histopathology was not performed.
5.4.2.2 Gavage—Stott  et al.  (1981)  found  dose-related  changes  characteristic
of hepatocellular hypertrophy at all dosage levels in mice.  This study included
dosages comparable  to  those used  in  the  National  Cancer Institute (NCI, 1976)
bioassay for carcinogenicity of TCI by gavage in B6C3F1 mice and Osborne-Mendel
rats (see Chapter 7).  In the Stott et al.  study, male mice were given 0,  250,
500, 1200,  or 2400  mg TCI/kg body  weight/day,  5 days/week,  for 3 weeks.
The rats  received 0  or 1100 mg TCI/body  weight/day.   Rats  did not show histo-
pathological changes but had  slightly increased  liver  weights.   The  authors
suggest that the  greater  sensitivity of mice reflects their greater capacity
for metabolic activation.  The  species difference could very well be due to
metabolic saturation in the rat (chapter 4).
     Relative liver  weight and  hepatic  triglyceride content were  used  as
parameters  in measuring liver injury in  the study by  Danni et al.  (1981).  No
injury in Wistar  rats  at  dose levels  of 85 mg/kg and 340  mg/kg was observed
unless the  animals were first pretreated with phenobarbital.  It  was  noted by
Moslen et al. (1977) and  by Adams et al. (1951)  that acute liver necrosis is
produced only when  TCI inhalation is preceded by pretreatment with phenobar-
bital.   In  phenobarbital-treated  rats,  hepatic  glutathione levels decreased
during TCI  exposure  but  rebounded during the postexposure period (Moslen et
al., 1977).
     In a 14-day  gavage  study with weanling CD-I and ICR outbred albino  male
and female  mice,  the only  effect noted was an apparent dose-related increase
in liver weight.  Doses given were 24 and 240 mg/kg (Tucker et al.,  1982).
5.4.2.3  Drinking Water—Tucker et al. (1982) evaluated the toxicity of TCI to
male and female mice administered TCI in a 1 percent solution of Emulphor.
Results of gross pathology  on mice sacrificed at  4 and 6 months were unremarkable.
Dose levels were 0.1, 1.0,  2.5,  and 5.0 mg  TCI/ml.  Decreased water consumption
was evident in  females at  the highest dose  level and in males at  the  two
highest dose levels, compared to vehicle controls.  Urinalysis and hematological
findings are indicated in  Table  5-1.
5.5  IMMUNOLOGICAL AND HEMATOLOGICAL EFFECTS
     The immunological status  of  male and female CD-I mice exposed to TCI in
drinking water for 4  and 6 months was investigated by Sanders et al.  (1982).
                                    5-21

-------
Details of exposure and discussion of toxicological effects noted were reported
separately by  Tucker  et  al.  (1982)  (see  Section  5.4.2).   In the  14-day  range-
-finding study,  male  CD-I mice had been dosed by p.o. gavage daily with 24 or
240  mg/kg  TCI.   Although the humoral immune  response  to sheep RBC was  not
significantly different from control, the cell-mediated  response was inhibited
in a dose-related manner.  To determine the immune response over longer periods
of time,  both  male and female mice were exposed to TCI  (0.1, 1.0, 2.5, and 5
mg/ml)  in  drinking water for 4 and  6 months.   Parameters evaluated  included
humoral  immunity,  cell-mediated  immunity,  lymphocyte responsiveness,  bone
marrow  function, and  macrophage function.  It was observed that male mice were
relatively unaffected after both  4 and 6 months  of exposure compared to results
in the  14-day  study.   On the other hand, females were affected, particularly
in the  4-mo  exposure.  Humoral  immunity  was  inhibited only at the  highest TCI
concentrations  (2.5  and  5  mg/ml),  whereas cell-mediated  immunity and  bone
marrow  stem  cell colonization were inhibited  at all  four concentrations of
TCI.   The proliferation  capacity of the  lymphocyte  was not affected when
challenged with T-cell mitogens.
     Hobara  et  al.  (1984)  investigated the acute effects of TCI  on hematologic
parameters  in  anesthetized, mature, cross-bred dogs.   A dose-related decrease
in  leukocyte counts was observed over the TCI  exposure  range  of 0 to 700 ppm
(0 to  3,766  mg/m ).   Five  dogs were exposed to  each  of the following treatments
for  one hour:    200,  500,  700,  1000,  1500, and  2000 ppm (1026,  2690,  5380,
                      3
8070,  and 10,760 mg/m ).   In addition,  five  dogs  were exposed to  700 ppm  for
one  hour.   Leukocyte counts decreased at the start of exposure and at the end
of  exposure reached  minimum values.   During postexposure, values gradually
 recovered.   No significant changes  were  observed in  erythrocyte  counts,  hemato-
 crit values, or thrombocyte  counts.   It was the  authors'  opinion that the
 decrease  in  leukocytes  reflects  the sequestering of  the  leukocytes in  tissues,
 as  a defensive mechanism.
      Koizumi et al.  (1984) found that 6-aminolevulinic  acid dehydratase (ALA-D),
 an  enzyme important in  heme synthesis in erythrocytes and liver, was inhibited
 in  rats exposed to 50,  400, or  800  ppm  (269,  2152, or 4304 mg/m3)  TCI  continu-
 ously for 48  or 240  hours.   A  dose-response relationship was observed.  TCA
 and TCE were  not  inhibitory j_n vitro.   Inhibition of ALA-D (38  percent or 48
 percent of  the control  in liver or  in  blood,  respectively)  observed after
                                     5-22

-------
240 hours  of  exposure to 400 ppm (2152 mg/m )  was  accompanied by the signifi-
cant elevation of 6-aminolevulinic acid synthetase (186 percent  of control) in
liver  and  an  increase in excretion of ALA  in  urine (142 percent of  control).
No effects on weight,  liver, or other hematological parameters were  observed.
     Fujita et  al.  (1984) also demonstrated an  inhibition  of ALA-D in rats
exposed to TCI  for 240  hours.   Exposure  levels  were   50, 377,  and  754 ppm
(270,  2140, and 4128 mg/m ).  TCI inhibited the holoenzyme only when incubated
with a mixed-function oxidase system, suggesting that metabolites play a role.
The apoenzyme  (without  the  zinc cofactor)  was inhibited  in  a dose-dependent
manner.  After  exposure  to  377  ppm  (2140  mg/m3), urinary  ALA increased to  142
percent of control.  In  livers of exposed rats, cytochrome P-450 concentrations
and heme saturation  of tryptophan pyrrolase were reduced.  Both measures are
indices of the  free  heme pool.   Lead levels in blood were normal, indicating
that lead  was  not  responsible for the inhibition.  Inhibition of ALA-D in  the
liver  was  observed at the lowest dose  and  after 48  hours exposure  to TCI.
5.6  DERMAL EFFECTS
     Industrial use  of  TCI  is often associated with dermatological problems
(Bauer  and  Rabens,  1974).   These include  reddening and  dermatographic skin
burns on contact (Maloof, 1949), generalized dermatitis resulting from contact
with the  vapor (McBirney, 1954;  Conde-Salazar  et al., 1983), and possible
scleroderma (Reinl,  1957;  Walder, 1983).  However, irritations,  burns,  and
rashes are usually the result of direct skin contact with concentrated solvent
and are, therefore,  probably  limited to  effects secondary to  the  solvent.  No
effects have been  reported with exposure to TCI in dilute aqueous solutions.
     Phoon et al.  (1984) reported recently on severe erythema (Stevens-Johnson
syndrome) in five  individuals occupationally exposed to TCI through vapor and
direct dermal  contact.   Duration  of exposure ranged from 2  to  5 weeks and
exposure levels were estimated to be between 9 and 169  ppm (50 and 912 mg/m ).
The common occupational  history and similarities in condition suggest  that TCI
may have  been  a causative  factor in eliciting a  hypersensitive  response.
5.7  MODIFICATION OF TOXICITY:   SYNERGISM AND ANTAGONISM
     The manifestation of adverse  effects  of TCI appears to depend,  in part,
on its metabolic  products.   Therefore,  drugs and chemicals that  enhance  or
                                    5-23

-------
inhibit the metabolism of TCI can have a corresponding effect on the toxicity
of TCI.   Reynolds  (1976) and Moslen  et al.  (1977) have reported  that  the
hepatotoxic potential of TCI can  be greatly enhanced by prior treatment with
drugs or chemicals that induce components of the liver mixed-function oxidase
(MFC1) system.   Carlson (1974) demonstrated  that pretreatment of male albino
rats with  phenobarbital  (50 mg/kg  i.p.) or  3-methylcholanthrene  (40 mg/kg)
exacerbated TCI-induced liver damage.   There was massive necrosis  and a  three-
fold increase  in SCOT and SGPT  levels.   Groups of animals were exposed to 6900
                3                                           3
ppm  (37,122 mg/m ) for 2  hours  and 10,400 ppm  (55,952  mg/m ) for 2 hours.
This was confirmed in  a  later  study of Moslen  et  al. (1977), who  found that
TCI  induced morphological changes  in  the liver of animals  treated with only
phenobarbital.  Reynolds  (1976)  also found acute liver injury following  a 2-hr
exposure to 1 percent TCI  in animals pretreated with other inducers of MFO,
including chlorinated biphenyls (Arochlor 1245).   These  findings suggest that
a metabolite(s) of TCI contributes to the toxicity of the compound.   In this
regard, Mikiskova  and Mikiska  (1966)  have  found  that  trichloroethanol,  a
metabolite of TCI, was three to five times more effective  in  inducing changes
in EKG and EEG.
     Masuda and Nakayama (1982) found that TCI  produced  moderate increases in
plasma glutamine pyruvate transminase activity  in  male  SPF  mice given an i.p.
dose of 2  ml/kg (dissolved in olive oil).   Diethyldithiocarbamate trihydrate
and  CS?,  when given to the  mice  orally, 30 minutes  before TCI,  exerted a
protective effect, in  that  the  increase in  plasma glutamine-pyruvate trans-
aminase was  lessened.   Diethyldithiocarbamate was effective  at each of the
three  dose  levels  used (10,  30, and 100 mg/kg).   In controls given only TCI,
no marked pathological alterations were reported upon histological examinations.
CS?, at levels of 3, 10,  and 30 mg/kg, acted similarly to diethyldithiocarbamate
in suppressing the TCI-associated increase in plasma glutamate-pyruvate trans-
ami nase.  The authors concluded that the protective mechanisms of  diethyldithio-
carbamate  and CS-  may involve their interference with metabolic activation of
a variety of  hepatotoxins.
     Novakova et  al.  (1981)  reported that  an  open-ended jjn vitro rat liver
perfusion  system  could be used to  demonstrate  the effect  of TCI  and other
halogenated solvents  on  liver enzyme activities.  TCI  in  such a  system was
reported  to  increase the activities  of aspartate aminotransferase, alamine
aminotransferase,  alkaline  phosphatase, and  ornithine  carbamoyltransferase.
                                     5-24

-------
     Pessayre et al.  (1982)  have obtained evidence that TCI is an effective
potentiator of hepatotoxic effects of CC1. jji vivo  in rats.  While liver his-
tology was normal 24 hours after i.p.  administration of TCI (1 ml/kg) and also
after administration  of  64  pi/kg CC1.,  both solvents coadministered produced
extensive centrilobular  necrosis.  TCI  given in vitro 5 hours prior to CC1.
                                              _"_                           *T
increased CCK-induced lipid peroxidation.   Similarly,  coadministration markedly
increased serum  alamine  ami no transferase and  decreased  hepatic  cytochrome
P450 concentrations.   Coadministration  also decreased  hepatic glutathione
levels.   It was  concluded  that the toxicologic effects produced by the con-
comitant administration  of TCI  and a small  dose of CC1.  resembled those pro-
duced by a  high  dose of  CC1., a  finding  consistent with  the view that TCI
potentiated the  hepatotoxicity of CCl^.   It was further concluded that this
potentiation may be the result of increased  CCl.-induced  rate of lipid peroxi-
dation.   Since stable  metabolites of  TCI  failed to stimulate such peroxida-
tion, it was apparently mediated by TCI  itself.
     The effect  of ethanol  and other substances  upon  the manifestation of
effects of  subsequent TCI exposure in animals has been studied by a number of
investigators.
     One of the earliest  studies  was  that  of Cornish and Adefuin (1966).   Male
Sprague-Dawley rats were orally pretreated with ethanol (50 percent in water)
in a single dose,  on the basis of 5  g/kg  body weight.  Sixteen to eighteen
hours after pretreatment, each group (6 rats) was exposed to TCI in a dynamic
exposure chamber.  Exposure levels were 100, 2000, 5000, and 10,000 ppm (538,
10,760,  26,900, and 53,800 mg/m ) TCI  (3 rats in the highest exposure category
died).   Animals  ingesting ethanol prior to exposure to TCI showed a statisti-
cally significant (p <0.05)  increase  in  SCOT, SGPT,  and serum isocitric dehydro-
genase (SICD)  at 5000 ppm (26,900 mg/m ).   After exposure  to 5000  ppm (26,900
mg/m ) for  4 hours,  the  SCOT of animals pretreated with ethanol was  about 20
times the control level,  whereas  exposed rats not receiving ethanol  had normal
SGOT values.  Animals  exposed  to 2000 ppm (10,760 mg/m )  or less  for 4 hours
showed no marked alteration  of serum enzyme levels.   Neither ethanol-treated
nor nontreated rats showed any alterations  of serum enzyme levels  when exposed
                                3                                             3
for 8 hours to 100  ppm (538 mg/m ).   All rats exposed to 5000 ppm  (26,900 mg/m )
TCI showed  slight to widespread degenerative lipid infiltration of the liver.
Ethanol-treated  rats  (before TCI  exposure)  exhibited early  centrilobular
necrosis with  slight acute inflammatory  cellular reaction.
                                    5-25

-------
     Muller et al. (1975) proposed that TCI and ethanol undergo similar meta-
bolic steps, thus allowing  for  a variety of interactions, depending on dose
levels and the sequence  of  exposure.   When simultaneous exposure occurs,  TCI
and ethanol may compete  for both alcohol  dehydrogenase and microsomal  mixed-
function oxidase.   Intolerance  of TCI-exposed workers  to alcohol has been
described by Defalque  (1961).   It is  often manifested as red flushes on the
face  and  upper parts of the body and has been referred  to  as  "degreasers
flush."  Stewart et al. (1974)  observed this effect in seven individuals given
as little  as 0.5  ml  ethanol/kg during exposure to 20, 100, and 200 ppm (108,
538,  1076 mg/m )  TCI  for 1, 3,  or 7.5 hours.   Alcohol intolerance among TCI-
exposed workers has also been described by Bardodej et al. (1952) and Bardodej
and Vyskocil (1956).
     Traiger and  Plaa (1974)  reported that isopropyl  alcohol  or acetone,
administered to  mice in an 18-hr  pretreatment, potentiated  the hepatotoxic
response to two doses of TCI, 11.1 mmol/kg and 16.7 mmol/kg.
      Cornish et  al.  (1977)  found that rats given i.p. doses of 0.5, 1.0, and
2.0 ml  TCI/kg  showed progressively increasing levels  of  SGOT.   Pretreatment
with  either phenobarbital  or ethanol  or both had no effect on this response.
This  lack  of  response with ethanol is  in  contrast to the earlier  study  by
Cornish  and Adefuin  (1966)  but may  reflect different routes of exposure.
      Utesch et al.  (1981),  in  developing  their animal  model  for solvent abuse
(see  Section 5.4.2),  found  that pretreatment of Wistar-Munich rats with ethanol
exacerbated the  CNS-depressant  effects of high doses  of TCI.   A group  of rats
was given  5 ml/kg ethanol orally  1 hour before initiation of the  TCI inhalation
regimen:   15  exposures cycles  in  a 5-hour period;  each cycle consisted of 5
minutes  of inhalation of 15,000  ppm  (80,700 mg/m3) TCI followed by a 15-min
solvent-free interval.   Ethanol was shown  to cause  significant  prolongation of
the  animals'  ability to regain the righting reflex after the second and  each
subsequent TCI exposure.  Administration  of ethanol alone did not result  in an
increase  in levels of SGOT  or SGPT in histopathologic  alterations of the  liver
or kidneys.  No  elevations  in SGPT were seen in any of the groups that  received
both  alcohol  and TCI.   Histopathologic examination of the liver and kidneys
revealed no evidence of  chemically induced injury in  any animals.   Fasting for
16 hours before  the  5-hr  intermittent TCI exposure regimen  had  no apparent
effect on the combined  CNS-depressant action  of  ethanol and TCI.  No histo-
pathological  effects were observed in the livers or kidneys of these animals.
                                     5-26

-------
     Moslen et al. (1977) also reported that fasting had little effect on SCOT
and SGPT levels in rats which inhaled 10,000 ppm (53,800 mg/m ) TCI continuously
for 2  hours.   However,  pretreatment  of  rats with phenobarbital  in  combination
with  fasting  produced significant elevations  in the levels of these  serum
enzymes.
     Cornish et al.  (1973)  reported that phenobarbital pretreatment produced
no additional  effect on SGOT above  that produced  by 1.0 and 2.0 ml TCI/kg,
given i.p.  to Sprague-Dawley rats.   Lower doses of TCI, 0.5 and 0.3 ml/kg, had
no effect on SGOT.
     Pretreatment of  male  albino rats with either phenobarbital or 3-methy1
cholanthrene was  shown  by  Carlson (1974) to potentiate the hepatotoxicity of
TCI.   Rats were injected with either 50 mg phenobarbital/kg i.p. for 4 days or
3-methyl chol anthrene  in  corn oil, 40 mg/kg i.p. for 2 days.   Animals were
exposed to TCI 24 hours after the last dose of phenobarbital or 48 hours after
the last dose  of  3-methylcholanthrene in a dynamic inhalation chamber.  Rats
were sacrificed 22 hours after exposure.  Phenobarbital pretreatment increased
SGOT  and SGPT  three-fold above TCI  controls.   TCI  exposure was 10,400 ppm
(55,952 mg/m ) for 2  hours.   Similarly, 3-methylcholanthrene increased SGOT
and SGPT.
5.8  SUMMARY OF TOXICOLOGICAL EFFECTS
     There is  no reliable  information concerning the  toxicological effects  in
humans of chronic  exposure to levels of TCI below the TL\r  (50 ppm).  Based
upon acute human overexposure information  and limited animal testing, it is
unlikely that  chronic  exposure  to TCI at levels found or expected in ambient
air would result in liver or kidney damage.   Such damage has not been generally
                                                ®
found even when exposure greatly exceeds the TLV .
     The first  sign  likely to be observed upon excessive exposure to TCI is
central nervous system (CNS) dysfunction.   Psychomotor function and subjective
complaints which have  been  studied in short-term controlled  human studies,  as
well as data  from  accidental  exposure and epidemiological studies, indicate
that nervous system function probably is affected by TCI concentrations ranging
from 200  to  500 ppm (1076 to 2690 mg/m ).   Dose-response  relationships  for
adverse health effects associated with TCI  have not been established in man or
experimental  animals.
                                    5-27

-------
5.9  REFERENCES


Adams, E.M., H.C. Spencer, V.K. Rowe, D.D. McCollister, and D.D. Irish. 1951.
     Vapor toxicity of trichloroethylene determined by experiments on labora-
     tory animals.  Arch. Ind. Hyg. Occup. Med.  4:469-481.

Atkinson, R.S. 1960. Trichloroethylene anaesthesia.  Anesthesiology 21:67-77.

Aviado, D. , S. Zakhari, J. Simaar, and A. Ulsamer. 1976. Methyl Chloroform and
     TCE in the Environment.  CRC Press, Cleveland, OH.

Ayre, P. 1945. Acute yellow fever after Trilene anesthesia.  Correspondence,
     Brit.  M.J. 2:784.

Baerg, R.D., and D.V. Kimberg. 1970. Centrilobular hepatic necrosis and acute
     renal failure in "solvent sniffers."  Ann. Intern. Med.  73:713-720.

Baetjer, A. M., Z. Annau, and H. Abbey. 1970. Water deprivation and trichloro-
     ethylene:  effect on hypothalamic self-stimulation.  Arch. Environ.
     Health 20:712-719.

Baker, A.B. 1958. The nervous system in trichloroethylene:  an experimental
     study.  Neuropathol. Exp. Neurol.  17:649-655.

Bardodej, Z. ,  and J. Vyskocil. 1956. The problem of trichloroethylene in
     occupational medicine:  trichloroethylene metabolism and its  effects on
     the nervous  system  as a means of  hygienic control.  AMA Arch.  Ind. Health
     13:581-592.

Bardodej,  Z.,  I.  Berka,  B. Chaluna, 0. Nesvadha, and J. Vyskocil.  1952. Recent
     advances  in  our knowledge of  the  effects  of trichloroethylene upon the
     health of workers.   Prac. Lek. 4:441-467.

Barnes, C.G.  and  J.  Ives. 1944. Electrocardiographic changes during trilene
     anesthesia.  Proc.  Roy.  Soc.  Med.   37:528-532.

Barret, L., Ph.  Arsac, M. Vincent, J.  Faure, S. Garrel, and F.  Reymond. 1982.
     Evoked trigeminal nerve  potential  in chronic  trichloroethylene intoxica-
     tion.  J. Toxicol.  Clin.  Toxicol.  19(4)-.419-423.

Bartonicek, V.J., and  A.  Brun. 1970. Subacute  and  chronic  trichloroethylene
     poisoning:   a  neuropathological study  in  rabbits.  Acta Pharmacol. Toxicol.
     28:359-369.

Battig,  K., and  E.  Grandjean.  1963.  Chronic effects  of trichloroethylene  on
      rat  behavior.   Arch.  Environ.  Hlth.  7:694-699.

Bauer,  M.,  and S.F.  Rabens. 1974.  Cutaneous manifestations of  trichloroethylene
     toxicity.  Arch.  Dermatol.   110:886-890.

Bell,  A.  1951. Death from trichloroethylene in a  dry-cleaning  establishment.
      N.Z.  Med. J.   50:119-126.
                                     5-28

-------
Berleb, M.  1953.  Eklampsie und Trichloran-Analgesie, Miinchen. Med. Wehnschr.
     13: 1225.

Bernstine,  M.L.  1954. Cardiac arrest occurring under trichloroethylene anal-
     gesia.   AMA Arch.  Surg.   68:262-266.

Bolt, H. M.  and J.  G. Filser. 1977. Irreversible binding of chlorinated ethy-
     lenes  to macromolecules.  Environ. Hlth. Persp.  21:107-112.

Boulton, T.B.,  and R.B.  Sweet. 1960. The place of trichloroethylene  in modern
     anesthesia.   J.  Mich. State Med.  Soc.  59:270-273.

Brittain, G.J.C.  1948.  Trichloroethylene (trilene) as anesthetic in  neuro-
     surgery.  Anesth.  & Analg.   27:145-151.

Carlson, G.  P.,  and J.  F. White. 1983. Cardiac arrhythmogenic action of ben-
     zo(a)pyrene in the rabbit.   Toxicol.  Lett.  15:43-43.

Carlson, G.P. 1974. Enhancement of the hepatotoxicity of trichloroethylene by
     inducers of drug metabolism.  Res. Commun. Chem. Pathol. Pharmacol.
     7:637-640.

Clearfield,  H.  1970.  Hepatorenal toxicity from sniffing spot remover (tri-
     chloroethylene).  Dig. Dis.  15:851-856.

Coleridge,  H.M.,  J.C.G.  Coleridge, J.C. Luck and J. Norman. 1968. The effect
     of four volatile anaesthetic agents on the impulse activity of  two types
     of pulmonary receptor.  Br. J. Anaesth.  40:484-492.

Conde-Salazar,  L.,  D. Guimaraens, L. V. Romero, and E. Sanchez Yus.  1983.
     Subcorneal  pustular eruption and erythema from occupational exposure to
     trichloroethylene.   Contact Dermatitis.  9:235-237.

Cornish, H.H.,  and J. Adefuin. 1966. Ethanol potentiation of halogenated
     aliphatic solvent toxicity.  Am.  Ind. Hyg. Assoc. J.  27:57-61.

Cornish, H.H.,  M.L. Barth, and B. Ling. 1977. Influence of aliphatic alcohols
     on the hepatic response to halogenated olefins.  Environ. Hlth. Persp.
     21:149-152.

Cornish, H.H.,  B.P. Ling, and M.L. Barth. 1973. Phenobarbital and organic
     solvent toxicity.   Phenobarbital  and organic solvent toxicity.  Amer.
     Ind. Hyg.  Assoc. J. 34:   487-492.

Cotter, L.H. 1960.  Trichloroethylene poisoning.  Arch. Ind. Hyg. Occup. Med.
     1:319-322.

Danni, 0.,  0. Brossa, E. Burdino, D. Milillo, and G. Ugazio. 1981. Toxicity
     of halogenated hydrocarbons in pretreated rats.  An experimental model
     for the study of integrated permissible limits of environmental poisons.
     Int. Arch.  Occup.  Environ.  Health 49:165-176.

Defalque, R.J.  1961.  Pharmacology and toxicology of trichloroethylene:  a
     review of the world literature.  Clin. Pharmacol. Ther.  2:665-688.
                                    5-29

-------
Dhuner, K.  G.  1951.  Cardiac irregularities due to trichloroethylene given
     during labor.   Acta Obst.  Ct Gynee.  Scandinav. 31:478-482.

Dhuner, K.  G.  1957.  Cardiac irregularities in trichloroethylene poisoning.
     Acta Anaesth.  Scand. 1:121-135.

DiGiovanni, A.  J.  1969.  Reversal of chloral hydrate.  Associated cardiac
     arrhythmia by a beta-adrenergic blocking agent.  Anesthesiology 31: 93.

Dobkin A.,  and P.  Byles. 1963.  Trichloroethylene anesthesia.  Clin. Anesth.
     1:44-65.

Oodds, G. H. 1945.  Necrosis of liver and bilateral massive suprarenal hemor-
     rhage in puerperium, Brit. M. J. 1:769-770.

Elam, J. 1949.  Correspondence.   Brit. M.  J. 1:546-547.

Ettema, J.  H.,  R.  L. Zielhuis,  E. Burer,  H. A. Meier, L. Kleerekoper, and M.
     A. de Graaf.  1975.  Effects of alcohol, carbon monoxide, and trichloro-
     ethylene on mental  capacity.  Int. Arch. Occup. Environ. Health 35:
     117-132.

Fink, R. 1968.  Toxicity of Anesthetics, Williams and Wilkins Co.,  Baltimore,
     MD.  p. 270.

Fishbein, L. 1976.  Industrial mutagens and potential mutagens.  I. Halogenated
     aliphatic derivatives.  Mutat. Res.  32:267- .

Fossa, A.A., J.F.  White, and G.P. Carlson. 1982. Antiarrhythmic effects of
     disulfiram on epinephrine-induced cardiac arrhythmias in rabbits exposed
     to trichloroethylene.  Toxicol.  Appl. Pharmacol.  66:109-117.

Fujita, H., A.  Koizumi, M. Yamamoto, M. Kumai, T.  Sadamoto, and M. Ikeda.
     1984.   Inhibition of 6-aminolevulinate dehydratase  in trichloroethylene-
     exposed rats, and  the effects on heme regulation.  Biochim. Biophys. Acta
     800: 1-10.

Gehring, P.O. 1968. Hepatotoxic potency of various  chlorinated hydrocarbon
     vapors relative to their  narcotic and lethal  potencies  in mice.  Toxicol.
     Appl.  Pharmacol.   13:287-298.

Goldberg, M.E., H.E. Johnson,  U.C.  Pozzani, and  H.F. Smyth,  Jr. 1964a.  Effect
     of  repeated inhalation of vapors of  industrial  solvents on animal  behavior.
     I.  Evaluation of  nine solvent  vapors and pole-climb performance in  rats.
     Am. Ind. Hyg. Assoc. J.   25:369-375.

Goldberg, M.E., H.E. Johnson,  V.C.  Pozzani, and  H.F. Smyth,  Jr. 1964b.  Behav-
     ioral  response of  rats during  inhalation of trichloroethylene and  carbon
     disulfide vapors.  Acta Pharmacol. Toxicol. 21:36-44.

Grandjean,  E., R. Muchinger, V. Turrian,  P.A. Haas,  H.K.  Knoepfel, and  H.
     Rosenmund. 1955. Investigations into  the effects of  exposure  to trichloro-
     ethylene  in mechanical engineering.   Brit.  J.  Ind. Med. 12:131-142.
                                     5-30

-------
Grandjean, E. 1960. Trichloroethylene effects on animal behavior.  Arch.
     Environ. Hlth.  1:106-108.

Grandjean, E. 1963. The effects of short exposures to trichloroethylene on
     swimming performances and motor activity of rats.  Am. Ind. Hyg. Assoc. J.
     24:376-379.

Guyotjeannin, C. , and J. Van Steenkiste. 1958. Action of trichloroethylene on
     proteins and lipoproteins.  Study of 18 workers working  in polluted
     atmosphere.  Arch. Mai. Prof. Med. Trav. Secur. Soc.  19:489-494.

Haglid, K.G., C. Briving, H.A. Hansson, L. Rosengren, P. Kjellstrand, D.
     Stavron, U. Swedin, and A. Wronski. 1981. Trichloroethylene:  long-lasting
     changes in the brain after rehabilitation.  Neurotoxicol. 2:659-673.

Heizer, H. 1951. Trichloraethylen in der Gebutshilfe, Munchen. Med. Wehnschr.
     93:2046-2050.

Hobara, T., H.  Kobayashi, E. Higashihara, T. Kawamoto, and T. Sakai. 1984.
     Acute effects of 1,1,1-trichloroethane, trichloroethylene, and toluene  on
     the hematologic parameters in dogs.  Arch. Environ. Contam. Toxicol.
     13:589-593.

Humphrey, J.H., and M. McClelland. 1944. Cranial-nerve palsies with herpes
     following general anaesthesia:  a report.  Br. Med. J.   1:315-318.

Ikeda, M., and T. Imamura. 1973. Biological half-life of trichloroethylene and
     tetrachloroethylene in human subjects.  Int. Arch. Arbeitsmed.  31:209-224.

Ikeda, T., C. Nagano, and A. Okada. 1969. Hepatotoxic effect  of trichloro-
     ethylene and perch!oroethylene in the rat and mouse.  Igaku to Seibutsugaku
     79:123-129.

James, W.R.L. 1963. Fatal addiction to trichloroethylene.  Br. J.  Ind. Med.
     20:47-49.

Jones, N. M., G. Margolis, and C. R. Stephens. 1958. Hepatotoxicity of inhala-
     tion anesthetic drugs.  Anesth. 19:715-723.

Joron, G.E., D.G. Cameron, and G.W. Halpenny. 1955. Massive necrosis of the
     liver due to trichloroethylene.  Can. Med. Assoc. J.  73:890-891.

Kanje, M. P. Kjellstrand, K. Fex and A. Waldorf. 1981. Neurotransmitter
     metabolizing enzymes and plasma butyrylcholinesterase in mice exposed to
     trichloroethylene.  Acta. Pharmacol. Toxicol. 49:205-209.

Kjellstrand, P., M. Kanje, L. Mansson, M. Bjerkemo, I. Mortensen,  J. Lanke,
     and B. Holmquist. 1981. Trichloroethylene: Effects on body and organ
     weights in mice, rats, and gerbils.  Toxicol. 21:105-115.

Kjellstrand, P., A. Edstrom, M. Bjerkemo, and B. Holmqvist. 1982.  Effects of
     trichloroethylene inhalation on acid phosphatase in rodent brain.  Toxicol.
     Lett. 10:1-5.
                                    5-31

-------
Kjellstrand, P., B.  Holmquist, N.  Mandahl, and M. Bjerkemo. 1983a. Effects of
     continuous trichloroethylene inhalation on different strains of mice.
     Acta Pharmacol.  Toxicol. 53:369-374.

Kjellstrand, P., B.  Holmquist, P.  Aim, M. Kanje, S. Romare, I. Jonsson, L.
     Mannson, and M.  Bjerkemo. 1983b. Trichloroethylene:  further studies of
     the effects on body and organ weights and plasma butyryl cholinesterase
     activity in mice.  Acta Pharmacol. Toxicol. 53:375-384.

Klaassen, C.D., and G.L. Plaa. 1967. Relative effects of various chlorinated
     hydrocarbons on liver and kidney function in dogs.  Toxicol. Appl.
     Pharmacol.  10:119-131.

Kledecki, Z., and H.  Bura. 1963. Ostre zatrucia wywolane polknieciem troje-
     hloroethylenu ("Tri").  Pol.  Tyg. Lek.  18:748-750.

Kleinfeld, M. and I.R. Tabershaw.  1954. Trichloroethylene toxicity:  report
     of five fatal cases.  AMA Arch. Ind. Hyg. Occup. Med.  10:134-141.

Koizumi, A., H. Fujita, T. Sadamoto, M. Yamamoto, M. Kumai, and M. Ikeda.
     1984. Inhibition of 6-aminolevulinic acid dehydratase  by trichloroethy-
     lene.  Toxicol.  30:93-102.

Konietzko, H.,  I. Elster,  K. Vetter, and H. Weichardt.  1974.  Field studies  in
     solvent factories: Second communication: Telemetric EEG monitoring of
     solvent workers at trichloroethylene basins.   Arbeitsschutz  23(5):129-133.

Kunz, E., and R. Isenschmid.  1935.  The toxic effect of  trichloroethylene  on
     the eye.   Klin. Monatsbl. Augenheilkd.  94:577-585.

Kylin, B., H. Reichard,   I.  Sumergi, and S. Yllner. 1963.  Hepatotoxicity  of
     inhaled trichloroethylene, tetrachloroethy1ene, and chloroform.   Single
     exposure.  Acta  Pharmacol.  Toxicol.   20:16-26.

Kylin, B. ,  K. Axell, fl.E.  Samuel, and  A.  Lindborg.  1967. Effect  of  inhaled
     trichloroethylene  on the CNS as measured by optokinetic  nystagmus.   Arch.
     Environ.  Hlth.   15:48-52.

Kyrklund, T.,  C. Ailing,  K.  Haglid, and  P.  Kjellstrand.  1983.  Chronic  exposure
     to  trichloroethylene:   lipid and  acyl  group composition  in  gerbil cerebral
     cortex and hippocampus.   Neurotox.   4:35-47.

Kyrklund, T.,  C. Ailing,  P.  Kjellstrand,  and  K.G.  Haglid.  1984.  Chronic
     effects of perchloroethy1ene on the composition  of lipid and acyl groups
     in  cerebral  cortex and hippocampus  of the  gerbil.   Toxicol.  Lett. 22:
     343-349.

 Lachnit,  V., and G.  Brichta.  1958.  Trichloroethylene  and liver damage.  Zbl.
     Arbeitsmed.   8:56-62.

 Lawrence, W.H.  and E.K.  Partyka.  1981.  Chronic  dysphagra and trigeminal
     anesthesia after trichloroethylene  exposure.   Ann. Int.  Med. 95(6):710.

 Lehmann, K.B.  1911.  Experimental  studies on the technical  influence and
     essential  hygienic gases and steam on the  organism.   Arch.  Hyg.  74:1-60.

                                     5-32

-------
Lewis, G.  D., R. C. Reynolds, and A. R. Johnson. 1984. Some effects of tri-
     chloroethylene on mouse lungs and livers.  Gen. Pharmac. 15:139-144.

Lilis, R. , D. Stanescu, N. Muica, and A. Roventa. 1969. Chronic effects of
     trichloroethylene exposure.  Med.  Lav.  60:595-601.

Longley, E.G., and R.  Jones. 1963. Acute trichloroethylene narcosis.  Arch.
     Environ.  Health 7(2):249-252.

Maloof, C.C. 1949. Burns of the skin produced by trichloroethylene vapors at
     room temperature.  J. Ind. Hyg. Toxicol.  31:295-296.

Masoero, V., and A. Lavarino. 1955. Acute toxicosis from trichloroethylene.
     Patol.  Sper. 43:124-129.

Masuda, Y. and N. Nakayama. 1982. Protective effect of diethyldithiocarbamate
     and carbon disulfide against liver injury induced by various hepatotoxic
     agents.  Biochem. Pharmac. 31(17):2713-2725.

Mazza, V., and A. Brancaccio. 1967. Characteristics of the formed elements of
     the blood and bone marrow in experimental trichloroethylene intoxication.
     Folia.  Med.  50:318-324.

McBirney,  R.S. 1954.  Trichloroethylene and dichloroethylene poisoning.  AMA
     Arch. Ind. Hyg.  Occup. Med.  10:130-133.

McCarthy,  T. B. and R. D. Jones. 1983.  Industrial gassing poisonings due to
     trichloroethylene, perchloroethylene, and 1,1,1-trichloroethane, 1961-1980.
     Br. J.  Ind. Med.  40:450-455.

Mikiskova, H. , and A.  Mikiska.  1966. Trichloroethanol in trichloroethylene
     poisoning.  Br.  J. Ind. Med.  23:116-125.

Mitchell,  A.B.S., and B.C. Parsons-Smith. 1969. Trichloroethylene neuropathy.
     Br. Med. J.  1:422-423.

Moslen, M.T., E.S. Reynolds, P.J. Boor, K. Bailey, and S. Szabo. 1977. Tri-
     chloroethylene- induced deactivation of cytochrome P-450 and loss of liver
     glutathione in vivo.  Res. Commun. Chem. Pathol. Pharmacol.  16:109-120.

Mroczek, H., and T. Fedyk. 1971. Przpadek ciezkiego, samobojczego zatrucia
     trojechloroetylenem ("Tri").  Pol. Tyg. Led.  26:1509-1510.

Muller, G.,  M. Spassovski, and D. Henschler. 1975. Metabolism of trichloro-
     ethylene in man.   III.  Interaction of trichloroethylene and ethanol.
     Arch. Toxicol.  33:173-189.

NIOSH. 1980. Registry of Toxic Effects of Chemical Substances, update.

Nomiyama,  K., and H.  Nomiyama.  1977. Dose-response relationship for trichloro-
     ethylene in man.   Int. Arch. Occup. Environ. Hlth.  39:237-248.
                                    5-33

-------
Novakova, V., J.  Musil, D.  Huckiova, 0., Taborsky, H., Sollova, and P. Vyborny.
     1981.  Effect of tetrachloromethane and other chlorinated hydrocarbons on
     the hepatic metabolism in the isolated perfused rat liver.  J. Hyg.
     Epidemiol.  Microbiol.  Immunol.  25(4):369-383.

Nowill, W.K., C.R.  Stephen, and G. Margolis. 1954. The chronic toxicity of
     trichloroethylene:  a study.   Anesthesiology  15:462-465.

Ogata, M. ,  Y. Takatsuka, and K. Tomokuni. 1971. Excretion of organic chlorine
     compounds in the urine of persons exposed to vapours of trichloroethylene
     and tetrachloroethylene.   Br. J. Ind. Med.  28:386-391.

Page, N. P. and J.  L. Arthur.  1978.  Special Occupational Hazard Review of
     Trichloroethylene.  U. S. Department of Health, Education and Welfare,
     Public Health Service, Center for Disease Control, National Institute of
     Occupational Safety and Health, January.

Patty, F.A. 1958. Industrial Hygiene and Toxicology.  3 volumes, 2nd Revised
     Edition.  Interscience Publishers, New York, NY.

Pendergast, J.A., R.A. Jones,  L.J. Jenkins, Jr., and J. Siegel. 1967. Effects
     on experimental animals of long-term inhalation of trichloroethylene,
     carbon tetrachloride, 1,1,1-trichloroethane, dichlorodifluoromethane, and
     1,1-dichloroethylene.  Toxicol. Appl. Pharmacol.  10:270-289.

Persson, H. 1934. On trichloroethylene intoxication.  Acta Med. Scand.
     59:410-422.

Pessayre, D., B. Cobert, V. Decatoire, C. Degott, G. Babany, C. Funck-Brentano,
     M. Delaforge, and D.  Larrey.  1982. Hepatotoxicity of trichloroethylene-
     carbon tetrachloride  mixtures in rats.  Gastroent. 83:761-772.

Phoon, W.H., M.O.Y. Chan,  V.S. Rajan, K.J. Tan, T. Thirumoorthy, and C.L. Goh.
     1984 Stevens-Johnson  syndrome associated with occupational exposure to
     trichloroethylene.  Contact  Dermatitis 10:270-276.

Plessner, W. 1915. On trigeminal  disease  due to trichloroethylene  intoxication.
     Neurol. Zentr.  34:916-918.

Ramadan, M.A., and M.I.A.  Ramadan. 1969.  Histochemical studies on  the effect
     of  some anaesthetics  on  rat  liver.   Acta  Histochem. 34:310-316.

Reinhardt, C. F.;  L. S.  Mullen; and  M.  E. Maxfield.  1973. Epinephrine-induced
     cardiac arrhythmia  potential of some common  industrial  solvents.   J.
     Occup.  Med. 15:953.

Reinl, W.  1957.  Scleroderma under the  influence of  trichloroethylene.
     Zentralbl.  Arbeitsmed.   7:58-60.

Reynolds,  E. 1976. A  return to trichloroethylene.   Br. J. Anaesth.   48:51
      (Letter).

Salvini, M., S.  Binaschi and  M. Riva.  1971.  Evaluation of the psychophysiolog-
      ical  functions  in  humans exposed  to  trichloroethylene.   Br.  J.  Ind. Med.
      28:293-295.

                                     5-34

-------
Sanders, V.M., A.N. Tucker, K.L. White, Jr. B.M. Kauffman,  P.  Hallett,  R.A.
     Carchman, J.F. Borzelleca, and A.E. Munson. 1982. Humoral and  cell-
     mediated immune status in mice exposed to trichloroethylene  in drinking
     water.  Toxicol. Appl. Pharmacol. 62:358-368.

Secchi, G.C., G. Chiappino, A. Lotto, and N. Zurlo. 1968. Actual  chemical
     composition of commercial trilene and their hepatotoxic effects:   clinical
     enzymologic study.  Med. Lav.  59:486-497.

Seifter, J. 1944. Liver injury in dogs exposed to trichloroethylene.  J.  Ind.
     Hyg. Toxicol.  26:250-252.

Silverman, A.P.  and H. Williams. 1975. Behaviour of rats exposed  to trichloro-
     ethylene vapours.  Br. J. Ind. Med.  32:308-315.

Steinberg, W. 1981. Residual neuropsychological effects following exposure to
     trichloroethylene (TCE):  a case study.  Clin. Neuropsy.  3(3):1-4.

St. Hill, C.A.  1966. Occupation as a cause of sudden death.  Trans.  Soc.
     Occup. Med.  16:6-9.

Stewart, R.D., and H.C. Dodd. 1964. Absorption of carbon tetrachloride,
     trichloroethylene, tetrachloroethylene, methylene chloride,  and 1,1,1-
     trichloroethane through the human skin.  Am. Ind. Hyg. Assoc.  J.   25:
     439-446.

Stewart, R.D., H.C. Dodd,  H.H. Gay, and D.S. Erley. 1970. Experimental  human
     exposure to trichloroethylene.  Arch. Environ. Hlth.   20:64-71.

Stewart, R.D., C.L. Hake,  and J.E. Peterson. 1974. "Degreasers1 flush":
     dermal response to trichloroethylene and ethanol. Arch. Environ. Hlth.
     29:1-5.

Stopps, G.J., and M. McLaughlin. 1967. Psychophysiological  testing  of human
     subjects exposed to solvent vapors.  Am. Ind. Hyg. Assoc. J.   28:43-50.

Stott, W. T., J. F. Quast, and P. G. Watanabe. 1982. Pharmacokinetics and
     macromolecular interactions of trichloroethylene in mice  and rats.
     Toxicol. Appl. Pharmacol. 62:137-151.

Striker, C., S.  Goldblatt, I.S. Wann, and D.E. Jackson. 1935.  Clinical  exper-
     ience with use of trichloroethylene in production of over 300  analgesias
     and anesthesias.  Anesth. and Analg.  14:68-71.

Stuber, K. 1931. Gesundheitsschadigen bei der Gewerblichen  Verwendung des
     Trichloroalthylens und die Monglichkelien ihrer Verhutung.   Arch.
     Gewerbepath. Gewerbehyg.  2:398-456.

Suciu, I. and L. Olinici.  1983. Hepato-renal involvement in acute occupational
     trichloroethylene intoxication. Med. Lav. 74:123-128.

Tham, R., B. Larsby, L. Odkvist, B. Norlander, D. Hyden, G. Aschan,  and
     A.  Bertler. 1979. The influence of trichloroethylene and  related drugs
     on the vestibular system.  Acta Pharmacol. Toxicol.  44:336-342.


                                    5-35

-------
Tolot, F., J. Viallier, A. Roullett, J. Rivoire, and J.C. Figueres. 1964.
     Hepatic toxicity of trichloroethylene.  Arch. Mai. Prof. Med. Trav.
     Secur. Soc.  25:9-15.

Tomasini, M. , and E. Sartorelli. 1971. Chronic intoxication from commercial
     trilene inhalation with compromise of the eighth cranial nerves.  Med.
     Lav.  62:277-280.

Traiger, G.J., and G.L. Plaa. 1974. Chlorinated hydrocarbon toxicity.  Arch.
     Environ. Health 28:276-278.

Triebig, G., P. Trautner, D. Weltle, E. Saure, and H. Valentin. 1982.  Investi-
     gations on the neurotoxicity of chemical substances at the workplace.
     III. Determination of the motor and sensory nerve conduction  velocity  in
     person occupationally exposed to trichloroethylene.  Int. Arch. Occup.
     Environ. Health 51:25-34.

Tucker, A.N., V.M. Sanders, D.W. Barnes, T.J. Bradshaw, K.L. White, I.E.  Sain,
     J.F. Borzelleca, and A.E. Munson. 1982. Toxicology of trichloroethylene
     in the mouse.  Toxicol. Appl. Pharmacol. 62:351-357.

Utesch, R. C.; F. W. Weir; and J. V. Bruckner. 1981. Development of an animal
     model of solvent abuse for use in evaluation of extreme trichloroethylene
     inhalation.  Toxicology 19(2):169-182.

Vallee, C., and J. Leclercq. 1935. Intoxication by trichloroethylene.  Ann.
     Med. Leg. Criminol. 15:10-12.

Vernon, R.J., and R.K.  Ferguson. 1969. Effects of trichloroethylene on visual-
     motor performance.  Arch. Environ. Hlth.  18:894-900.

Walder, B. K. 1983. Do solvents cause scleroderma. Int. J. Dermat.  22:157-158.

Werch, S. C., G. H. Marquardt, and J. F. Mallach. 1955. The action of  trichloro-
     ethylene on the. cardiovascular system.  Am. J. Obst. and Gyn. 69:352-364.

White, J. F. and G. P.  Carlson. 1979. Influence of alterations in  drug metab-
     olism on spontaneous and epinephrine-induced cardiac arrhythmias  in
     animals exposed to trichloroethylene.   Toxicol. Appl. Pharmacol.  47:
     515-527.

White, J.F.  and G.P. Carlson. 1981. Epinephrine-induced cardiac arrhythmias in
     rabbits exposed to trichloroethylene:   potentiation by ethanol.   Toxicol.
     Appl.  Pharmacol. 60:466-471.

White, J.F.  and G.P. Carlson. 1982. Epinephrine-induced cardiac arrhythmias in
     rabbits exposed to trichloroethylene:   potentiation by caffeine.  Funda-
     mental  Appl. Toxicol. 2:125-129.

Whitteridge, D. , and E. Bulbring.  1944. Changes in activity of pulmonary  re-
     ceptors in anesthesia and their  influence on respiratory behaviour.   J.
     Pharmacol. Exptl. Therap.  81:340-359.

Wirtschafter, Z.T., and M.W.  Cronyn.  1964.  Relative  hepatotoxicity.  Arch
     Environ. Hlth.  9:180-185.

                                    5-36

-------
Yacoub, M.,  J.  Faure, R.  Rollux, J. Mallion, and C. Marka. 1973. L1intoxication
     aigue par le trichloroethylene.  Manifestations cliniques et paracliniques,
     surveillance de 1'elimination du produit.   J.  Eur. Toxicol.  6:275-283.
                                    5-37

-------
           6.   TERATOGENICITY, EMBRYOTOXICITY, AND REPRODUCTIVE EFFECTS
     Because of  its widespread  use,  TCI has been  studied  for teratogenic
potential.   Teratology studies have been performed  in rats, mice, and rabbits
using doses of TCI  which,  in some studies,  produced slight signs of maternal
toxicity.  Also,  TCI is known to be metabolized in the maternal host by hepatic
metabolizing enzymes (and possibly also in the fetal liver) to chloral hydrate
and then  to trichloroethanol and trichloroacetic acid.   These metabolites,
particularly trichloroethanol, have also been shown  to readily cross the human
placenta into the  fetal  circulation and amniotic fluid (Bernstine and Meyer,
1953; Bernstine  et al., 1954) and also into the breast milk of nursing mothers
(Bernstine et al.,  1956).  There  is little  information available on the feto-
toxic and teratogenic potential of these metabolites.  Chloral hydrate and its
metabolite, trichloroethanol, have been in common medical use for a great many
years as  hypnotics, including use  during pregnancy, but with no  reported
adverse teratogenic, fetotoxic,  or  reproductive effects (Goodman and Gil man,
1980).  Trichloroethanol  has been administered  to  three animal species at
various stages  of pregnancy,  at levels as high as  700  mg/kg/day,  without
dose-related effects  (Physicians Desk  Reference, 1981).   Other studies in
avian embryo systems  (Fink,  1968; Elovaara et al.,  1979) have indicated that
TCI disrupts embryogenesis  in a  dose-related manner.  Another study  in the
avian embryo system (Bross et al., 1983) indicated that TCI is embryotoxic and
teratogenic in chick embryos.  However, because administration of TCI directly
into the air space  of chicken embryo  is not comparable to administration of
dose to animals  with  a placenta, it is not possible to interpret this result
in relationship  to the potential  of TCI to cause adverse effects in animals or
humans.   The  following  discussion  of  studies  subscribes  to  the  basic
viewpoints and  definitions  of the  terms "teratogenic"  and  "fetotoxic" as
summarized and  stated by the  U.S.  Environmental Protection Agency (1980):

     Generally,  the term "teratogenic" is defined as the tendency to produce
physical and/or  functional  defects in offspring jri utero.   The term "fetotoxic"
has traditionally been used to describe a wide variety of embryonic and/or
fetal divergences from the normal which cannot be classified as gross terata
(birth defects)  -- or which are of unknown or doubtful significance.   Types of
effects which fall under the very broad category of  fetotoxic effects are
death, reductions in fetal  weight, enlarged renal pelvis edema, and increased
                                     6-1

-------
incidence of supernumerary ribs.   It should be emphasized, however, that the
phenomena of terata and fetal toxicity as currently defined are not separable
into precise categories.   Rather, the spectrum of adverse embryonic/fetal
effects is continuous, and all deviations from the normal must be considered
as examples of developmental toxicity.  Gross morphological terata represent
but one aspect of this spectrum,  and while the significance of such structural
changes is more readily evaluated, such effects are not necessarily more
serious than certain effects which are ordinarily classified as fetotoxic--fetal
death being the most obvious example.

     In view of the spectrum of effects at issue, the Agency suggests that it
might be useful to consider developmental toxicity in terms of three basic
subcategories.  The first subcategory would be embryo or fetal lethality.
This is, of course, an irreversible effect and may occur with or without the
occurrence of gross terata.  The  second subcategory would be teratogenesis and
would encompass those changes (structural and/or functional) which are induced
prenatally, and which are irreversible.  Teratogenesis includes structural
defects apparent in the fetus, functional deficits which may become apparent
only after birth, and any other long-term effects (such as carcinogenicity)
which are attributable to jm utero exposure.   The third category would be
embryo or fetal toxicity as comprised of those effects which are potentially
reversible.  This subcategory would therefore include such effects as weight
reductions, reduction in the degree of skeletal ossification, and delays in
organ maturation.

     Two major problems with a definitional scheme of this nature must be
pointed out, however.  The first is that the reversibility of any phenomenon
is extremely difficult to prove.   An organ such as the kidney, for example,
may be delayed in development and then appear to "catch up."  Unless a series
of specific kidney function tests are performed on the neonate, however, no
conclusion may be drawn concerning permanent organ function changes.  This
same uncertainty as to possible long-lasting aftereffects from developmental
deviations is true for all examples of fetotoxicity.  The second problem is
that the reversible nature of an embryonic/ fetal effect in one species might,
under a given agent, react in another species in a more serious and irrever-
sible manner.
6.1 ANIMAL STUDIES

6.1.1  Mouse
     Schwetz and  his  associates (Schwetz et al.,  1975;  Leong et al. , 1975)

investigated the  potential  of TCI to adversely  affect  fetal  development in

Swiss-Webster mice  (Table  6-1).   Thirty to forty  bred  mice were exposed by

inhalation  to  TCI (99.24-percent pure)  at  an  air concentration of 300  ppm

(1614 mg/m ) for 7 hours daily on days 6 through 15 of gestation.  The results

from two  separate trials of 12 and  18  pregnant animals were reported.  The

first day  of  pregnancy was designated as  the  first day a  vaginal  plug  was

observed.  Concurrent pregnant control mice (30) were exposed to filtered air.

The mice  were  observed daily throughout pregnancy, and maternal body weights


                                     6-2

-------
TABLE 6-1.   SUMMARY  OF  ANIMAL  STUDIES  OF  FETOTOXIC  AND  TERATOGENIC  POTENTIAL OF  TCI

Fink (1968)

Elovaara
(1979)

Schwetz et al .
(1975)




cn
do Bell (1977)


Dorfmuel ler
et al. (1979)




Bellies et al .
(1980)
TCI purity
Unknown

Reagent grade


Technical
99.24 % TCI
0.76 % stabi-
lizers and
impurities
(Neu-tri)

Technical
(Trichlor 132)

Technical
99 %+ TCI
0.2 % epichloro-
hydri n
(Neu-tri)

Technical
99.9 %
Conditions
(mode of administration, dosage
Species and duration of exposure)
Chick embryo Vapor exposure, 10,000 ppm

Chick embryo Injected; 5 - 100 umole per
egg in 25 ul olive oil

S - D Rat 9 Inhalation; 300 ppm,
7h/d on 6 - 15-d gestation
S - W mouse 9




CR - SO rat 9 Inhalation; 300 ppm,
6h/d on 6 - 15-d gestation

Rat 9 Inhalation, 1800 ppm
(Long-Evans) a) Premating: 6h/d, 5d/wk
for 2 wk
b) Premating for 2 wk + first
20-d gestation, daily
c) First 20-d gestation, daily
CR - SD Rat 9 Inhalation; 500 ppm, 7h/d,
Rabbit 9 5d/wk
Measures Results
Mortality Increase
anomalies Slight increase
LD50 50 - 100 umole/egg
(16 % malformations in
total survivors)
Embryo toxicity, +
teratogenicity
(maternal toxicity) (+)




Embryo toxicity, +
teratogenicity
(maternal toxicity) (+)
Embryo toxicity, +
teratogenicity, offspring
behavioral evaluation
(maternal toxicity) (+)


Embryo toxicity, +
teratogenicity
                            Premating  3  wk + first 18-d
                            gestation,  daily (rat)
                            + first 21-d gestation,  daily
                            (rabbit)
(maternal  toxicity)

-------
were recorded on  days  6,  10,  16 and 18 of gestation as an index of maternal
toxicity.   Following caesarean  section  on  day 18, the fetuses were weighed,
measured,  (crown-rump length),  sexed,  and then examined for external  anomalies.
One-half of  the fetuses were examined  for  soft  tissue malformation, using
free-hand sectioning, and one-half  of  the fetuses were cleaned,  stained,  and
examined for skeletal malformations.   One  fetus  in each litter was subjected
to microscopic examination.
     Exposure to  TCI had  no effect on either the mean, absolute, or relative
weight of  the liver of the maternal  animals when measured at the  time  of
caesarean section. Exposure  had no effect on the  average number of implantation
sites per litter,  litter  size,  the incidence of  fetal  resorptions, fetal  sex
ratios, or fetal body measurements.   The incidence of gross anomalies observed
by  external  examination was not  significantly greater than  among control
litters.   TCI exposure  had  no  effect on the incidence of skeletal anomalies,
and microscopic examination revealed  no abnormalities of organs, tissues, or
cells.   These investigators concluded that 300 ppm (1614 mg/m  ) TCI caused no
significant maternal, embryonal  or fetal toxicity, and that TCI was not terato-
genic in Swiss-Webster mice  at this dose level.

6.1.2  Rat
     Schwetz et al.  (1975)  carried out a study in Sprague-Dawley  rats.   The
design of the  study was similar to their protocol for mice described above.
TCI (300 ppm;  1614 mg/m ) was  administered  by inhalation,  7  h daily, to 18
pregnant rats during organogenesis, the period when the greatest susceptibility
to  teratogenic  effect  is  present (days 6 to 15 of gestation).  Controls  (30)
were provided filtered air.   Day 0 of gestation was defined as the day sperma-
tozoa were  observed in  vaginal  smears.  The animals were killed on day 21 of
gestation,  caesarean sections were  performed, and fetuses  in  each  litter  were
examined for  external  malformations.   One-half of the fetuses in each litter
were then  examined for  soft tissue  malformations,  and  the  other  half  examined
for skeletal  malformations.  One  fetus  from  each  litter was randomly  selected
for  serial  sectioning  and  histological  evaluation.   Schwetz et  al.  (1975)
reported  a  small  but statistically significant  inhibition of maternal body
                                                        3
weight gain when  dams were exposed to 300 ppm (1614 mg/m ) TCI.  However, this
dose had  no effect on either the  mean,  absolute, or relative weight of  the
maternal host livers when measured at the time of caesarean section.  Exposure
to  TCI  had  no effect on the average number  of implantation sites per litter,
                                     6-4

-------
litter size, the  incidence  of fetal resorptions,  fetal  sex  ratios,  or  fetal
body measurements.  No increased incidence of soft tissue or skeletal anomalies
was observed, and histopathological  examination revealed no abnormalities of
organs, tissues,  or cells.    Schwetz et al.  (1975)  concluded  that 300 ppm
          3
(1614 mg/m ) TCI  causes  little or no significant maternal, embryonal  or fetal
toxicity in Sprague-Dawley rats.
     Bell  (1977)  reported a similar study, conducted by Industrial  Bio-Test
Lab. Inc.  (Northbrook,  111.),  on the teratogenic  and fetal  toxic  effects  of
TCI on Charles  River  rats.   TCI (300 ppm; 1614 mg/m ) was administered to 17
dams by inhalation  exposure, 6 h per day,  from  day 6 through day 15  of  gesta-
tion.   Control  pregnant animals  (21)  were killed on day 20 of gestation.
Caesarean  sections  were performed and  the number of fetuses,  implantation
sites, resorption  sites, and corpora lutea were counted.  Viable fetuses were
counted,  and all  fetuses were removed and weighed.  External  examination  of
the fetuses was made for gross anomalies.  Two-thirds of the fetuses from each
litter were examined for skeletal development,  and the remaining one-third of
the fetuses were  evaluated  for soft tissue abnormalities.  The investigators
found that 300 ppm (1614 mg/m ) TCI  exposure produced little or no evidence of
maternal  toxicity,  although  a  small  but significant reduction of mean weight
gain by gestation day  15,  in comparison to control dams, was recorded.  There
were no mortalities or adverse behavioral reactions,  gross fetal anomalies, or
differences in mean weights  of fetuses exposed to TCI in comparison to unexposed
controls.   The numbers of corpora lutea, implantation sites,  resorption sites,
and fetuses were  similar among exposed  and  control  dams.  No evidence  for
teratogenicity was  observed  in fetuses  from either exposed or  unexposed dams.
     Beliles et al. (1980)  reported on  a  study of teratogenic  and fetotoxic
potential  of TCI  in Sprague-Dawley  rats conducted by Litton Bionetics,  Inc.
(Kensington, MD.). The experimental  design of the study included pregestational
as well as gestational  exposure (Table 6-2).
                                        3
     Exposure to TCI,  500 ppm (2690  mg/m ) in air, was 7 h/d, 5 d/wk, during a
3-wk pre-gestational period, and 7 h/d each day for days 0 to 18 and days 6 to
18 of  gestation.   Positive  identification  of spermatozoa in  the vaginal  canal
was taken  as evidence  of mating and  designated as day 0 of gestation. The dams
were killed on day  20 of gestation;  caesarean section was performed,  visceral
and thoracic  organs were examined,  and the numbers corpora lutea per dam
determined.  The maternal livers, kidneys and lungs were  weighed.   The numbers
                                     6-5

-------
        TABLE 6-2.   EXPERIMENTAL DESIGN OF BELILES et al.  STUDY (1980)
                            ON SPRAGUE-DAWLEY RATS
Group
1
2
3
4
5
6
No.
21
16
20
16
20
32
TCI
ppm
0 (control)
0 (control)
500
500
500
500
Days of Exposure

(Controls exposed to air only)
Pre-gestational Gestational
None
(5 day/wk) 3 wk
None
(5 day/wk) 3 wk
None
(5 day/wk) 3 wk
0-18
0-18
0-18
0-18
6-18
6-18
of implantation sites.,  live  and dead fetuses, and resorption sites were re-
corded.   The fetuses were  examined for gross external  anomalies and weighed,
and crown-rump length  and  sex were determined.  One-half of  the fetuses of
each litter were  sectioned and examined for changes  in the  soft tissues; the
remaining fetuses  were examined for skeletal abnormalities.
     The findings reported  in this study indicate that maternal toxicity was
not a factor in data interpretation; means of body weight,  liver, kidneys,  and
lungs of test dams were within  normal expected variation of control dams.  No
deaths  occurred  in  the exposed groups.  The  investigators  concluded that
neither the frequency  nor  character of the macro- or microscopic findings in
the treated groups indicated an adverse effect on fetal growth and development,
or a teratogenic  potential for  TCI  at the dose level of 500 ppm (2690 mg/m3).
     Dorfmueller  et al. (1979)  conducted a  study of the teratogenic and  feto-
toxic effects of  TCI in Long-Evans  rats exposed to a inhalation concentration
of 1800 ppm  (9684 mg/m )  in air.  Their experimental  design  was similar to
that of Beliles et  al. (1980) insofar as the animals were exposed during the
pre-gestational period in addition to during gestation (Table 6-3).
     The premating exposure was conducted for 6 hr/day, 5 days/wk for 2 weeks,
and the gestational  exposure was 6 hr/day each day for days  0 to 20.  Day 1 of
pregnancy was considered to be the day on which sperm from mating were observed.
On day  21  of gestation, 15 out  of  30  dams/treatment group  were sacrificed;
maternal livers were  removed, weighed,  and frozen for later enzyme analysis.
Caesarean  sections were performed and the number of live, dead, and resorbed
                                     6-6

-------
      TABLE 6-3.  EXPERIMENTAL DESIGN OF DORFMUELLER  et al. STUDY (1979)
                              ON LONG-EVANS RATS
Group
A
B
C
D
No.
30
30
30
30
TCI
(ppm)
1800 ± 200
1800 ± 200
1800 ± 200
none
Days of
Pre-gestational
5 day/wk; 2 wk
5 day/wk; 2 wk
none
none
Exposure
Gestational
1-20
none
1-20
none
fetuses counted.   Each fetus was weighed, the sex was recorded, and then exam-
ined for external anomalies.  Four fetuses from each  litter were examined for
soft tissue changes, and four fetuses were evaluated for skeletal  abnormalities.
Livers were removed from the remaining fetuses for measurement of mixed-function
oxidase activity.
     Postnatal behavioral  evaluation was conducted on the offspring  of the
remaining 15 dams/treatment group which were allowed to litter.  At weaning, 2
male and 2 female pups from each litter were randomly selected for observation
of activity measurements  at 10  days of  age  (Motility Meter)  and 20 and 100
days of age (Activity Cage).
     Dorfmueller et al. (1979)  found that the female  rats exposed to 1800 ppm
          3
(9684 mg/m ) TCI before mating  and/or during pregnancy did not  exhibit any
signs of maternal toxicity.  Weight  gains throughout  pregnancy were normal  in
relation to  filtered air controls,  and  the  relative and absolute maternal
liver weights indicated that no significant  treatment effects  occurred. There
was  no  indication  of  embryotoxicity;  no significant treatment  effects or
interactions were found in  numbers of corpora lutea  or  implantation  sites/
litter, fetal  body weights, resorbed fetuses/litter, or sex ratios.   No signi-
ficant treatment effects  were, observed in the analysis of total soft tissue
anomalies.   However, the  incidence  of total  skeletal anomalies was increased
in group C because of an increased incidence of incomplete ossification of the
sternum, but when the  frequency of this  particular  anomaly alone was compared
in group C versus group D,  the  investigators found  no significant difference.
Incomplete ossification of  the  sternum is thought to be indicative of delays
                                     6-7

-------
in the  general  process of skeletal ossification  which  are not necessarily
                                                             3
irreversible.   Thus, prenatal exposure to 1800 ppm (9684 mg/m ) TCI caused an
elevation of minor  anomalies  indicative  of development  delays  in maturation,
but not of major malformations.   Also,  the study indicated no treatment effect
on the  general  activity of the offspring  at  10,  20, and 100 days  of age.
There was, however, a  small but statistically significant depression  in post-
natal weight gains  of  offspring  in the premating exposure groups (A  and B),
the  biological  significance  of which was  not  apparent.   The offspring are
being maintained until  18  months  of age  to further assess postnatal behavior
and transplacental  carcinogenicity potential.  This  study revealed  no results
indicative of  treatment-related  maternal  toxicity,  embryotoxicity, terato-
genicity,  or significant  postnatal  behavioral  deficits  from  exposure,
pre-gestational  and/or gestational,  to  1800  ppm (9684 mg/m ) TCI  in air.

6.1.3   Rabbit
     Beliles et al  (1980)  also studied the  fetotoxicity and teratogenicity of
TCI  in  the female New Zealand white rabbit.   Their experimental design is
presented in Table 6-4.
        TABLE 6-4.  EXPERIMENTAL DESIGN OF BELILES et al. STUDY (1980)
                         ON NEW ZEALAND WHITE RABBITS
Group
1
2
3
4
5
6
Animals/ TCI
group ppm
21
18
23
19
23
25
0 (control)
0 (control)
500
500
500
500
Days of Exposure
Pre-gestational Gestational
None
5 day/wk;
None
5 day/wk;
None
5 day/wk;

3 wk

3 wk

3 wk
0-21
0-21
0-21
0-21
7-21
7-21
     Exposure to TCI, 500 ppm (2690 mg/m ) in air, was 7 hr/day, 5 day/wk during
a 3-week pre-gestational period, and/or daily for days 0 to 21 or days 7 to 21
of gestation.  Observed copulation and/or the presence of sperm in the vaginal
canal  established  day 0 of gestation.  On day 30 of gestation, the dams were
killed, caesarean-sectioned,  and the corpora lutea per  dam determined.   The
maternal  livers,  kidneys,  and lungs were removed, weighed, and examined; the
remaining  visceral and thoracic organs  were also examined.  The  number  of
                                     6-8

-------
uterine implantation sites,  live  and dead fetuses, and resorption sites were
recorded.   The fetuses were examined for gross external anomalies and weighed,
and then crown-rump  length and sex ratios were determined.  One-half of the
fetuses of each  litter were sectioned and examined  for  changes in the soft
tissues; the  remaining fetuses were  examined for  skeletal  abnormalities.
     The number  of deaths in  the  TCI-treated groups was not significantly
greater than among  controls.   Deaths occurring during the experiment account
for the  unequal  numbers of  animals  in the groups.  The  mean  body weights
during pre-gestational  and gestational periods were determined,  and there were
no remarkable differences between control and treated rabbits.
     This study provided no  evidence  of  maternal toxicity from  500 ppm  (2690
mg/m ) exposure to  TCI.   Means of body  weight, liver, and lungs were within
normal expected  variation of  control  dams.   A small  increase  in  means of
kidney weights for  groups  2, 4, and  6  (compared  to group 1) was found, but
these increases were judged  to be unrelated to treatment.  No difference was
observed in the mean weight of fetuses exposed to TCI in comparison to those
of unexposed  controls.   The numbers  of  corpora  lutea, implantation sites,
resorption sites,  fetuses and sex ratio were similar among exposed and control
dams.
     Gross examination  of fetuses, with regard to external  morphology and
internal abnormalities, revealed  no  unusual  changes except for hydrocephalus
in 4  fetuses of  2  litters of group 3.  This anomaly occurred only in the one
treatment group 3  and  the malformation was not observed  in  the litters of
                                                3
rabbits that were  exposed  to 500 ppm  (2690 mg/m )  TCI during  a 3-week pre-
gestational  period plus the  same  gestational  period (Group 4).  The authors
confirmed the anomaly as an external  hydrocephalus, which, in their experience,
does  not  occur in  untreated control  rabbits.  They, therefore, felt that,
although not statistically  significant,  this  finding could not be discounted
entirely as occurring by chance.   It  is  not known if the  increase in external
hydrocephalus observed  in  this study is a significant observation of a rare
biological  anomaly.  Additional  studies  would be needed  to  strengthen  this
conclusion.
     Examination  of the fetuses  for  soft tissue anomalies showed no changes
(except external  hydrocephalus discussed  above)  that were  not frequently
observed in 30-day-old  rabbit  fetuses of the strain and  source used in the
study.  Skeletal  examination revealed no  unusual  features, except  for an
                                     6-9

-------
increased incidence of  changes  related to retarded bone ossification, which
were not considered malformations as such.   Neither frequency nor the character
of these changes  in  the treated groups indicated an adverse effect on fetal
growth and development  or of a teratogenic potential.  In summary, this study
in rabbits exposed to 500 ppm (2690 mg/m ) TCI revealed little or no evidence
of maternal toxicity or embryotoxicity; the authors did report the occurrence
of hydrocephalus in a few fetuses of one of the study groups, but no definitive
conclusion was made regarding this occurrence.   A definite conclusion concerning
this effect cannot be determined without additional  studies that would strengthen
this observation.
6.2  FETOTOXICITY IN HUMANS
     There are no epidemiologic studies in the literature specifically investi-
gating fetotoxicity or overt birth anomalies that may occur from TCI exposure
in the workplace.
     Corbett and  his  associates  (1972; 1974) reported an increased incidence
of miscarriages in operating room nurses exposed to various general anesthetics
(volatile halogenated aliphatic  hydrocarbons).   TCI was only one  of several
anesthetics in use, and,  therefore, it is impossible to make a specific associa-
tion of  TCI with the  miscarriages.  Cote  (1975)  has pointed out weaknesses  in
this study due to the lack of direct physician examination of the children and
a  lack of  stringent categorization of the  types of birth defects.  It was
suggested by  Cote that  additional long-term studies be  conducted  to resolve
these issues.
6.3  SUMMARY
     Inhalation exposure studies in the mouse and rat by different investigators
indicate that TCI does not produce significant signs of developmental toxicity
except  at  levels  which affect maternal  well-being.   One  investigator,  Beliles
et  al.  (1980),  reported the  occurrence of an unusual observation in rabbits,
hydrocephalus, in a few fetuses from one exposure group.  However, this observa-
tion was  not dose-related and the significance  of  this  finding could  not be
assessed from this study.
                                     6-10

-------
     At present, the  available  studies do not indicate that TCI is toxic to
the fetus at  levels  below the maternally toxic  level.   Also,  no definitive
clinical evidence of fetotoxicity or teratogenicity from TCI exposure has been
reported.   Therefore, the available information does not  indicate  that the
conceptus is  uniquely susceptible to the effects of TCI.   It should be  noted,
however, that the potential of TCI to produce external hydrocephalus in rabbits
has been discussed by Bellies et al. (1980) and could not  be resolved by that
study.   In addition, TCI  has been implicated in producing  adverse male  repro-
ductive effects (see Chapter 7).
                                     6-11

-------
6.4  REFERENCES


Beliles, R. P.,  D.  J.  Brusick, and F. J. Mecler. 1980. Teratogenic-Mutagenic
     Risk of Workplace Contaminants:  Trichloroethylene,  Perch!oroethylene and
     Carbon Disulfide.  U.S.  Department of Health, Education and Welfare (Contract
     No. 210-77-0047).

Bell, Z. 1977. Written Communication  with contractor  reports  of:  (a)  Dominant
     Lethal Study with Trichloroethylene in Albino Rats Exposed via Inhalation,
     February 9, 1977, and  (b) Teratogenic  Study via  Inhalation with  Trichlor
     132,.  Trichloroethylene in Albino Rats, March 8.

Bernstine,  J.  B. and  A.  E.  Meyer. 1953.  Passage of metabolites  of chloral
     hydrate into  amniotic  fluid. Proc.  Soc.  Exp. Biol.  Med. 84:456-457.

Bernstine,  J.  B. ,  A.  E.  Meyer and R.  L. Bernstine. 1956.  Maternal  blood  and
     breast milk estimation  following the administration  of  chloral  hydrate
     during the puerperium.  J.  Obst.  Gyn. 63:228-231.

Bernstine,  J.  B. ,  A.  E.  Meyer and H.  B.  Hayman. 1954. Maternal  and  foetal
     blood estimation following  the  administration of chloral hydrate during
     labour.  J.  Obst.  Gyn.  61:683-685.

Bross,  G. ,  D.  D. Franceisco,  and  M.  E.  Desmond. 1983.  The  effects of low
     dosages of  trichloroethylene  on  chick development. Toxicol.  28:283-294.

Corbett, T.  H.  1972.  Anesthetics  as  a  cause  of  abortion.  Fert. Steril.
     23:866-869.

Corbett, T.  H. ,  R. G. Cornell,  J.  L. Endres,  and  K.  Leiding.  1974.  Birth
     defects among  children of nurse  anesthetists.  Anesthesiology 41:341-344.

Corbett, T. , G.  Hamilton, M. Yoon, and J. Endres.  1973.  Occupational  exposure
     of operating  room personnel  to  trichloroethylene. Can.  Anesth.  Soc. J.
     20:657-678.

Cote, C. J. 1975.  Birth  defects  among infants of nurse anesthetists.  Anesthe-
     siology 42:514-515.

Dorfmueller, M.  A., S.  P.  Henne,  R.  G.  York,  R.  L.  Bornschein,  and J.  M.
     Manson.  1979.  Evaluation  of teratogenicity and behavioral toxicity  with
     inhalation  exposure of maternal  rats to  trichloroethylene. Toxicol.
     14:153-166.

Elovaara, E., K.  Hemminki, and H. Vainio. 1979.  Effects of methylene  chloride,
     trichloroethane,  trichloroethylene,  tetrachloroethylene and toluene on
     the development of chick embryos. Toxicology 12:111-119.

Fink, R. 1968.  Toxicity  of Anesthetics. Williams andWilkins Co., Baltimore.
     MD, pp.  270.

Goodman, L. S., and Gilman, A.  G. 1980. The Pharmacological Basis of  Therapeutics.
     Ed. A. G.  Gilman, L. S. Goodman  and A. Gilman, 6th ed. New York: McMillan
     Pub. Co., pp.   361-363.

                                      6-12

-------
PDR, Physician's Desk Reference. 1981.  Oradell,  N.Y.:  Medical Economics Co.,
     p.  1258.

Schwetz, B. A.,  B.  K.  J.  Leong,  and P.  J. Gehring.  1975.  The effect of
     maternally inhaled trichloroethy1ene, perchloroethy1ene, methyl chloroform,
     and methylene chloride  on  embryonal  and fetal  development in mice and
     rats.  Toxicol. Appl.  Pharmacol. 32:84-96.

U.S. Environmental Protection Agency.  1980. Determination not  to  initiate  a
     rebuttable presumption  against registration  (RPAR)  of pesticide  products
     containing carbaryl; Availability  of decision  document.  Fed.  Regist. 45:
     81869-81876, December 12.
                                     6-13

-------

-------
                               7.   MUTAGENICITY
     Trichloroethylene (TCI) has  been  tested for its ability  to cause gene
mutations in bacteria, yeast,  higher plants, insects, and  rodents.   It has
also been evaluated for its ability to cause chromosome aberrations in insects,
rodents, occupationally exposed workers,  and in cultured cells,  and  for its
ability  to  cause other effects indicative  of  DMA damage (gene  conversion,
mitotic  recombination, sister  chromatid  exchange,  unscheduled DNA synthesis,
and DNA  alkylation).   The  results of these tests will be discussed below and
summarized in Tables 7-1 to 7-9.  Consideration will  also be given to studies
bearing  on  the  potential  of TCI  to reach the germinal  tissue of mammals.

7.1  GENE MUTATION STUDIES
7.1.1  Prokaryotic Test Systems (Bacteria)
     Eight reports  were  evaluated  that  concerned testing  of TCI for  its
mutagenic potential in bacteria.   Seven used Salmonella typhimurium (Henschler
et a!., 1977; Simmon et al., 1977; Margard, 1978; Waskell,  1978; Baden et al.,
1979; Bartsch et al.   1979; Kline et al., 1982), and  one employed Escherichia
coli (Greim et al., 1975).
     Because of  the possibility  of  mutagenic  contaminants  in TCI samples,
Henschler et al.  (1977) chemically analyzed and subsequently tested samples  of
TCI for  mutagenicity.  They determined that a sample of technical-grade TCI
used in  the  National  Cancer Institute (NCI) carcinogenicity  bioassay (NCI,
1976) contained the  following  impurities (% w/w) by  gas chromatography-mass
spectrometry analysis: epichlorohydrin,  0.22;  1,2-epoxybutane, 0.20;  carbon
tetrachloride,  0.05; chloroform,  0.01;  1,1,1-trichloroethane,  0.035;  diisobuty-
lene, 0.02;  ethylacetate,  0.052;  pentanol-2, 0.015;  butanol,  0.051.   These
impurities were  tested for mutagenicity with and without S9 mix from PCB-induced
Wistar rat livers.   In Salmonella typhimurium,  two impurities, epichlorohydrin
and 1,2-epoxybutane, were  found to  be mutagenic in tests  conducted  without
activation.   Increases of  twenty-five  and eight times over the spontaneous
revertant rates  were observed for the two compounds,  respectively (extrapolated
from Figure  1 of the paper).   Negative responses were observed when the tests
were repeated using S9 mix.  Negative responses were  also obtained with carbon
                                    7-1

-------
tetrachloride, chloroform, and 1,1,1—trichloroethane,  both  with and without
activation, but no data were presented. Similarly, no  increased mutation fre-
quency was obtained with "purified"  TCI (source and purity not given).   Roughly
140 TA100  revertants were  observed  at each point tested over the entire dose
                     -3
range (from 1.12 x 10  M to 1.12  M), both with and without metabolic activation;
the spontaneous revertant count was 138 ± 8 (without activation) and 242 ± 10
(with activation)  colonies  per plate.  In  their  testing,  Henschler et al.
(1977) added the chemical  substances directly to the top agar.   No precautions
were reported to have been taken to prevent excessive  evaporation of the test
compound,  and, thus, the  significance of the negative response obtained with
TCI is  weakened by  the  possibility that the material  may  have evaporated
before  adequately  exposing the indicator cells.   This aspect of the study
design could  not  be  evaluated  because no toxicity data were presented.  The
positive response after exposure  to other volatiles like 1,2-epoxybutane may
reflect the potency of these materials compared to TCI.  Based on their results,
the authors concluded that the  carcinogenic activity of technical-grade TCI in
the NCI  bioassay  experiment  was  probably predominantly, if not exclusively,
due to  the epoxide contaminants.  The  basis for reaching these  conclusions is
inadequate  for a  number  of reasons including flaws in the study design, such
as those mentioned above.
     Margard  (1978)  also addressed  the hypothesis  that the mutagenic activity
of technical-grade samples of TCI is  due  to mutagenic  stabilizers rather than
TCI by  comparing  the mutagenic properties of  technical  and purified grades of
TCI  in  the Salmonella  typhimuriurn/mammalian  microsome plate  incorporation
assay. He  used  tester  strains  TA1535, TA1537, TA1538, TA98, and TA100 in the
presence  and  absence of a metabolic  activation  system (S9 mix) from adult
Sprague-Dawley rats  induced with Aroclor-1254.  Prior to plating, two aliquots
of cells  were taken.   One was  diluted and grown  on complete media,  for  deter-
mining  cell  survival,  and the  other was  grown  on  histidine-deficient plates
for selection of  his  revertants.  Cells were  then incubated  for 48 hours  at
37°C  and colony  counts were made.  The technical-grade sample contained 0.08
percent  (w/w) epichlorohydrin, 0.02  percent N-methylpyrrole,  0.13 percent
N-propyl alcohol, 0.05 percent nitromethane, 0.10 percent 1,1,1-trichloroethane,
and 0.45 percent  butylene oxide.   The purified grade  contained no detectable
epoxides  nor other  stabilizing  ingredients (written  communication with  L.
Schlossberg,  Detrex Chemical  Industries,  Inc.).   A  concentration-related
                                    7-2

-------
increase in the number of revertants was observed, as described in Table 7-1,
in strain TA1535,  both with and without metabolic activation after exposure to
technical-grade TCI.  This  sample  was  not mutagenic in strains TA98, TA100,
TA1537,  and  TA1538.   Purified  TCI  was not mutagenic in  any  strains.   TCI
levels >0.1 ml/plate were toxic, as indicated by reduced cell  viability (49-86
percent  killing).  It  has been reported that precautions were taken to inhibit
evaporation of the  test  samples during treatment of the  cells (Schlossberg,
personal communication),  but it was not specified in writing what those precau-
tions were. The toxicity data suggest that the cells were exposed to the test
material, but  it  is  not  possible to determine toxicity adequately in a plate
incorporation  test.   To  do  this,  a liquid suspension assay must  be done.
Appropriate positive  controls were  reported  and demonstrated that the 59 had
metabolic activity.    From the data presented, it is not clear how many plates
were employed  for each dose nor is it clear  whether or not replicate testing
was performed.   Under the conditions of test, "purified" TCI was  not mutagenic.
     To overcome the concern that the volatility of certain halogenated hydro-
carbons may prevent them  from being adequately tested in standard plate incor-
poration tests, Waskell  (1978) investigated the mutagenic properties of commonly
used volatile anesthetics,  including TCI,  and several  of their known metabolites
in testing conducted in sealed chambers.   Salmonella typhimurium  strains TA100
and TA98 were  used  in tests for gene  reversion.   Anesthetic-grade TCI from
Ayerst was tested in closed containers at doses up to 10 percent  in the atmos-
phere. The purity of  the sample tested was  not  reported, but it would have
been high, with an approximate minimum purity of 99.5 percent and stabilization
by 0.1 percent thymol  (personal  communication with Ayerst).   Holes were punched
in the top of the Petri  dishes to allow the test vapor access to  the bacterial
strains.  Tests were  conducted  in  the presence and absence of phenobarbital-
and PCB-induced Sprague-Dawley  rat  liver microsomes.  No  toxicity testing  was
conducted nor  were  the actual  concentrations of TCI measured in  the exposure
chamber, so it is not known whether the bacteria received a sufficiently high
dose.    After  a 48-hr exposure,  TCI  did  not  produce a  mutagenic  effect in
strains TA100  and TA98,  with or without metabolic  activation.   However,  no
data were presented to support this conclusion.   With respect to  the metabolites
of TCI,  the  nonvolatile  compound 1,1,1-trichloroacetic acid  (>99.0 percent
pure, obtained from Matheson,  Coleman,  and Bell, personal communication with
Matheson, Coleman, and Bell), chloral hydrate (recrystallized from chloroform
                                    7-3

-------
            TABLE 7-1.  MUTAGENICITY OF TECHNICAL-GRADE TCI  IN SALMONELLA/MAMMALIAN
                                       MICROSOME ASSAY
Activation
Toxicity
Tester
Strain
TA1535





TA100





Cone.
(ml/plate)
1.0
0.5
0.1
0.05
0.01
0
1.0
0.5
0.1
0.05
0.01
0
Plate
Count


27
142
185
196


53
207
255
283
Rel.
Via.*


0.14
0.72
0.94
--


0.19
0.73
0.90
—
MUTAGENICITY OF
TA1535





TA100





0.1
0.5
0.1
0.05
0.01
0
1.0
0.5
0.1
0.05
0.01
0


97
172
22
246


175
175
225
317


0.39
0.70
0.90
--


0.55
0.55
0.71
"" ™"
Mutagenicity
No.
Rev.t


58
40
17
24


184
263
123
167
PURIFIED-GRADE


17
19
21
22


70
142
157
154
Nonactivation
Toxicity
Plate
Count


22
122
149
157


66
167
232
234
TCI


89
121
156
175


166
162
354
305
Rel.
Via.*


0.14
0.78
0.95
--


0.28
0.71
1.0
—



0.51
0.69
0.89
--


0.54
0.53
0.82
"
Mutagenicity
No.
Rev.t


58
42
21
23


173
181
136
160



15
20
20
19


114
158
145
158
*Relative viability of tests compared to the control  plates.
tNumber of revertants observed (average).

Source:  Adapted from Margard, 1978.
                                            7-4

-------
six times, to achieve a purity of >99.0 percent), and trichloroethanol (purity
not  determined,  personal  communication with  L.  Waskell,  V.A.  Hospital, San
Francisco, CA) were tested at maximum nontoxic quantities in plate incorporation
tests.   Doses  of trichloroethanol  at 7.5  mg/plate  and  trichloroacetic acid  at
0.45  mg/plate  were  not mutagenic in strains TA98 and TA100.  Chloral hydrate
at 10 mg/plate increased the number of TA100 revertants above control (200) by
121 with metabolic  activation  and 89 without metabolic activation.  The data
are inconclusive but are suggestive of a weak positive response.  These strains
were  responsive  to  the positive-control  chemicals, diethyl  sulfate and
2-aminofluorene.
      Further testing was done to evaluate the dose-response of chloral hydrate-
induced  mutagenesis  in  strain  TA100.  The number  of revertants  per  plate in
the control  groups  were 190 and 210  with  and without metabolic activation,
respectively.  Doses of chloral hydrate from 0.5 mg/plate to 10.0 mg/plate did
not result  in  a  mutation frequency greater than  twofold above the  spontaneous
level.   However,  increases in the number of revertants per plate above control
values were observed both with and without metabolic activation: 255 to 260 at
0.5 mg;  285 to 290 at 1.0 mg; 300 to 320 at 5.0 mg; and 290 to 310 at 10.0 mg.
Although  the author  reported these increases as statistically significant by
the t test  (P <0.01),  the  increases merely  suggest,  at  best,  a  positive
response.  The negative response for the induction of gene mutations in bacteria
induced  by  exposure  to  TCI reported by Waskell (1978) is  consistent with the
results  of  Henschler  et al.  (1977) and Margard  (1978).   However,  equivocal
weak-positive responses suggestive of a positive effect have been  reported  by
other investigators using airtight exposure chambers and metabolic activation.
      In  tests  employing Salmonella tester strain TA100, both with  and without
an exogenous S9 metabolic activation, Bartsch et al.  (1979) assessed the muta-
genic potential of over 20 chemical substances, including  a TCI  sample obtained
from  Merck,  Darmstadt,  Federal  Republic  of Germany.   The  sample was reported
to be 99.5 percent pure and contained no detectable amounts of epichlorohydrin
or 1,2-epoxybutane at levels >I ppm.   Experiments were conducted using a vapor
exposure assay in which Petri dishes  containing a lawn of  TA100  with or without
S9 mix were placed,  uncovered,  in 10-1 desiccators for various lengths of time
at 37°C, in the dark.   Upon completion of the exposure period the compound was
removed, and the plates were then covered, inverted, placed in an incubator
for an  additional 48 hours at 37°C,  and  scored.   TCI  was not  found to  be
mutagenic under the  conditions  used:
                                    7-5

-------
Concentration in Air
(5% v/v)
Time
(hr)
Revertants/plate
+S9 -S9
Trichloroethylene



Vinyl


0
8
20
Bromide
2
20

16
16

16
16
75
85
55 (toxic)

240
715
70
72
50 (toxic)

190
695
When the test conditions were modified, such that plates containing a liver S9
fraction from  phenobarbital-treated OF-1 mice,  an  NADPH-generating system,
TA100, and 0.1 mM EDTA were preincubated at 37°C for 4  hours and then exposed
to 5  percent TCI vapor in air for 2 hours, a nearly twofold increase over the
spontaneous mutation frequency was observed (see table).
Concentration Air
(% v/v)
0
5*
5*
Time
(hr)

2
2t
Revertants/pl ate
+S9
75
100
135
-S9
70
80
70
*0.1 mM EDTA added to soft agar.
tPreincubated with S9, TA100, and 0.1 mM EDTA for 4 hours, then exposed
 to TCI vapors.

The  results  of only one  test  were  presented, but it was  stated that near
twofold  increases over the  spontaneous mutation  frequency were  observed
consistently.
     Baden  et al.  (1979) evaluated  the  mutagenic properties  of TCI in
Salmonella  typhimurium  strains TA100 and TA1535,  both  with and without S9
metabolic activation  from PCB-induced male Sprague-Dawley  rats.   The sample
was  >99.5 percent pure,  contained no detectable epoxides, and was stabilized
with  0.1 percent (w/w) thymol  (personal  communication  with M.  Kelly, V.A.
                                    7-6

-------
Hospital, Palo  Alto,  CA).  Liquid TCI was  directly  added to bacterial cell
cultures for incubation  at  37°C for 2 hours,  and  concentrations in the gas
phase above the  incubator,  which were the same  as  those in the desiccator
experiment described below, were monitored for the  first 2 hours of exposure,
showing a variation of <10 percent.  These cultures were plated  and incubated
for 2 days  at  37°C before counting revertant colonies.   TCI was not found to
be mutagenic in  this  experiment.   However, no monitoring was conducted after
the initial 2  hours,  and it is not known whether the desired exposure levels
were maintained  throughout  the  2-day incubation period.  Toxicity data were
not reported, and, therefore,  it is not known whether  the target cells were
actually exposed to a  sufficiently high concentration  of the test material.
In another  experiment, reported  in the same article, cultures were exposed to
vapor concentrations  ranging from 0.1 to 10.0 percent (± 10 percent tolerance)
in a desiccator for 8 hours at 37°C, with further incubation without treatment
for 40  hours before  counting of colonies.   The  experiment was repeated, and
each test  concentration  was  evaluated  with triplicate assays.   TCI  vapor
produced a  less  than twofold  increase in the number of  revertant colonies, to
approximately 30 percent  above  background  (149 ± 12, n=15) at concentrations
of 1 percent (196  ±  13,  n=15),  and  3 percent  (194  ± 12, n=15) in TA100 with
metabolic activation.   No thymol  control  was reported, and the  response was
not dose-related.  At best this  study  provides  only suggestive supportive
evidence for mutagenic activity caused by TCI.
     Simmon et al.  (1977) also reported that reagent-grade TCI,  with no detec-
table epoxides,  gave  less  than a twofold increase in  revertant counts  in
Salmonella typhimurium strain  TA100.   The  sample was 99.85 percent pure, and
contaminants included 0.09 percent chloroform, 0.02 percent tetrachloroethy1ene,
0.01 percent pentachlorobutadiene, and 0.03 percent hexachlorobutene;  epoxides
were not detected (personal  communication with R. Spanggord, SRI International).
Plates  containing  TA100 with  and without metabolic  activation were exposed to
TCI vapor evaporated from a known amount of liquid in a desiccator for 7 hours
at 37°C  before  counting.   TCI was detected as mutagenic only with metabolic
activation.   The number  of  revertants per  plate  was increased above the  spon-
taneous control  value  of about 140, both with and without activation, by TCI
treatment as follows:   180,  230, 240, and 200 revertants at 0.5, 1.0,  1.5, and
2.0 ml  TCI/91 desiccator, respectively, in the presence of S9 mix from B6C3F1
male mice;  and  160, 210,  200, and  170 revertants at 0.5, 1.0, 1.5, and 2.0 ml
                                    7-7

-------
TCI/91 desiccator, respectively,  in  the  presence of S9 mix derived from PCB-
induced male Sprague-Dawley rats.   The revertant counts were extrapolated from
Figure 22 of  the  paper.   The results of these assays were reproduced several
times (personal communication with V. Simmon, Genex Corp.), but no estimation
of the variance of  the results (stated to be low) was presented and no data
are presented to allow its calculation.
     Greim et al.  (1975) exposed Escherichia coli strain K12 to 3.3 mM analyt-
ical  grade TCI during 2 hours of incubation at 37°C in the presence of metabolic
activation  from phenobarbital-induced male NMRI mouse  liver  microsomes.   A
chemical  analysis  of the test sample was not done by the investigators (written
communication with D.  Henschler, University of Wurzburg); however,  the approxi-
mate composition  of the  TCI product was probably:   TCI, 99.9 percent; either
<100 ppm  diisopropylamine  or £120 ppm of a combination  of diisopropylamine,
N-methylpyrrole,  and  4-tertbutylphenyl  as  stabilizers;  <500 ppm  of other
chlorinated hydrocarbons combined (chloroform, carbon tetrachloride, 1,2-dichloro-
ethane, 1,1,2-dichloroethylene  [sic]  ,  and tetrachloroethylene);  no epoxides
present at  a  detection level of 1 ppm (written communication with Dr.   Fries
and  Dr.  Koppe, E. Merck Company,  the producer of the  sample  used in this
study). Several loci  in  £._ coli K12 were used to test for mutagenicity: back
mutation  at three loci,  gal ,  arg ,  and nad  ,  and the forward mutation MTR
system, which yields resistance to 5-methyl-DL-tryptophan.  Survival of bacteria
on complete medium at  the TCI  concentration used (3.3 mM) was 76 ± 4 percent.
A  weak twofold mutagenic effect  of TCE,  expressed as  a percentage of the
spontaneous mutation  rate,  was observed for arg , 232 ± 36, but the signifi-
cance of  this response cannot  be  ascertained  because  no  increase was  observed
for the other markers; gal  , 123 ± 23; MTR, 114 ± 18; nad , 100.  Descriptions
of the methodology  and the observed  results  were  inadequate.   For instance,
information  on spontaneous  mutation  frequencies,   number  of replications,
revertant count data,  etc.  for each  test was  not reported.   Only one  dose  was
used,  so  it was not possible to obtain  dose-response  data.  Thus, the  signifi-
cance of  the  results cannot be assessed adequately.
     None of  the  tests described previously clearly showed TCI to  be mutagenic.
However,  the  weak,  less than twofold, increases in revertant counts observed
in several  studies,  using several samples  of TCI  and different strains and
species of  bacteria,  suggest that commercially  available  TCI may be weakly
mutagenic  to  bacteria after metabolic activation.   The data for  concluding
                                    7-8

-------
that  TCI  itself is mutagenic are at best suggestive; the fact that weak, but
less  than twofold responses were obtained with such samples only after metabolic
activation argues against the possibility of the responses being caused by the
presence of epoxide stabilizers, because these agents are direct-acting mutagens.
      It is  important  to note  that only very weak positive  responses  (
-------
their data, r was found to equal  0.35, which is not statistically significant.
Furthermore, the number of mutants in the negative controls for the technical
product was much lower,  by 44 percent and 35 percent, respectively, than in
the controls  for the "purified"  TCI and epichlorohydrin.   The  experimental
values for  the  "purified"  samples  were in the  same  range as those for the
technical material, and they showed no increase in mutation frequency,  compared
to the concurrent negative controls.   The mutagenic response to  epichlorohydrin
                                                           -4
tested by  itself was,  however,  positive (3.71 ±  1.05  x  10   mutants at 50
        as  was the  response  of the  positive  control,  dimethylnitrosamine
                  -4
(16.73 ± 3.24 x 10   mutants at 100 mg/kg), indicating that, for these chemicals,
the assay was  working  properly.  No data were presented on the ability of the
administered TCI to  reach  the indicator organisms (e.g., chemical binding).
The negative response in this host-mediated assay obtained for the pure sample
and the equivocal  response for the technical sample in this study may be due
to a number of reasons, including the possibility that TCI (or its metabolites)
did not reach  the  indicator organisms  in the peritoneum under the test condi-
tions employed.
     Subsequent testing by the same laboratory (Rossi et al., 1983) was con-
ducted to  investigate  further the mutagenic potential of pure  and technical
samples of  TCI and two stabilizers contained  in  the  technical -grade  sample,
epichlorohydrin and  1,2-epoxybutane.   Schi zosaccharomyces pombe strain PI  was
used  in i_n  vitro  testing and  in  intraperitoneal  and  intrasanguineous host-
mediated assays.  S9 mix from phenobarbital or B-napthoflavone- induced mice or
rat livers  was used in the j_n vitro studies in which TCI was tested at doses
up to 100 mM.  No increases in the frequency of forward mutations over negative
controls (1.38 ±  0.67  revertants) were found.   By contrast,  epichlorohydrin
and 1,2-epoxybutane  caused dose-related increases in mutant  frequency  up  to
about a tenfold increase,  compared to  negative  controls  at 6.4 mM and 12.8
mM, respectively.
      Negative  responses were  obtained when 2  g/kg doses  of TCI were  used  in
intraperitoneal and  intrasanguineous  host-mediated assays with  CD-I  x  C57BL
mice  (Rossi et al.,  1983).    Negative  responses were also  obtained when
epichlorohydrin and  1,2-epoxybutane were used  in  intraperitoneal host-mediated
assays  at  doses up to 2 g/kg.   These data suggest that the  intraperitoneal
host-mediated  assay  in mice with  Schi zosaccharomyces pombe may not be appro-
priate  for  testing halocarbons and their stabilizers.
                                     7-10

-------
       TABLE 7-2.  MUTAGENICITY OF TRICHLOROETHYLENE AND EPICHLOROHYDRIN
        IN S._ POMBE BY MEANS OF THE HOST-MEDIATED ASSAY IN B6C3F1 MICE
                           (16 hours of incubation)
       Treatment
     Dose
   (mg/kg)
        Mutants
M.F. x 10   (M ± S.E.)
Tri chloroethy1ene,
 technical grade
Methylmethanesulfonate
None (control)
  500
1,000
2,000
   50
      0.94 ± 0.08*
      1.33 ± 0.28
      1.32 ± 0.29
      1.61 ± 0.45
      4.16 ± 0.54
Tri chl oroethy 1 ene,
purified grade

Dimethylnitrosamine
Epichlorohydrin



None (control)
1,000
2,000
100
None (control)
2
10
50
1.45 ± 0.06
1.59 ± 0.22
1.36 ± 0.21
16.73 ± 3.24
1.66 ± 0.03
1.50 ± 1.20
2.05 ± 0.51
3.71 ± 1.05
*Lower than the other negative control values.
Source:  Loprieno et al., 1979.

     Mondino (1979) also reported negative results for purified TCI in a yeast
host-mediated assay.  He used Schizosaccharomyces pombe and adult  male B6C3F1
mice.  The components of the TCI sample were  identified as follows: TCI, 99.9
percent; water,  evaporation  residue,  chloroform, 1,2-dichloroethane, carbon
tetrachloride, 1,1,2-trichlorethane,  other impurities, all £20 ppm; diisopro-
pylamine, 198 ppm; diisobutylene, 104 ppm; butylhydroxytoluene, 12 ppm.   Yeast
cell suspension was injected into the peritoneum of the mice,  who had received
500, 1000, or  2000  mg/kg dosages of  TCI  by gavage.  Three animals comprised
each treatment group as  well  as the  untreated  control  and positive control
(100 mg/kg of  dimethylnitrosamine)  groups.   After a 16-hr incubation period,
the  animals were  sacrificed  to  remove yeast  cells  from the peritoneum for
plating and counting.  Plates were  incubated at 32°C for 4 to 5 days before
                                                   ~4
mutant colony counting.   Mean numbers (±  S.D.) x 10   of mutant colonies were
as follows:  control, 1.80 ±  0.48;  low dose, 1.18  ± 1.18; mid-dose,  1.03 ±
                                    7-11

-------
0.91; high-dose, 1.84 ± 0.28; positive control (dimethylnitrosamine 100 mg/kg
31.1 ±  3.31.   As was the case  for  the studies above,  the  negative  result
obtained by Mondino (1979) may result from TCI not being mutagenic, or it may
be that the compound  or  its  metabolites did not reach  the test organisms in
the peritoneum of the test mice.
     The mutagenic properties of technical-grade TCI  in the XV 185-14C haploid
strain of Saccharomyces cerevisiae  were  assessed i_n vitro by Shahin and Von
Borstel  (1977).  The  ability of TCI to induce mutations at the lys 1-1, his
1-7, and horn  3-10  loci was assessed in this study.  Stock cultures were incu-
bated in liquid YPG medium at 30°C until  a stationary growth phase was achieved.
                                                                    o
Cells were harvested, washed, and adjusted to a population of 1 x 10  cells/ml.
In order to estimate  its mutagenic effect in the absence of metabolic activa-
tion, cells treated with 1.11, 5.56, 111.1, or 222.2 mM of TCI during incuba-
tion at 30°C were withdrawn at 2, 6, 24,  and 48 hours for plating and counting
of colonies.  No precautions were reported to have been taken to prevent exces-
sive evaporation of the compound and ensure exposure to the test organism,  but
the reported toxicity indicates this was not a problem and that the cells were
adequately exposed.   Treated cells  were compared with  cells  from similarly
maintained untreated  control and positive control experiments  (222.2  mM of
2-acetylaminofluorene for tests  without  metabolic activation and 8  ug/ml  of
ethyl methanesulfonate  for  tests with and without  metabolic activation).
Plates were maintained at 30°C for 5 days to estimate survival, and the assays
were performed in triplicate.
     In the  study described  above,  concentrations of  111.1  mM and 222.2 mM
TCI, in the presence  of metabolic activation, increased the reversion frequencies
at the  lys 1-1,  horn 1-10, and his 1-7 loci (Table 7-3)  but only at levels that
were also highly toxic (>99  percent cell death).  The revertants/10  survivors
for  lys 1-1,  his 1-7, and horn 3-10, corresponding to 1-hr treatments  of 111
and  222 mM,  were 2552 and >23,200,  3553  and >29,000,  and 5313 and >25,000,
respectively,  compared  to negative  control  values  of 15.3,  19.5,  and 8.6.  The
significance  of  the  reported increased incidences of mutations in tests conduc-
ted  with  metabolic activation is uncertain, because the increases were noted
only when  there was greater  than 99 percent killing.   The positive control in
the  tests  with metabolic activation (EMS)  was  inappropriate because it is a
direct-acting mutagen.   TCI  without metabolic activation was ineffective as a
mutagen at all doses tested, but increased numbers  of revertants were noted
                                     7-12

-------
        TABLE 7-3.   MUTAGENICITY OF  TECHNICAL-GRADE  TCI  AT THE HIS 1-7
                       LOCUS OF SACCHAROMYCES CEREVISIAE

Dose
(mM)
0
1.11
5.56
111.1
222.2
Positive control
-S9
(6-hr
incubation)
Revertant
14.3
16.7
22
0
0
83*
Percent
Survival
100
90
72
<0.13
<0.013
61
+S9
(4-hr
incubation)
Revertants
19.2
15.9
16.8
22,000
>29,000
6300t
Percent
Survival
99
72
65
<0.1
<0.1
0.23
 2-acetylaminofluorene at 20 ug/ml.
'''EMS at 8 pi/ml-
Source:  Adapted from Shahin and Von Borstel,  1977.
with activation.  With respect to the 48-hr treatment without activation, the
maximum number of revertants per 10  survivors for the 3 loci individually was
10 to 15  for untreated controls (98 percent survival),  12.8 to 21.4 for 1.11
mM (73 percent  survival),  11.4 to 17.1 for 5.56 mM  (64 percent survival), and
0 for 11.1 mM and 222.2 mM (<0.1 percent survival) of TCI.   The lack of rever-
tants with the  two highest concentrations appears to correspond to  failure of
the cells to survive after plating.  The largest  number of  revertants per 10
survivors produced by 2-acetylaminofluorene (222.2 mM) was 102 (lys 1-1) after
a 6-hr treatment, and 673  (his 1-7) and 488 (horn 3-10) after a 24-hr treatment.
It may be that there is a  "window" of effective concentrations in which TCI is
mutagenic but not  excessively toxic.   Such a  concentration effect  cannot be
ruled out, because levels  between 5.56 mM (ineffective) and 111.1 mM (suggestive
response but highly toxic) were not tested; it was in this dose range that the
weak positive responses  discussed below were  obtained  by Bronzetti et al.
(1978) and Call en  et al.   (1980) in strain D7.  It  is important to  note that
the composition  of the TCI sample used in the  study by  Shahin and Von Borstel
(1977) was not  determined, and the possible  contribution of contaminants to
                                    7-13

-------
the response cannot be  ascertained  or ruled out (personal  communication with
R.C.  Von Borstel,  University of Alberta).   These results do not allow a conclu-
sive evaluation to be  made.
     An increase  in point mutations and gene conversion by  a  certified ACS
reagent grade of TCI in Saccharomyces cerevisiae strains D4 and D7 was observed
by Bronzetti et al.  (1978)  in cell  suspension and in host-mediated assays.  The
authors did  not report a detailed chemical analysis of the sample; however, a
description  of  the  product obtained  by personal  communication with Fisher
Scientific Company, the producer of the test sample,  indicates  a purity of
>99.5 percent and stabilization with  either 25 ppm diisopropylamine or 20 ppm
triethyl amine.   In the cell-suspension experiment, strain 07 cells, which can
detect mitotic gene conversion  at  the trp locus, point (reverse) mutation at
the ilv locus, and  mitotic  recombination  at the ade locus, were incubated in
the presence of TCI with  or without  S9 mix derived from livers of uninduced
male CD-I  mice  for  4  hours  at 37°C  before  plating  on selective  media  (for
counting  of  ilv revertants, ade recombinants,  and  trp convertants) and on
complete  media  (for survivor counts).  The  results  for ilv revertants  are
presented  in Table  7-4.  The other results will  be  discussed  later, in the
section on  tests  indicating DMA damage.   In the  host-mediated assay,  yeast
cell culture was  instilled  into the  retroorbital sinus  of male  adult  CD-I
mice.  Two  experiments  were included in this assay:  an acute  experiment, in
which a single gavage  dose  of 400 mg/kg was given immediately  after  introduc-
tion of strain  D4 or  D7 cells on the day of the assay.  Mice were sacrificed
4  hours  following introduction  of  the yeast cells,  and  liver,  lungs, and
kidneys were removed  for homogenization and subsequent plating of yeast cells
as  described for the  suspension assay.  Assays  in this study  were  done in
triplicate or quadruplicate, and the data for gene revertants are presented in
Table 7-4  as mean ± S.E.
     The  results  of Bronzetti  et al. (1978) were as follows.  In the absence
of metabolic activation, 10 mM and 40 mM  concentrations  of TCI  effected a
dose-related reduction in ce.ll viability in the cell suspension system but did
not  induce a mutagenic response.   Survival was  reduced to approximately 87
percent with the  10-mM dose and to 65 percent with the 40-mM dose, compared to
100 percent  for the controls.  TCI did produce concentration-dependent  increases
in  point  mutation frequencies with metabolic activation  at doses of 10, 20,
30,  and 40 mM.   Positive responses were  also  observed in the host-mediated
                                    7-14

-------
       TABLE 7-4.   MUTAGENICITY OF REAGENT-GRADE TCI AT THE ILV LOCUS IN
                           SACCHAROMYCES CEREVISIAE
Suspension Tests Strain D7
Concentration
(mM)
Revertants/106
(Mean x ± S.E.)
Percent Survival
Without Metabolic Activation
0
10
40
With Metabolic Activation
0
10
20
30
40
Host-Mediated Assay in CD-I
Strain Organ
Acute Exposure (400 mg/kg)
D7 Liver
Lungs
Kidneys
0.75 ± 0.12
0.85 ± 0.15
0.87 ± 0.06

0.85 ± 0.06
1.16 ± 0.16
1.77 ± 0.12
2.24 ± 0.17
3.23 ± 0.32
Mice
No. of
Treatment Mice

Control TCE 3
4
Control TCE 2
3
Control TCE 2
4
100
87.5 ±1.0
64.1 ±2.1

100
81.6 ± 2.9
67.1 ± 2.8
55.5 ± 1.7
50.3 ± 1.4

Revertants/106
x ± S.E.

0.08 ± 0.04
0.56 ± 0.52
0.08 ± 0.002
0.40 ± 0.19
0.21 ± 0.02
0.79 ± 0.37
Source:   Adapted from Bronzetti  et al.,  1978.
                                    7-15

-------
assays.   Fourfold  to  eightfold  increases  were obtained  in  the number of
revertants compared to  the  negative  controls.   The  highest response was
Observed in  yeast  recovered from the  kidney,  in  which 0.79 ±  0.37  x 10
revertants were found in the treated animals versus 0.21 ± 0.02 x 10   in the
negative controls.
     Like Bronzetti et  al.  (1978),  Call en et al.  (1980) also employed a TCI
sample obtained from the  Fisher Scientific Company.   They used Saccharomyces
cerevisiae strains D4 and  D7 to assess TCI's genetic activity  and strain D5
intact yeast cells  or  microsomal suspensions to study its metabolism indirectly.
Gene conversion at the  trp 5 locus, mitotic recombination at the ade  2 locus,
and gene reversion of the ilv 1 locus (Table 7-5) of strain D7 were assessed
by exposures of the cells to 0,  15, and 22 mM solutions  in suspension at 37°C
for 1 hour  followed by  plating on appropriate media, allowing for expression
of genetic  activity  and,  finally, scoring.  A  twofold increase in reverse
mutations at the ilv 1  locus over the  spontaneous control values was  observed
at 15 mM  (7.4  vs.  3.6  revertants x 10   survivors,  respectively) without an
exogenously supplied metabolic  activation  system  (67 percent cell  survival).
A severely toxic effect was observed at 22 mM.  A positive  response  without
activation was observed by  Call en et al.   (1980) but  not by  Bronzetti et al.
(1978).   The reason for this is not  known,  but it  may be due to the growth
stage of the yeast cells  when  tested.  Call en  et al.  (1980) found TCI to be
mutagenic to cells from log phase cultures but not  to cells from stationary
cultures of 07. Bronzetti  et al. (1978) did not report the growth phase of the
cells used  in  their test,  but if they employed stationary  phase  cells, a
negative response without activation would be expected.
     Callen et al.  (1980)  provided additional suggestive evidence that Saccharo-
myces can metabolize  TCI.   They analyzed whole-cell suspensions of strain D5
in the log phase of growth in a  spectrophotometer, after exposure to  unspecified
amounts of  TCI.   Peaks  with maximums  of 450  and  505 mM were observed.  The
peak  at 450  nm appeared to  reach a maximum after  roughly 5 minutes,  and  after
this  period  an additional  peak at  423 nm was  observed.   This spectrum is
similar to  that  observed  when  TCI is  added to mammalian microsomes.  These
results  indicate  that  TCI  enters exposed  yeast  cells and suggest TCI  is
metabolized  by a  yeast cytochrome P.5Q-independent monooxygenase system.
                                    7-16

-------
              TABLE 7-5.  MUTAGENICITY OF TCI AT THE ILV LOCUS IN
                          SACCHAROMYCES CEREVISIAE D7

Survival
Concentration Total
(mM) Colonies
0
15
22
1291
802
3
Percent
199
67
0.3
ilv 1
Total
Revertants
43
59
Revertants
106 Survivors
3.6
7.4
Source:  Adapted from Callen et al., 1980.

     The yeast  studies  show that commercially available TCI  is mutagenic  in
Saccharomyces cerevisiae.   The data are consistent with TCI metabolites  being
mutagenic, because  the samples  tested are represented as  being  free from
epoxide stabilizers (which are direct-acting mutagens), and metabolic activation
appears to be required for the genetic activity of TCI.  The doses required to
elicit twofold  to  fourfold  increases  in gene reversion (i.e., 10 to 40 mM in
suspension tests and 400 mg/kg in the host-mediated assay) show the commercial
samples of TCI  to be weakly active.  The  negative responses obtained with  the
Schi zosaccharomyces pombe host-mediated assays are not considered to diminish
the positive results in Saccharomyces because of the inherent insensitivity of
the S_. pombe test (fprward mutation assay with high spontaneous control levels
and no selection scheme).
7.1.2.1.2  Angiosperms.   Testing for  somatic  cell  mutations  in the stamen
hairs of Tradescantia clone 4430, an interspecific hybrid (T.  subacaulis x T.
hirsutiflora),  was  done  by  Schairer et al.  (1978)  to assess the mutagenic
potential of TCI.   A group of unrooted fresh cuttings containing young inflores-
cences with flower  buds in  a range of  developmental  stages was exposed to  0.5
ppm TCI  for 6 hours.  Following treatment, the cuttings were  grown  in aerated
Hoagland's nutrient solution and analyzed each day as they bloomed for 3 weeks
after treatment.   One  hundred forty-eight mutations  were  observed out of
44,000 stamen hairs scored.   The mutation  frequency  (± S.E./100 cells), after
the negative control values  (not reported) were subtracted,  was reported to be
positive (0.112 ± 0.036;  P  <0.01).   Based on  the  data given, however, this
                                    7-17

-------
value cannot be verified.   The  mutation frequencies (± S.E./100  cells)  for
6-hr exposures to 5  ppm EMS and 75 ppm vinyl  chloride  were 1.012  ± 0.133 and
0.112 ± 0.046,  respectively.  Because the purity and source of the sample were
not provided, it is  not possible to ascribe with certainty the observed positive
effects to the genotoxic action  of TCI.
7.1.2.2  Animals.
7.1.2.2.1  Insects (Drosophila).  Abrahamson and Valencia  (1980) fed 1000 ppm
(5380 mg/m ) TCE  in  0.2 percent sucrose for 72 hours to male Canton-S (wild
type) Drosophila melanogaster,  and subsequently mated 102  of these flies with
female FM6 (a marked and balanced X-chromosome) flies,  screening the F_ descen-
dents for mutagenicity using the sex-linked recessive lethal test.   Ninety-six
of the treated males  were fertile.   Administration of the  test compound was
performed in a glass  shell  vial containing  filter material  saturated with the
feeding  solution, which was renewed twice  daily.   However, because of the
volatility of TCI,  the actual  dose given to the  flies  might have been much
less than expected  in the feeding study.  Treated and untreated control males
were mated  individually with 3 females each  and then  transferred without
anesthetization to  fresh  females at 2-  to 3-day intervals,  until a total of 4
broods of progeny were produced.   The fertile females  were left in the vials
for 1 week.  In a  simultaneous experiment, 0.3  ul of TCI was injected into the
abdomen of the male flies before mating.  A technical-grade sample from Allied
Chemical Co. was  used, and the purity and composition  of  the sample were not
ascertained (personal communications with L.Valcovic, Food and Drug Administra-
tion, sponsor of the "study; and Allied Chemical Co.).
     TCI was not mutagenic  in the above test.   The frequency of lethals (number
of  lethals/number of  chromosomes  tested) was  0.12  percent (19/9248)  in the
feeding  experiment,  0.17  percent  (4/2344)  in  the injection experiment,  0.31
percent  (31/10,009)  in concurrent controls, and 0.24 percent (230/94,491) in
the  historical controls used in the testing laboratory for 2 years prior to
completion of this study.   The  negative result obtained in this test indicates
that either  the test substance  did  not  reach the target site  (testes)  or that
TCI  is not mutagenic.
     Beliles et al.  (1980) also tested TCI in Drosophila for its  ability to
cause  sex-linked  recessive lethals  and, in  addition, for  its  ability to  cause
X  or Y chromosome loss.  Threehundred  one-day-old  B Y y  ac ,  (K  1+2)  /X.Y  ,
INEN2,f  males  were  placed  in  each of  three  screen cages and subsequently
                                    7-18

-------
                                                      3
exposed to 0,  100,  and  500 ppm (0,  538,  and 2690 mg/m ) TCI vapor from North
Strong (estimated to  be 99.9 percent pure  by  infrared spectroscopy) for 7
hours.   An unspecified  number  of  males  serving as positive controls were fed
0.025 M EMS in 1 percent sucrose for 8 hours.   Two hundred males per treatment
                                                                  8    8
group were  individually mated  to  sequential  groups  of 3 In(l)sc , sc  1
              Ml          Ml
B/In(l)dl-49 B  ,  y  ac v B   virgin  females.  The brooding scheme consisted of
a 2-3-3-2  day mating sequence, enabling specific sperm  cell stages  to be
tested (i.e.,  spermatozoa,  spermatid,   spermatocyte,  and  spermatogonia).
Losses of the  entire  X  or Y chromosome  and the short or  long arm of the Y
chromosome were detected by  scoring for  specific phenotypic classes in  the F..
                                                Ml
generation.   F-. males of  the genotype y ac v  B  /In(l)EN2,f  were mated to
        S    L
virgin Y /X.Y  , In(l)dl-49,  v  f B females, and  the  occurrence  of recessive
lethal mutation on  the  In(l)EN2,  f  chromosome was detected by  the lack of
males with forked bristles  in  the F2 generation.  The  frequencies of X or Y
chromosome loss were 0.07 percent and 0.17 percent for the low and high doses,
respectively, compared  to  0.06  percent  for the  concurrent  negative control.
These values are  not  significantly  different; however,  the 0.17 percent was
close to a  significant  value.   The  positive control  EMS did  not induce sex
chromosome loss  under these test conditions;  thus,  it  is  not  possible to
determine whether the test  was  conducted properly.  The  frequencies of  losses
of the long arm of the Y chromosome, 0.03 percent and 0.05 percent for the low
and high doses, respectively, and losses of the short arm of the Y chromosome,
0.02 percent  for  both dose levels,  were not significantly different from the
0.01 percent value of the concurrent negative control for losses of both arms.
Like the study  by Abrahamson and Valencia  (1980), Beliles et al.  (1980) found
TCI to be  ineffective in causing an  increase  in the  incidence of  sex-linked
recessive lethal mutations.  The frequencies were 0.03 percent and 0.16 percent
for the low and high doses, respectively, compared to 0 percent for the concur-
rent negative  control and 0.11 percent  of  the  historical negative control.
The frequency of sex-linked recessive lethal mutations for the positive control
(EMS) was 21.56 percent showing a significant positive response.   No significant
variations were observed  among the  four germ cell (brood) stages employed in
any of the assays.
7.1.2.2.2  Rodents.    TCI  was tested in  a  mouse  spot test by Fahrig (1977).
Virgin females  of the inbred C57BL/6J Han strain (a/a, wild type) were mated
with males of  the Han-rotation bred T-stock  (a/a=non-agouti, a/a=brown, c
                                    7-19

-------
p/cc  p=chinchilla and pink-eyed dilution;  d se/d se=dilute and  short  ear;
s/s=piebald spotting) to yield  embryos  of the genotype a/a,  b/+,  c   p/+ +,  d
se/+ +,  s/+ (black coat,  dark eyes).   Each male was mated with  two females.  On
day 11 of  gestation,  a  single intraperitoneal (i.p.) injection of TCI (99.5
percent pure) was  made  into  50 and 26 females at doses of 140  and 350 mg/kg,
respectively.   Impurities detected  in this test sample were 60  to  150 ppm
ethanolamine,  0.0003 percent chloride (C12), and 0.0001 percent free chloride
(Cl-1) (personal  communication with  R.  Fahrig,  University  Tubingen,  West
Germany).  Litters were born on day  21 of gestation, and pups  were examined for
color spots, the  mutagenic  endpoint,  twice weekly between 2 and 5  weeks of
age.
     The number of offspring (number of litters) surviving for  examination was
144 (23)  in the  untreated control  group, 145 (38) in the low-dose group, and
51  (13) in the  high-dose group.  The number of  coat color (brown or gray)
spots found were  none in concurrent control pups,  2/794 historical  control
pups (0.25 percent), 2 low-dose pups (1.3 percent), and 2 high-dose pups (3.92
percent).   Known mutagens,  including ethyl  methanesulfonate, were  positive (26
percent at 0.9  mM EMS).   Results of  this  study  indicate  that  the TCI sample
tested or one or  more  of  its metabolites is  able  to  produce  mutations in
somatic tissue of  an intact mammal.
     Based on the  weak positive responses obtained in Saccharomyces, Tradescantia,
and mice,  commercial  samples of TCI are judged to be capable of  causing gene
mutations  after metabolic activation.  Because of the requirement of activation,
it  is  not likely  that epoxide  stabilizers  present in  the commercial  samples
account for the effect.   However,  because high doses of  commercial-grade TCI
were  required  to elicit a positive effect, the material  must be  considered  to
be  only weakly mutagenic.  The  data for  purified samples  of TCI do not  conclu-
sively show it to  be  positive.  However, weak, less  than  twofold  increases are
reported  with metabolic  activation at high doses.
 7.2  CHROMOSOME ABERRATION  STUDIES
      TCI has been  examined for  its ability to cause  changes  in chromosome
 number (chromosomal loss tests with Drosophila, micronucleus test in  mice,  and
 hypodiploid cells  in humans), and structure (in studies employing mice, rats,
 and humans).  The  results  of the chromosome loss test were described in the
 previous section (Bellies et al., 1980).   The  rest of  the studies are presented
 next.
                                     7-20

-------
7.2.1  Experimental Animals
7.2.1.1  Bone Marrow.   Bellies  et al.  (1980) tested  TCI  for its ability to
cause chromosome  aberrations  in  rat  bone marrow  cells.  Adult male  and  female
rats [CRL:  COBS  CD (SD) BR] were  exposed  to 0,  100, or 500 ppm (0, 538, or
2690 mg/m ) TCI  by inhalation, both  acutely  and  subchronically.  The positive
control was TEM at 0.3 mg/kg given i.p. in 0.85 percent saline.   Acute exposure
consisted of a single 7-hr session in the exposure chamber. Subchronic exposures
consisted of 5 daily exposures of 7 hours.   Bone marrow cells were then collec-
ted; slides were  prepared  and coded, and 50 cells/animal  were scored.   There
was no  increase  in the  number of aberrations  in  either  study.  The  percentage
of cells with aberrations were 0.4 to 2.4 in the negative control group compared
to 0 to 2.7 in the TCI-treated group.  The positive control TEM was clastogenic,
causing 7.2 to 8.5 percent  aberrant  cells.   TCI  failed  to  cause  the induction
of aberrations or aneuploidy  under the conditions of this study; however, it
is not  considered to be  an  adequate  test of  the  clastogenic  potential of  TCI.
The positive control was not appropriate, because it was given by i.p.  injection
and TCI was administered via inhalation.   In addition, the doses used may have
been too low to detect weak activity.  The lowest concentration in air reported
to cause death  in the  rat  is 8000 ppm/4 hour (American Industrial Hygiene
Association, 1969).   It would have  been appropriate  to have tested higher
doses of TCI and to have incorporated a volatile and preferably a chemically
related substance as the positive control.
7.2.1.2  Dominant  Lethal Test.   The  precise  nature of damage-causing dominant
lethal  effects is not known, but there is a good correlation between chromosome
breakage in germ  cells  and  dominant  lethal effects (Matter and Jaeger,  1975).
When dominant lethal effects  are observed, it can be concluded that the test
substance reaches the gonads and likely causes genetic damage.
     Slacik-Erben  et al. (1980)  assessed the ability of 0,  50,  202, and  450
ppm (0, 269, 1086, and 2421 mg/m ) TCI in air to cause dominant lethal  mutations
in NMRI-Han/BGA mice.   The  sample contained 99.5 percent TCI stabilized with
100 mg/1 triethanolamine.   Analysis  of a liquid sample also identified 5 ppm
1,2-dichloroethylene, 9  ppm cis-l,2-dichloroethylene, tetrachloride,  10  ppm
1,1,2-trichloroethane,  and  815 ppm tetrachloroethylene.  Fifty males per  dose
were exposed for 24 hr in a 1000-1 volume glass chamber equipped with a ventila-
tor.  The temperature was kept between 22-24°C and the humidity between 70-80
percent. For  rapid equilibration  the  calculated volume of  liquid  TCI  was
                                    7-21

-------
placed in the  chamber,  and mixtures of TCI and air were streamed through the
chamber at a rate  of 200 (50 and 202 ppm) and 300 (450 ppm) 1/hr.   After the
24-hr exposure  period, the male mice were mated individually with one untreated
NMRI-Han female for  days.   At this time,  the mated females were replaced by
fresh virgin females  who were  kept  for 4  days and then  replaced  (and this was
repeated, etc.)> until  each male had been mated  singly to 12 females for 4
days each (up to 48 days after treatment).  By this mating regimen, all stages
of  spermatogenesis  were sampled, from mature  sperm  to gonial cells.  Each
pregnant female was sacrificed by cervical dislocation, dissected, and examined
for dead and living implants.   The results of  the  testing are presented in
Table 7-6.  Under  conditions  of  this  test, TCI  was not  detected  as mutagenic,
but there are several important points to consider with respect to this finding.
The males exhibited no  visible effects during  the exposure,  and doses up to
450 (2421 mg/m3) ppm did not  alter the  fertilization  rate.   Thus, the doses
selected for test may not have approached the maximum tolerated dose.  Even if
the maximum tolerated dose was approached, the dominant  lethal  test is not
recognized to  be a sensitive  test  for detecting mutagens  (Russell  and  Matter,
1980), due to the high  spontaneous  level  of dominant lethal events.
7.2.1.3  Micronucleus Test.   Duprat and Gradiski  (1980) used TCI in a micro-
nucleus  test.   CD-I mice were divided into 9  groups, 6 of which  were given  up
to  3000  mg/kg  TCI  (analytical grade  99.5  percent purity;  unspecified source)
in  gum  arabic  orally in two dosings,  separated by 24 hours.   The animals were
sacrificed  16  hours  later;  bone marrow  smears were made and stained,  and
polychromatic  erythocytes  (PCEs) and mature erythrocytes  (MEs)  were examined
for the  presence  of  micronuclei.   The  incidence  of micronuclei  in  PCEs was
higher  in  the  vehicle control group (i.e., treated with 10 percent gum arabic
solution)  than in  the untreated control  (1.0  ±0.5 versus 0.5 ±  0.4 percent).
Also,  the  incidence  of what  were  considered micronuclei  in  PCEs and MEs  from
treated  animals was  greater than  the vehicle  control  levels  at  all  doses and
showed a dose-response  effect.   At the  highest  dose (3000  mg/kg) the percentage
of  PCEs  with "micronuclei"  was 15.8 ± 5.7 compared to  the  vehicle  and  untreated
control  values presented above.  The  corresponding values  for "micronuclei"  in
MEs were 3.4 ± 1.8,  0.2 ± 0.3, and 0.3 ± 0.3,  respectively (P <0.01).  Because
of  their occurrence  in MC,  the micronuclei  observed by  Duprat  and  Gradiski
 (1980)  are  likely  artifactual  (see discussion  by  Schmid,  1975).
                                     7-22

-------
                      TABLE 7-6.  MUTAGENICITY OF TCI IN THE DOMINANT LETHAL ASSAY, USING NMRI MICE






1
ro
GO




Dose
in ppm
(rate)
(200 1/h)
0
50
202
(300 1/h)


0
450
*Dominant
F. = 1-
Female
Total
Mated % w/Implants Implants

498
482
493



503
488
lethal factors (FL).
(live implants per

89.6
90.1
91



93.4
92.4

female

5578
5360
5703



6142
5927

of test group)
Total
Dead
Implants

443
479
531



502
533

x 100
Total
Live
Implants

5135
4881
5172



5640
5394


Live % Dominant
Implants/ Lethal
Female Factors*

11.5
11.2 2.6
11.5 0



12.0
11.97 0.25


Source:  Adapted from Slacik-Erben et al., 1980.

-------
7.2.2  Humans
     Konietzko et al.  (1978)  studied the incidence of chromosome aberrations
(structural and numerical) in peripheral lymphomycytes of 28 chemical workers
engaged in the manufacture of TCI.   The workers ranged in age between 23 and
67 years,  with an  average age of 42.5.  Their  exposures  ranged in duration
from 1 to  21 years; the average length of exposure was approximately 6 years.
Since many of the workers had been exposed over a period of many years, it was
not known to what other chemical  substances they may have been exposed.  Thus,
the study  could only address the long-term effects associated with working  in
a factory  that manufactured  TCI.   A control group of 10  healthy young men
working in the Institute of Human Genetics, where the study was conducted, and
ranging in age from 15 to 45 years  (average age 28.3  years), was selected for
comparison.  However, this group  was not a matched control group.   The exposed
group had a frequency of 10.96 ±4.4 percent (X ± SD) hypodiploid cells, which
was significantly elevated compared to the control group, 6.5 ± 3.2 percent
hypodiploid cells.  Concerning chromosome  breaks/100 cells,  the frequencies
were 3.1 ± 3.6  and 0.6 ± 0.69, respectively.  Because the control group was
not a matched control group and the mean frequency of chromosome breaks observed
in the exposed workers was within the normal range, the increased incidence of
chromosome breaks was not considered significant by the authors, although they
did consider  the  increase in  hypodiploid cells significant.  Because of this,
it was  thought appropriate  to  further investigate  factors which  may have
contributed to the  increased incidence of hypodiploid cells.   Based  on the
frequency of hypodiploid cells in the controls, it was calculated that indivi-
duals possessing  >13  percent hypodiploid cells were  abnormal.  Workers were
divided according  to  whether or  not they  possessed  >13  percent hypodiploid
cells, and these groups of workers  were  subsequently further characterized
(Table 7-7).   Nine  individuals had  greater than  13 percent hypodiploid cells.
These individuals  worked  under  higher daily peak  concentrations  of TCI (as
measured on  three occasions  over  a  2-week  period  for 8 hr each  time)  than did
their fellow  workers.  The average daily concentrations for these 2 groups of
workers were  75  ± 14 ppm versus  61 ± 23 ppm, respectively.   These exposure
levels were used as estimates of previous years of exposure.   This was thought
to be a good approximation of the actual exposures received, because no techni-
cal  changes  had  been made in the plants.  This report may provide suggestive
evidence that the exposures  resulting from the  manufacture of technical-grade
                                    7-24

-------
              TABLE 7-7.  NUMERICAL CHROMOSOME ABERRATIONS IN TCI WORKERS
   Parameter
                         >13% Hypodiploid Cells
                     	(n* = 9)	
                     Mean Value     Standard Error
 (X)

    >13% Hypodipoloid Cells
	(n* - 18)	
Mean Value     Standard Error
   (X)              (SJ
Daily average
 concentration
 (ppm)

Average maximum
 concentration
 (ppm)
 75.0
205.8
17.4
96.3
   61.4
  115.7
*n = Number of individuals studied.

Source:   Adpated from Konietzko et al.,  1978.
23.1
 57.6
Exposure 5.9
(years)
Total dose
(ppm x day) 0.55 x 106
Age 38
(years)
Alcohol 18.9
(g/day)
Nicotine
(cigarettes/day) 12.8
4.8 5.9 5.1
0.72 x 106 0.31 x 106 0.39 x 106
9.4 40.4 9.7
29.3 33.6 37.5
12.7 10.1 10.2
                                         7-25

-------
TCI may cause chromosomal loss or nondisjunction in humans.   However,  the lack
of an appropriate matched control group and the possibility that the incidence
of hypodiploid cells was due to  preparation of the chromosomes rather than to
exposure to TCI renders this study inconclusive.
7.3  OTHER STUDIES INDICATIVE OF MUTAGENIC ACTIVITY
     Additional studies have been conducted on the genotoxicity of TCI.   These
studies do not measure mutagenic events per se, in that they do not demonstrate
the induction  of  heritable  genetic alterations, but positive results in these
test  systems  show that  DNA has been affected.   Such  test  systems provide
supporting evidence useful for qualitatively assessing genetic risk.

7.3.1.  Gene Conversion in Yeast
     An increase  in the frequency of mitotic gene conversion has been observed
after  exposure  to chemical  mutagens  as  a  consequence of  genetic  repair (Fabre
and Roman,  1977).   Bronzetti et al.   (1978)  tested  TCI  for its ability  to
induce  this end  point  at the  ade  and trp loci  in  strains D4 and D7 of
Saccharomyces cerevisiae.  Strain D7 cells were incubated with TCI  (ranging up
to 40  mM) with or without  CD-I  mouse  liver S9 mix for 4 hours at  37°C in a
liquid  suspension assay.   A dose-related increase in trp  convertants was
observed  in the tests conducted with metabolic activation  (41.0 ±  2.6 x  10
convertants at  40 mM  compared to 16.3 ± 0.42 x 10   convertants for the negative
control).   No  increase was  noted in the  test conducted without activation.
Toxicity  was  observed in both tests  with  a 40 to  50  percent reduction  in  cell
survival  at the highest dose (40 mM).  Twofold to fivefold increases  in  the
incidences  of  gene conversion were  noted  in these tests  and also in the  host-
mediated  assays where strains D4 and D7  were  instilled  into the retroorbital
sinus  of  adult male  CD-I mice.  TCI was  given  by gavage  in an acute-exposure
experiment  (single dose of  400  mg/kg)  or a subacute-exposure experiment (22
pretreatment  dosages  of 150 mg/kg 5 days/week followed  by  a  single 400 mg/kg
dose).  The largest increases  (fourfold over  spontaneous  values) were found in
yeast isolated from the  liver  and  kidney  (target  organs  for TCI  toxicity), and
gene  conversion  frequencies were  higher with  subacute  dosing,  compared  to
acute dosing.
                                     7-26

-------
     Call en et  al.  (1980) treated D7 cells  to  0,  15,  and 22  mM solutions  of
TCI  in liquid suspension tests  at  37°C for 1  hour  followed  by plating on
selective media.  Cell survival was reduced by 30 percent at 15 mM and by  >99
percent at  22 mM  TCI.  A fivefold increase  in trp  convertants  was  detected at
15 mM  compared  to the negative controls (76 x 10   versus 14 x 10   , respec-
tively). There  was  also  a fourfold increase in mitotic  recombination at the
ade  2  locus at  15 mM  TCI compared to  the negative  controls  (12 x 10   vs.  3  x
  -3
10   ,  respectively).   These end points were not reported for  the cultures
exposed to 22 mM TCI because of the extreme toxicity at this dose.   The positive
responses for  gene  conversion  and  mitotic  recombination in Saccharomyces
strain D7 indicate TCI is capable of interacting with DMA in yeast.

7.3.2.   Sister Chromatid Exchange (SCE) Formation
     White et al.  (1979) evaluated the ability of anesthetic-grade TCI (source
and  purity  not  reported) to cause sister chromatid exchanges  (SCE) in Chinese
hamster ovary cells.  The material was  administered as a  gas to exponentially
growing cells at  0.17 percent of the  atmosphere  in the presence of 19 percent
oxygen, 5 percent carbon dioxide, and 75.83 percent nitrogen for 1 hour with S9
mix prepared from Aroclor 1254-treated male rats.  After exposure,  the concen-
tration of TCI  gas  in the flask was measured by gas chromatography and found
to be  59 percent  of that delivered.   The cells  were allowed to go two rounds
of cell division  in the  presence of bromodeoxyuridine (BrdU).   Subsequently,
chromosome preparations were made and stained with Giemsa.  Onehundred control
and  100 treated metaphases  were scored for SCEs by  a single observer.   The
number of SCEs/chromosome were 0.536 ±  0.018 and 0.514 ±  0.018, respectively.
Under  the  conditions  of  this  test,  anesthetic-grade TCI did  not  yield an
increased incidence of SCEs.  However,  it should be noted that  the conditions
of the test may not have provided an adequate  assessment of TCI's  ability to
cause  sister  chromatid  exchanges.   In  the  first place,  only  one  dose  was
administered,  and it may have been too  low.  There was no indication of toxi-
city in the treated cells,  further  suggesting  that they  did  not receive  an
adequate dose.   In  the second place, no concurrent  positive  controls were
employed to ensure  that  the  test system was working properly.   These defici-
encies  reduce the weight of  the negative response reported by White et al.
(1979).
                                    7-27

-------
     Gu et  al.  (1981) collected blood  from  six individuals occupationally
exposed to TCI, cultured their peripheral lymphocytes, and scored the chromo-
somes for SCEs.  Blood was  also collected from nine  additional  persons who
served as the  control  group and as blood donors for testing trichloroethanol
and chloral hydrate,  metabolites of TCI, i_n  vitro for the induction of SCEs.
A significantly higher number of SCEs/cell was reported for the exposed group
compared to the controls  (9.1 ±0.4 versus 7.9 ± 0.2).  There was no information
about the levels of  TCI  to  which the subjects were exposed in the workplace,
but there was  a  good correlation between the average number of SCEs/cell  per
person and the levels of  trichloroacetic acid and trichloroethanol measured in
the blood. Those with the  highest mean  value  for  SCEs also had the highest
level of  these metabolites  of TCI  in their blood.   Cultured lymphocytes were
exposed to trichloroethanol  and chloral  hydrate at  178 and 54.1 mg/1,  respec-
tively, for 68 to  72 hours, and increases in SCEs were observed, compared to
the untreated cultures.  The results are tabulated as follows:

1.
2.
3.
4.
5.
Group
Control
Trichloroethanol (1)
Trichloroethanol (2)
Chloral hydrate
TCI in vivo

No. of
Karyo types
212
24
72
52
194
X ± SD
7.
9.
9.
10.
9.
9 ±
8 ±
2 ±
7 ±
1 ±
3.
3.
3.
4.
4.
0
2
3
0
9
Comparison
Between

1
1
1
1

&
&
&
&

2
3
4
5
P Value
from
t Test*

<0.05
<0.01
<0.001
<0.01
*Calculated by Gu et al.

     The  results obtained by Gu et al.  (1981) are  suggestive of a positive re-
sponse.  The  chemical  substances  to  which the TCI workers in this  study (Group
5)  were exposed and the extent to which  such exposures  occurred are  unknown.
It  is  likely that they were exposed to other substances which have been shown
to  be  mutagenic (e.g., 1,2-epoxybutane and epichlorohydrin), and these may be
responsible  for the weak effects.   However, the suggestive responses obtained
by  Gu  et al. (1981)  in in  vitro studies  with trichloroethanol  (>99.5 percent
                                     7-28

-------
pure, Fluka; Groups  2  and 3) and crystalline chloral  hydrate  (Group 4) show
that these metabolites  of TCI  could be responsible for the induction of SCEs
in TCI workers.

7.3.3.  Unscheduled DNA Synthesis (UPS)
     Four tests have been performed to assess the  ability of TCI to cause
unscheduled DNA  synthesis (Beliles et al.,  1980;  Perocco and Prodi, 1981;
Williams and Shimada,  1983;  Williams 1983).  Although demonstration  that a
chemical causes unscheduled  DNA synthesis does not provide a  measurement of
its  mutagenicity,  it does  indicate that the material  interacts  with DNA.
     Beliles et al.  (1980)  exposed human WI-38 cells to TCI at doses ranging
from 0.1 to 5.0  pi/ml  to evaluate  its ability  to cause UDS.   The positive
controls N-methyl-N-nitro-N-nitrosoguanidine  (MNNG) and benz(a)pyrene  (BaP)
were  used  at  concentration ranges  between  1.25  to 10 pg/ml and  2.5  to 20
jjg/ml,  respectively.  Both agents were positive;  however,  BaP  produced  a  non-
linear dose-response effect.   All studies were conducted at the same time on a
single  batch of WI-38  cells.   TCI yielded  a weak  increase at  the two lowest
doses (1.5- to 1.8-fold increase),  both with and  without activation (Table
7-8).  The response is only suggestive of an effect, but the testing done with
S9 may  not  have  provided an optimal test of its genotoxic potential, because
the  positive control (BaP) was  only effective at  one dose  (2.5 (jg/ml);  higher
doses resulted in a negative response, although this may have been due to cell
killing.
     Perocco and  Prodi  (1981)  collected blood samples from  healthy humans,
separated the lymphocytes, and cultured 5 x 10  of them in 0.2-ml  medium for 4
hours at 37°C in  the presence or absence  of TCI  (Carlo Erban,  Milan,  Italy  or
MerckSchuchardt,  Darmstadt, FRG, 97-99 percent pure).   The tests were conducted
both  in the presence and in the  absence  of PCB-induced rat liver 59 mix.   A
comparison was made between treated  and  untreated cells   for  scheduled DNA
                                             *
synthesis (i.e.,  DNA replication) and unscheduled DNA  synthesis.   A difference
was  noted between the groups, with respect to scheduled DNA synthesis measured
as dpm  of  [ H] deoxythymidylic  acid (TdR) after  4  hours of culture  (2661  ±  57
dpm  in  untreated  cells, compared to  1261 ± 36  dpm in cells treated with 5
pi/ml TCI).  Subsequently, 2.5,  5,  and 10  pi/ml  (0.028,  0.056,  and 0.11
jjmole/ml) TCI was added to cells which were cultured  in 10 mM hydroxyurea to
suppress scheduled DNA synthesis.  The amount of unscheduled DNA synthesis was
                                    7-29

-------
                                     TABLE  7-8.   UNSCHEDULED DMA SYNTHESIS IN WI-38 CELLS
 I
co
o
Test
Nonactivation
Solvent control, DMSO
MNNG
TCI
Activation
Solvent control, DMSO
BaPb
TCI
Compound
Concentration
(Ml/ml)
0
5 ug/ml
0.1
0.5
1.0
5.0
0
2.5 ug/ml
0.1
0.5
1.0
5.0
DNA
23.76
6.27
24.75
23.76
26.07
13.20
24.75
29.70
28.38
21.78
21.78
DPM
988.7
472.9
1561.3
1616.1
1221.3
150.6
834.5
1500.3
1622.1
860.6
1006.8
DPM/
(ug DNA)
41.6
75.4
63.1
68.0
46.8
11.4
33.7
50.5
57.2
39.5
46.2
Percent
of
Control Response
100
181 +
152 +
164 +
113
27
100
170 +
100
170 +
117
137
     Source:  Adapted  from  Beliles et al., 1980.

-------
                                               3
estimated by measuring  dpm  from incorporated [ H]TdR 4  hours  later.   At 10
ul/ml TCI, 392 ± 22 and 439 ± 23 dpm were counted, without and with exogenous
metabolic activation,  respectively.   Both values were lower than corresponding
negative controls of 715  ±  24 and 612 ± 26 dpm, respectively.   No positive
controls were run to ensure that the system was  working properly, although
testing of  chloromethyl  methyl  ether (CMME) with activation resulted in a
doubling of dpms over the corresponding  negative  control values (1320 ± 57 at
5 ul/ml CMME versus 612 ± untreated). The authors calculated an effective DNA
repair  value (r)  for  each chemical, based  on  the control  and experimental
values with and without metabolic activation.   TCI was evaluated by the authors
as positive in  the test, but they did not state their criteria for classifying
a chemical as positive.   It  is important to note that none of the experimental
values from cells treated with TCI without metabolic activation had higher dpm
values than the  controls.   Furthermore,  only one out of three  experimental
values was greater than the controls with metabolic activation (668 ± 34 at 5
Ml/ml, compared to control value of 612  ± 26). The increase was not statisti-
cally significant.   Thus,  the positive finding reported in this work is judged
to be inconclusive.
     In unpublished papers, Williams  and Shimada (1983) and Williams (1983)
tested samples of TCI for their ability  to  cause  unscheduled DNA synthesis in
rat and mouse primary  hepatocyte cultures (HPC DNA repair assay).   Williams and
Shimada (1983) report two samples  were tested using rat primary hepatocytes.
One was fully stabilized,  and the other,  referred to as "non-stabilized,"  con-
tained 99.9 percent TCI.   No information was provided about the remaining com-
ponents.  Two types of  tests  were conducted.   One was a conventional  HPC DNA
repair assay in that the hepatocytes were exposed to TCI added to  the culture
medium.  The other was  a  modified assay, and the cells  were placed in small
glass dishes in a sealed incubator and exposed to TCI vapor.  Tritiated thymidine
 3
( H-TdR) was added to  the culture medium in both types of tests.   If TCI  caused
DNA damage repaired by  the excision repair pathway,  H-TdR would be incorporated
into the  repaired DNA.   To detect such incorporation, the cells were fixed on
glass slides and subsequently coated with radiotrack emulsion.   After an  incuba-
tion period, the slides were  developed and examined microscopically to count
developed silver  grains  in the  radiotrack emulsion over the  cell nuclei.
                                    7-31

-------
     Williams and Shimada (1983) used 2-acetylaminofluorene and vinyl  chloride
as positive controls for the conventional  and modified exposure assays,  respec-
tively.   2-acetylaminofluorene was highly  active (170 grains/nucleus at  10  M)
and by comparison, vinyl chloride was weakly active at 2.5 percent  (5 grains/
nucleus)  and 5 percent  (11 grains/nucleus) in increasing unscheduled DNA syn-
thesis compared to the negative controls (3 grains/nucleus).
     In the modified HPC DNA repair assay, cells were exposed to dosages up to
2.5 percent TCI  in the  air for 3 or 18 hours.  No increase in the grain count
was observed (the number of grains/nucleus ranged from 0.2 ± 0.2 to 3.2 ± 2.5
in treated cells, compared to 2.75 ± 2.76  in the untreated controls).   Similar-
ly, negative results  were  reported in the conventional assay,  in  which the
cells were exposed  to concentrations  as high as 1 percent TCI in the medium.
     There are several  factors to consider  in  the  analysis  of this assay.
First, the system  is  designed to detect the occurrence of a specific type of
repair (i.e., "long-patch"  excision  repair).   Xenobiotics which cause damage
that is not  repaired  or that is repaired  by another mechanism will not be
detected as possessing genotoxic activity  in the system.
     Second, the end  point measured may not  actually  reflect repair synthesis
of chromosomal DNA  in the nucleus (Lonati-Galligani et al.,  1983).   Enhanced
                 3                                           3
incorporation of  H-TdR into  mitochondrial DNA or suppressed  HTdR  incorpora-
tion  into  mitochondrial DNA, without  a concomitant increased or decreased
incorporation into nuclear DNA, can lead to false negative or positive results,
respectively.  Thus,  both nuclear  incorporation  and cytoplasmic  incorporation
should be measured and the dose-responses  plotted for each to serve as a basis
for deciding whether or not the compound has induced UDS.
     Third, nuclear grain  counts vary with nuclear size (Lonati-Galligani et
al., 1983); thus, it may be appropriate to express grain counts as a percentage
of the measured  nuclear area.  In the studies by Williams and Shimada (1983)
the values were transformed by subtracting cytoplasmic grain counts from total
nuclear grains counts to give a net  grain count;  as a result, the data are
difficult to evaluate,  and the way they are presented may preclude the detection
of weakly active agents.  Thus, although the results of this study are negative,
they do not provide convincing evidence that TCI does not induce UDS.  This is
a  relevant  point,  because  vinyl  chloride,  a  structurally  related compound  and
known mutagen, was  only weakly active in  this test and then only at doses as
high or higher than the  highest dose tested  for TCI.
                                    7-32

-------
     In the  Williams  (1983) study, TCI  (source  and purity not reported)  was
assayed in the HPC DMA repair test using a conventional liquid exposure protocol
Primary hepatocytes from B6C3F1 mice and Osborne-Mendel  rats were tested in
two  separate  test  series.   A positive dose-related response was  obtained in
the  mouse  cells  between  10~4M and  10~2M  (18.24 ± 3.95 grains/nucleus  at 10"2M
vs.  0.78 ± 1.18  grains/nucleus  in the concurrent negative control).  TCI was
                                   _c         _p
reported to  be nontoxic between 10  M and  10 M.   A negative response was
obtained in  a test conducted in rat cells.   Positive and negative controls
were employed for  both sets of experiments  and  responded appropriately,  but
the  aforementioned concerns regarding  the Williams  and Shimada (1983) study
apply here as well.   In spite of  this,  the positive response in  the mouse
cells shows that a commercially available sample of TCI is genotoxic  in B6C3F1
mouse liver cells.
     The positive  responses for gene conversion and mitotic recombination in
yeast, and UDS in mouse cells, provide evidence that commercially available TCI
is weakly active in damaging DMA.
7.4  EVIDENCE THAT TCI REACHES THE GONADS
     Although not bearing on the ability of TCI to cause mutations, two studies
have been conducted to assess its ability to induce sperm morphological abnormal-
ities in mice  (Land  et al., 1979; Beliles  et  al. ,  1980).  A  statistically
significant increase was  observed in both studies, strongly suggesting that
the compound or  an  active metabolite reaches the gonads.  Land et al. (1979)
exposed groups of  (C57BL/C3H)F1 male mice  (13  weeks  old) to air  (negative
control),  0.02 percent and 0.2 percent anesthetic-grade TCI vapor 4 hr/day for
5  consecutive  days.  Twenty-eight days after the  first day of exposure the
animals were sacrificed.  Both cauda epididymides were removed,  minced, strained,
and stained in 1 percent Eosin Y.  Slides were prepared, mounted, coded,  and
evaluated by  scoring 1000  sperm/slide  at 400  X  magnification.   Data were
recorded as percent abnormal sperm.  The means  were calculated for each group
and compared  with  controls  using the sample t-test.   As  noted  below, TCI
exposure resulted in statistically significant elevations in the percentage of
morphologically abnormal  sperm.
                                    7-33

-------

Agent
Air
TCI


Concentration %

0.02
0.2
Number of
Animals Studied
15
5
5
% Abnormal Sperm
± S.E.
1.42 ± 0.08
1.68 ± 0.17
2.58 ± 0.29*
*Statistically significant difference compared to controls (P <0.001).

This study provides suggestive evidence that exposure to a high concentration
of  TCI  during the period of meiotic  division damages early spermatocytes.
     Beliles et al. (1980) also observed a significant increase in sperm head
abnormalities in male  CD-I  mice,  but not albino rats [CRL: COBS CD(SD) BR].
Groups of 12  animals  were dosed 7 hr/day for 5 days to 0, 100, or 500 ppm
(0,538, or 2690 mg/m  ) TCI.  After dosing, groups of 4 animals were killed at
the end  of  1 week, 4 weeks, and  10  weeks.   Sperm were  collected  from the
caudae epididymides and  suspended in  0.85 percent saline, stained, and scored
for abnormalities.   At least 2000 sperm were analyzed/group (500/animal).   The
negative responses obtained  in  the  rat study are not considered adequate for
assessing the ability of TCI to reach the testes, because the positive controls
gave a negative response.  Increased incidences of abnormal sperm were observed
in  the mice of the 500 ppm (2690 mg/m ) group compared to the negative controls
for weeks 1  and 4 (i.e., 18.9 percent  versus 6.8 percent and 23.5 percent
versus 8.1  percent,  respectively).   A significant increase was also noted in
the 100-ppm  exposure  group,  compared  to  the  negative  control  for week  4  (14.8
percent versus 8.1 percent).

7.5 MUTAGENICITY OF METABOLITES
     Trichloroethylene  is metabolized to trichloroethanol,  trichloroethanol
glucuronide,  and trichloroacetic acid.  There are several intermediates leading
to  the formation of these end products  of  metabolism,  including some which
have  been  identified in man (e.g.,  chloral  hydrate)  and some which  have not
been  identified but which are  thought to occur (e.g., TCI oxide).  Some  muta-
genicity  studies  bearing on a few of TCI's metabolites  are discussed below.
See the  chapter on metabolism for a more detailed discussion.
                                     7-34

-------
     Kline et  al.   (1982)  tested  the mutagenic potential of  TCI  oxide in
Salmonella strain TA1535  and  E.  coli strain WP2 uvrA.  They observed no in-
creases in revertant colonies at doses ranging from 0.25 to 5 mM.   Preincubation
tests were conducted,  and toxicity was observed at the highest doses, demon-
strating that the bacteria were exposed to sufficient concentrations of the
test material.  In the  same paper,  epoxide metabolites of other chloroalkenes
were found to be  mutagenic under the same conditions of test (i.e., cis  and
trans-1-chloropropene oxide, and cis and trans 1,3-dichloropropene oxide).  The
genotoxicity of  these  compounds was also assessed  by  differential growth
inhibition of the DNA  polymerase-deficient E.  coli  mutant polA  compared  to
its polymerase  proficient polA  parent.  TCI oxide was positive in  this assay
yielding a dose-related decrease  in the survival of polA  cells  (up to 40
percent) compared to the  polA  cells, at  doses  ranging  from 0.006 to  0.11
mM/ml.
     Both Waskell (1978)  and  Gu et  al.  (1981) assessed trichloroethanol for
its ability to  cause genetic  effects.  As discussed in  the section on gene
mutation studies in Salmonella  (using TA98 and TA100),  Waskell (1978) tested
trichloroethanol at doses up  to 7.5 mg/plate and  found  no increases in the
revertants, compared to the negative controls.
     Gu et al.  (1981), on the other  hand,  found  suggestive evidence that tri-
chloroethanol  (at 178  mg/1) may be an inducer of SCEs in primary cultures  of
human  lymphocytes.   In two experiments  (of 24  and 72 karyotypes analyzed,
respectively), Gu et al.  (1981) found 9.8 ± 3.2 and 9.2  ± 3.3 SCEs/nucleus,
compared to the control value of 7.9 ± 3.0.
     Trichloroacetic acid and chloral hydrate were  also tested by Waskell
(1978) in  the  study  mentioned above.  Trichloroacetic acid at 0.45 mg/plate
was not mutagenic to either TA98 and TA100.  Chloral  hydrate, however, was
weakly active, yielding consistent, but always less than twofold,  increases in
the number of TA100 revertants,  compared to the spontaneous level, over a dose
range from 0.5  to  10  mg,  both with  and without metabolic activation.   These
increases  suggest chloral hydrate  is weakly mutagenic to Salmonella strain
TA100.
     In their testing  of  chloral hydrate,  Gu et  al.  (1981) found that chloral
hydrate at 54.1 mg/1 caused increases in the number  of SCEs in cultured human
lymphocytes, compared  to  untreated control cells (i.e., 10.7 ± 4.0  versus  7.9
± 3.0).  Again,  the response was merely suggestive of an effect.
                                    7-35

-------
7.6  BINDING TO DNA
     Stott et  al.  (1982)  studied  the pharmacokinetics  and macromolecular
Interactions of TCI in B6C3F1 male  mice and male Osborne-Mendel  rats.   As part
of their study, mice were given 1200 mg/kg [  C] TCI (>500 uCi/animal,  specific
activity = 2.4  mCi/mmol)  in  corn oil.  Five  hours  later,  the animals were
killed, their  livers  were  removed  and frozen, and DNA was extracted two days
later.   After  enzymatic  digestion,  fractions  corresponding to nucleosides or
purine and pyrimidine bases  were separated and  identified  by high-pressure
liquid chromatography.  Aliquots were placed  in scintillation fluid and the
14
  C content was determined by  liquid  scintillation counting.  Analysis of the
data indicated  that  the  ability of TCI to  alkylate  DNA was low.   A maximum
estimate of the average  DNA  alkylation level  was 0.62 ± 0.42 alkylations/10
nucleotides.  These data  suggest TCI or its  metabolites  are  capable of binding
to DNA, but only  to a limited extent, and are consistent with the suggestive
mutagenic responses reported for TCI and certain of its metabolites.
7.7  SUMMARY AND CONCLUSIONS
     Commercially available samples of TCI have been shown to cause responses
suggestive  of  a positive  effect  in gene mutation  studies  using bacteria,
fungi,  higher  plants,  and  somatic  cells of mice iji vivo.  These responses
occurred with  metabolic activation only, suggesting the involvement of one or
more  metabolites  of TCI.   This requirement for metabolic  activation  argues
against  the possibility that  epoxide  stabilizers  found in many commercial
samples  are responsible.   The marginally increased  incidences  of  revertant
counts  were  only  observed  at  high doses.  Thus, commercially available TCI is
only weakly mutagenic at most. TCI was not shown to cause structural chromosomal
aberrations in the one test conducted to assess this end point.
     Other tests provide evidence that commercially available TCI damages DNA.
Suggestive  and weak-positive  responses  have  been  observed  in  yeast (gene
conversion  and mitotic  recombination),  mice (UDS), and humans (SCE and UDS).
Metabolic activation was  required to obtain the positive responses.  Certain
metabolites of TCI have been tested for their mutagenic potential,  and sugges-
tive  positive  effects have been shown.   TCI or a metabolite(s)  may  be  capable
of binding  to  DNA, but only minimally.
                                    7-36

-------
     TCI causes  weak increases in morphologically  abnormal  sperm providing
evidence that its  reaches  the gonads.   A synopsis  of  the results of these
studies is presented in Table 7-9.
     The available data provide suggestive evidence that  commercial-grade TCI
is a weakly active indirect mutagen, causing effects in a number  of different
test systems representing  a  wide  evolutionary range of organisms, including
humans.  Based on this, commercial  TCI  may have the potential to cause weak or
borderline increases  above the spontaneous  level of  mutagenic  effects  in
exposed human tissue.  The observation that TCI causes adverse effects in the
testes of mice suggests TCI may cause adverse testicular  effects  in man, too,
provided that the pharmacokinetics  of TCI in humans also results in its distri-
bution to the gonads.   The data on pure TCI do not allow a  conclusion to be
drawn about its  mutagenic  potential.   However,  mutagenic potential cannot be
ruled out.   If it is mutagenic, the available data suggest TCI would be a very
weak, indirect mutagen.
                                    7-37

-------
                                       TABLE 7-9.  SUMMARY OF TESTS FOR MUTAGENICITY OF TCI
CO
CO
Test
Category
I . Gene
Mutations






















Organism Type of Test
Salmonella Reverse mutations
typhimurium in vitro


Plate.
incorporation
tests






Vapor
exposure


Escherichia Forward and
coli reverse mutations


Schi zosaccharomyces Forward mutations
pombe (Host-mediated
assays)
Purity
of TCI Results
Technical
grade




Technical +
grade
Purified
Anesthetic
grade
Purified *

Purified3 *

Reagent *
grade
Analytical *
grade


Technical *
grade

Comments
No control to test
effectiveness of
S9 mix. No
precautions to
prevent evaporation.

Twofold increase




1.8-fold increase

1.3-fold increase

1.7-fold increase

'Positive for reverse
mutations only at
arg locus (twofold
increase).
1.7-fold increase


Reference
Henschler
et al . , 1977




Margard, 1978

Margard, 1978
Waskell, 1978

Bartsch
et al . , 1979
Baden
et al . , 1979
Simmon et
al., 1977
Greim
et al . , 1975


Loprieno
et al . , 1979

    + = Positive
    - = Negative
    * = Suggestive
    a = No detectable epoxides
    x = Inconclusive

-------
                                                       TABLE 7-9.   (continued)
i
CO
Test Purity
Category Organism Type of Test of TCI Results
Purified
Purified
Purified3

Saccharomyces Reverse mutations Technical x
cerevisiae (in vitro) arade

ACS reagent +
grade
Technical +
grade
Tradescantia Forward mutations Unknown *

Drosophila Sex- linked Technical
melanogaster recessive lethals grade
Technical
grade
Mouse Spot test Technical +
grade
Comments

Epichlorohydrin
and epoxybutane
were also negative.


High toxicity
Fourfold increase
both host-mediated
assay and liquid
suspension test.
Twofold increase





Sixfold increase
Reference
Loprieno
et al . , 1979
Rossi et al. ,
1983
Mondino,
1979
Shah in and
Von Borstel ,
1977
Bronzetti
et al . , 1978
Callen
et al . , 1980
Schairer
et al . , 1978
Abrahamson
and Valencia,
1980
Beliles
et al . , 1980
Fahrig
et al . , 1977
    + =  Positive
    - =  Negative
    * =  Suggestive
    a =  No detectable  epoxides
    x =  Inconclusive

-------
                                                       TABLE 7-9.  (continued)
-J

o
Test
Category Organism
II. Chromosomal Drosophila
Aberrations melanogaster
Rat

Mouse




Human


III. Other Saccharomyces
Studies cerevisiae
Indicative
of Mutagenic
Damage
Purity
Type of Test of TCI • Results
Chromosome Technical
loss grade
Bone marrow Technical x
grade
Dominant Purified
lethal

Micronucleus Analytical x
grade

Breaks Occupational
exposure
Hypodiploid cells Occupational x
exposure

Gene conversion ACS reagent +
grade

Comments

Positive control
given by different
route of exposure.
Doses of TCI may
have been too low.



Positive response
reported by authors
may be due to arti-
facts in mature
erythrocytes.

Unmatched control
group. Hypodiploid
cells can be caused
by preparation of
chromosomes.
Twofold increase
with metabolic
activation.

Reference
Bellies
et al . , 1980
Bellies
et al . , 1980
Slacik-
Erben
et al., 1980
Duprat and
Gradiski,
1980

Konietzko
et al . , 1978
Konietzko
et al . , 1978

Bronzetti
et al . , 1978

     + = Positive
     - = Negative
     * = Suggestive
     a = No detectable epoxides
     x = Inconclusive

-------
                                                  TABLE 7-9.   (continued)
Test
Category Organism Type of Test


Mitotic
recombination
Mouse Sister chromatic!
exchange


Human Sister chromatid
exchange





Unscheduled DNA
synthesis


Rat HPC DNA
repair
assay





Purity
of TCI
Technical3
grade
Technical3
grade
Anesthetic
grade


Occupational
exposure





Technical
grade
Technical
grade
Stabilized


Unstabilized


Technical
grade
Results Comments
+ Fivefold increase

+ Fourfold increase

- No positive
controls. No
evidence of
toxicity.
* Increases
correlated with
presence of TCI
metabolites tri-
chloroethanol and
trichloroacetic
acid in the blood
* 1.5 to 1.8-fold
increases
X

Vinyl chloride
only weakly
active.
-


-

Reference
Callen
et al . , 1980
Callen
et al . , 1980
White
et al . , 1979


Gu et al. ,
1981





Bellies
et al . , 1980
Perocco and
Prodi, 1981
Williams and
Shimada,
1983
Wi 1 1 i ams and
Shimada,
1983
Williams,
1983
+ = Positive
- = Negative
* = Suggestive
a = No detectable epoxides
x = Inconclusive

-------
                                                         TABLE 7-9.  (continued)
i
-E»
ro
Test
Category



IV. Evidence
TCI Reaches
the Gonads


V. Mutagenicity
A. TCI-oxide






Organism
Mouse


Mouse




of Metabolites
Salmonella
typhimurium
Escherichia
coli



Purity
Type of Test of TCI
HPC DNA Technical
repai r grade
assay
Morphological Anesthetic
sperm grade
abnormalities
Technical
grade

Reverse mutations

Reverse mutations

Differential
killing of repair
deficient bacteria
Results Comments Reference
+ 8- to 20-fold Williams,
increases 1983

* 1.8- fold increase Land
et al . , 1979

+ Threefold increase Beliles
et al . , 1980

Kline
et al . , 1982
Kline
et al . , 1982
+ 40% decreases in Kline
survival of Pol- etval., 1982
vs. Pol + cells
B. Tri chl oroethanol




Salmonella
typhimurium
Human
lymphocytes
Reverse mutations

Sister chromatid
exchange
Waskell,
1978
* Gu et al . ,
1981
       + = Positive
       - = Negative
       * = Suggestive
       a = No  detectable epoxides
       x = Inconclusive

-------
                                                  TABLE 7-9.   (continued)
Test
Category Organi sm
C. Trichloroacetic Acid
Salmonella
typhimurium
-J D. Chloral Hydrate
00 Salmonella
typhimurium
Human
lymphocytes
Purity
Type of Test of TCI Results
Reverse mutations
Reverse mutations *
Sister chromatid *
exchange
Comments Reference
Waskell,
1978
1.6-fold increase Waskell ,
1978
Gu et al . ,
1981
+ = Positive
- = Negative
* = Suggestive
a = No detectable epoxides
x = Inconclusive

-------
7.8  REFERENCES
Abrahamson, S. and  R.  Valencia.  1980. Evaluation  of  substances of interest
     for  genetic  damage  using Drosophila  melanogaster.  Final  sex-linked
     recessive lethal  test  report  to FDA on 13  compounds.  FDA Contract No.
     233-77-2119.

Baden, J.M., M. Kelley,  R.I.  Mazze,  and V.F.  Simmon.  1979.  Mutagenicity of
     inhalation anesthetics: trichloroethylene,  divinyl  ether, nitrous oxide
     and cyclopropane.  Br. J.  Anaesth. 51:  417-421.

Bartsch, H. , C. Malaveille,  A.  Barbin, and G.  Planche.  1979.  Mutagenic and
     alkylating metabolites of halo-ethylenes, chlorobutadienes and dichloro-
     butenes produced  by  rodent  or human liver  tissues.  Arch.  Toxicol. 44:
     249-277.

Beliles,  R.P.,  D.J.  Brusick,  and  F.J.  Mecler.  1980.  Teratogenic-mutagenic
     risk of workplace contaminants: trichloroethylene, perchloroethylene,  and
     carbon disulfide,  contract report, contract no. 210-77-0047. U.S.  Department
     of HHS, NIOSH, Cincinnati, OH 45226.

Bronzetti, G. , E. Zeiger, and D. Frezza. 1978. Genetic activity of  trichloro-
     ethylene in yeast. J. Environ. Pathol.  Toxicol. 1: 411-418.

Callen, D.F., C.R.  Wolf,  and R.M.   Philpot. 1980.  Cytochrome P-450 mediated
     genetic activity and cytotoxicity of seven  halogenated aliphatic hydrocar-
     bons in Saccharomyces cerevisiae. Mutat.  Res. 77: 55-63.

Duprat, P., and D.  Gradiski. 1980.  Cytogenetic effect of  trichloroethylene  in
     the mouse as  evaluated by the micronucleus test.  IRCS  Medical Science
     Libr. Compendium 8: 182.

Fabre, F., and H.  Roman. 1977.  Genetic evidence  for inducibility of recombina-
     tion competence in yeast.  Proc.  Natl.  Acad. Sci.  (U.S.A.) 74(4): 1667-1671.

Fahrig, R.  1977.  The mammalian spot  test  (Fellfleckentest)  with mice.  Arch.
     Toxicol.  38:  87-98.

Greim, H., G.  Bonse, Z. Radwan, D.   Reichert, and D. Henschler.  1975. Mutageni-
     city iji vitro and potential carcinogenicity of chlorinated ethylenes as a
     function of metabolic oxirane  formation.  Biochem. Pharmacol. 24: 2012-2017.

Gu, Z.  W. ,  B.  Sele,  P.  Jalbert,  M.  Vincent, C.  Marka, D.  Charma,  and J. Faure
     1981.  Induction d'echanges entre  les  chromatides soeurs  (SCE)  par  le
     trichloroethylene et ses metabolites.  Toxicol. Eur.  Res.  3: 63-67.

Henschler, D. ,  E.  Eder, T.  Neudecker,  and  M.  Metzler. 1977.  Carcinogenicity
     of  trichloroethylene:  fact or  artifact? Arch.  Toxicol.   37:  233-236.

Kline, S.A., E.C. McCoy, H.S. Rosenkranz, and  B.L. Van Duur&n.  1982. Mutageni-
     city of chloroalkene epoxides  in bacterial  systems.  Mutat.  Res. 101: 115-
     125.
                                    7-44

-------
Konietzko, H. W. , W.  Haberlandt, H. Heilbronner, G. Reill, and H. Weichardt.
     1978. Cytogenetlsche Untersuchungen  an Trichlorathylen-Arbeitern. Arch.
     Toxicol. 40: 201-206.

Land, P.  E.,  E.  L.  Owen, and H.W. Linde. 1979. Mouse sperm morphology  following
     exposure to  anesthetics during  early spermatogenesis.  Anesthesiology
     51(3):  S259.

Lonati-Galligani, M. ,  P.  H.  M.  Lohman, and F. Berends. 1983. The  validity  of
     the  autoradiographic method for detecting  DMA repair synthesis  in  rat
     hepatocytes in primary culture. Mutat. Res. 113: 145-160.

Loprieno, N., R. Barale, A.M. Rossi, S. Fumero,  G.  Meriggi, A. Mondino, and S.
     Silvestri.  1979.  Iji  vivo mutagenicity studies  with trichloroethylene and
     other solvents (preliminary results). Unpublished.

Margard, W.  1978. Summary report on i_n  vivo bioassay of chlorinated hydrocar-
     bon solvents.  Unpublished.

Matter, B. E. and  I.  Jaeger. 1975.  The cytogenetic basis of dominant lethal
     mutations  in mice.  Studies wth TEM,  EMS,  and 6-mercaptopurine.  Mutat.
     Res. 33: 251-260.

Mondino, A.  1979.  Examination  of the live mutant activity of the trichloro-
     ethylene compound with "Schizosaccharomyces pombe."  Unpublished.

National Cancer  Insititute  (NCI).  1976.  Carcinogenesis bioassay of trichloro-
     ethylene. CAS no. 79-01-6.  NCI-CG-TR-2.

Perocco, P.  and G.  Prodi. 1981.  DNA damage by haloalkenes in  human lymphocytes
     cultured i_n vitro. Cancer Lett. 13: 213-218.

 Rossi, A. M. ,  L.  Migliore, R.  Barale, and N. Loprieno.  1983.  In  vivo and  i_n
     vitro mutagenicity  studies  of a possible carcinogen, trichloroethylene,
     and  its stabilizers,  epichlorohydrin  and 1,2-epoxybutane.  Teratog.
     Careinog. Mutagen. 3: 75-87.

Russell, L.B., and B.E. Matter.  1980. Whole-mammal  mutagenicity tests.  Evalua-
     tion of five methods. Mutat. Res.  75: 279-302.

Schairer,  L.A.,  J.  Van't Hof, C.G.  Hayes, R.M.  Burton,  and  F.J.  deSerres.
     1978. Measurement of biological activity of  ambient air mixtures using
     a mobile laboratory for jji situ exposures. Preliminary  results from the
     Tradescantia plant test sytem. EPA Pub. 600/9-78-027. Pp.  421-440.

Schmid, W. 1975. The micronucleus test. Mutat. Res.  31: 9-15.

Shahin,  M.M.,  and  R.C. Von  Borstel.  1977. Mutagenic and lethal  effects  of
     -benzene  hexachloride,  dibutyl phthalate  and  trichloroethylene in
     Saccharomyces cerevisiae.  Mutat. Res. 48: 173-180.

Simmon,  V.F.,  K. Kauhanon,  and R.G. Tardiff.  1977.  Mutagenic activity  of
     chemicals  identified in drinking water.  Pages 249-258 in D.  Scott, B.  A.
     Bridges,  and  F.  H. Solves,   eds.  Progress  in genetic toxicology.
     Elsevier/North Holland Biomedical  Press.

                                    7-45

-------
Slacik-Erben, R., R. Roll, G. Franke,  and  H. Uehleke.  1980.  Trichloroethylene
     vapours do  not produce dominant  lethal mutations in male mice.  Arch.
     Toxicol. 45:  37-44.

Stott, W.T., J.F. Quast,  and P.G. Watanabe. 1982.  The pharmacokinetics and
     macromolecular interactions of trichloroethylene  in mice and rats. Toxicol
     Appl. Pharmacol.  62:  137-151.

Waskell,  L.  1978.  A  study of  the  mutagenicity of  anesthetics and their
     metabolites.  Mutat. Res. 57: 141-153.

White, A.S., S. Takehisa,  E.I.  Eger II, S.  Wolff,  and W.C. Stevens.  1979.
     Sister-chromatid exchanges induced by inhaled anesthetics. Anesthesiology
     50: 426-430.

Williams,  G.M.  1983.  Draft  final report TR-507-18. DNA  repair test of 11
     chlorinated hydrocarbon analogs.  For  ICAIR Life  Systems,  Inc., and U.S.
     Environmental  Protection Agency. Dr. Harry Milman, Project Officer, 401  M
     St., SW, Washington,  DC 20460.  Unpublished.

Williams,  G.M.,  and T. Shimada.  1983.  Evaluation  of  several  halogenated
     ethane  and  ethylene  compounds  for  genotoxicity.  Final report  for PPG
     Industries, Inc.,  Dr.  A.P.  Leber,  PPG Industries,  Inc., One  Gateway
     Center, Pittsburgh, PA 15222. Unpublished.
                                     7-46

-------
                              8.  CARCINOGENICITY

8.1  DESCRIPTION AND ANALYSIS OF ANIMAL STUDIES AND CELL TRANSFORMATION STUDIES
     There have been a number of laboratory investigations of the carcinogenic
potential of trichloroethylene (TCI)  in experimental  animals.  These studies
have been done using rats, mice, and  hamsters, with TCI administered by inhala-
tion, gavage, subcutaneous injection, and topical  application.  The results  are
summarized in Table 8-1.  Of the studies done, the evidence for the carcinoge-
nicity of TCI consists of statistically significant increases of hepatocellular
carcinomas in male and female B6C3F1  mice [National Toxicology Program (NTP),
1982; National Cancer Institute (NCI), 1976;  Bell  et  al., 1978], malignant
lymphomas in female NMRI mice (Henschler et al., 1980), and renal  adenocarci-
nomas, significant by life table and  incidental tumor tests,  in male Fischer
344 rats (NTP, 1982).  However, inadequacies  in the NTP study in rats and the
IBT study in mice and limitations in  the interpretation of the data in the
Henschler et al. study in mice are evident, as discussed herein.

8.1.1  Oral  Administration (Gavage):  Rat
8.1.1.1  National Toxicology Program  (NTP, 1982)—A carcinogenicity bioassay on
TCI in Fischer 344 rats was recently  completed for the National  Toxicology
Program, and a 1982 draft report of the study has  been reviewed by its Board of
Scientific Counselors.
     The TCI test sample was described as a high-purity "Hi-Tri" product  obtained
in two lots, one being Dow Chemical Company Lot No. TB05-206AA,  used for  a pre-
liminary dose-finding study and for the first 7 months of the bioassay, and the
other being  Missouri  Solvents Lot No. TB08-039AA,  used for the duration of the
bioassay.  Samples of TCI were chemically analyzed by gas-liquid chromatography,
                                      8-1

-------
TABLE 8-1.   TCI CARCINOGENICITY BIOASSAYS IN ANIMALS
TCI chemical
Study purity
NTP, 1982 Purified






NCI, 1976 Technical grade






Bell et al., Technical grade
(MCA) 1978






Maltoni, 1979 Purified



Henschler et al., Purified
1980







Species
Mice, B6C3F1
Males
Females
Rats, Fischer
344
Males
Females
Mice, B6C3F1
Males
Females
Rats, Osborne-
Mendel
Males
Females
Mice, B6C3F1
Males
Females

Rats, Charles
River
Males
Females
Rats, Sprague-
Dawley
Males
Females
Mice, Han:NMRI
Males
Females
Rats, Han:Wist
Males
Females
Hamsters, Syrian
Males
Females
Dose levels,
route
1000 mg/kg/day
gavage, 103 wk

500, 1000 mg/kg/day
gavage, 103 wk


1119, 2339 mg/kg/day
869, 1739 mg/kg/day
gavage, 78 wk
549, 1097 mg/kg
gavage, 88 wk


100, 300, 600 ppm
inhalation, 24 mo


100, 300, 600 ppm
inhalation, 24 mo


250, 50 mg/kg
gavage, 52 wk


100, 500 ppm
inhalation, 78 wk

100, 500 ppm
inhalation, 78 wk

100, 500 ppm
inhalation, 78 wk

Results
Treatment -related
hepatocellular carcinomas
in males and females
Renal adenocarcinomas
in treated males


Treatment-related
hepatocellular carcinomas
in males and females
Negative



Increased incidence of
hepatocellular carcinomas
in males and females
with dose
Negative



Negative



Increased incidence of
malignant lymphomas
in females
Negative


Negative


                                                [continued on the following page)
                       8-2

-------
                                              TABLE 8-1.  (continued)
Study
TCI chemical
   purity
Species
                                                                  Dose  levels,
                                                                     route
                                                                                           Results
Van Duuren et al.,
 1979, 1983
NTP, 1982
Maltoni, 1979,
 1985
Henschler et al.,
 1984
Purified
                        Purified TCI
                         epoxide
                        Purified
                       Purified
Purified and
 stabilized
Swiss mice
ICR/Ha
Female

Female
                                                 Female

                                                 Female
                                                 Male

                                                 Female
                                                 Female
                                                 Female
Rats, Osborne-
 Mendel
Marshall  540,
August 28807,
AC I

Mice, B6C3F1
Swiss albino

Rats, Sprague-
 Dawley

Swiss mice
ICR/Ha
                                           1 mg, 3x/wk,  581 d
                                           topical
                                           1 mg,  3x/wk,  14 d
                                           2.5 ug phorbol
                                           myristate acetate,
                                           topical  452 d

                                           0.5 mg sc/wk, 622 d

                                           0.5 mg,  once  wk
                                           gavage,  622 d

                                           1 mg  TCI  epoxide,
                                           3x/wk, 2.5 pg
                                           phorbol  myristate
                                           acetate,  topical,
                                           452 d

                                           2.5 mg TCI epoxide
                                           3x/wk, topical  for
                                           526 d

                                           0.5 mg TCI epoxide,
                                           once wk,  sc,  547  d

                                           Gavage,  104 wk
Inhalation, 78 wk


Inhalation, 104 wk


Gavage, 78 wk
Negative



Negative




Negative

Negative


Negative





Negative



Negative


In preparation





In press





Negative
                                                    8-3

-------
mass spectroscopy, infrared spectrophotometry, and nuclear magnetic resonance
spectroscopy.  Analytical results showed a TCI purity of greater than 99.9%.
Epichlorohydrin was not found by gas chromatography-mass spectroscopy at a
detection level of 0.001% (v/v).  The TCI product contained 8 ppm of
diisopropylamine as a stabilizer.  Refrigeration at 4°C created no analytically
detectable decrease in purity over the course of product usage.
     Stock solutions of TCI in corn oil  were made once weekly for the first 16
weeks and once monthly for the remainder of the bioassay.   Analysis by gas
chromatography showed that these solutions were stable for 7 days at room
temperature and for 4 weeks at 4°C.  Stock solution levels of TCI were within
10% of nominal concentrations.
     Based on survival, body weights, and pathology in a preliminary 13-week
dose-finding study, estimated maximally tolerated and one-half maximally
tolerated doses were selected for the bioassay.  For the bioassay, 50 males
and 50 females per group were given nothing (untreated controls), corn oil
alone (vehicle-controls), 500 mg/kg/day of TCI in corn oil (low dose), or 1,000
mg/kg/day of TCI in corn oil (high dose).  Rats were dosed daily, 5 days per
week, for 103 weeks.  The volume of administered corn oil  was 1 ml.
     Rats were randomly assigned to dose groups when 8 weeks old at the start
of the study.  The animals were caged in groups of 5, and  decedents were
replaced during the first 1.5 weeks.  Temperature and humidity in the animal
quarters were 22°C to 24°C and 40% to 60%, respectively; room air was exchanged
10 to 15 times every hour.
     The animals were observed daily.  Body weights were recorded weekly
for the first 12 weeks and monthly thereafter.  Each animal was necropsied,
and tissues and organs, as well as all lesions and tumors, were evaluated
histopathologically.
                                    8-4

-------
     Terminal sacrifice of survivors occurred at 111 to 115 weeks.
     Body weight trends in Figure 8-1 show reduced mean weight gains in high-dose
males and in both treatment groups of females.  A dose-related effect on mean
body weights in females became evident after 60 weeks, and mean weight gain
between vehicle-control and low-dose males was similar (mean body weights at the
start of the study were 161 and 141 g for vehicle-control  and low-dose males,
respectively).
     A dose-related reduction in survival of treated male  rats and a smaller
decrease in survival of high-dose females were evident through survival  pro-
babilities estimated by the Kaplan and Meier (1958) technique (Figure 8-2).
Differences in survival among female groups were not significant (p < 0.05);
however, survival was significantly (p < 0.05) lower in low-dose males
(p = 0.005) and high-dose males (p = 0.001) compared to vehicle-control  males.
The following numbers of rats which were killed by "gavage error" were censored
from the survival curves in Figure 8-1:   1 vehicle-control  male, 3 low-dose
males, 10 high-dose males, 2 vehicle-control  females, 5 low-dose females, and 5
high-dose females.  The numbers of rats  alive at terminal  sacrifice were as
follows:  35/50 vehicle-control males, 20/50 low-dose males, 16/50 high-dose
males, 36/50 vehicle-control  females, 33/50 low-dose females, and 26/50 high-
dose females.
     It is indicated in the NTP (1982) report that, because of reduced survival
in treated male rats, the statistical  methods used which adjust for intercurrent
mortality (life table analysis, incidental  tumor test) would provide more mean-
ingful results, statistically, than unadjusted statistical  methods (Fisher
Exact Test, Cochran-Armitage Trend Test).  Kidney tumor data in Table 8-2 show
a significant (p < 0.05) increase in kidney tubular adenocarcinoma incidence in
high-dose males (3/16)  compared to vehicle-control  males (0/33) at terminal
                                     8-5

-------
500
                                        0
                                        O
                                          ODD
              *
                                                2
          n n
          o°
          A A
              MALE RATS
              D VEHICLE CONTROL
              O LOW DOSE
              A HIGH DOSE
          10
                 I
                 20
                       30
500-
400-
  Till
 40     50     60     70
TIME ON STUDY (WEEKS)
80
      90
             100
                   110
                                              Q  Q  aonao D  D
                                         fi
                                              2
                          O n O O  O  O
                     °AAAAAAA
               FEMALE RATS
               D VEHICLE CONTROL
               O LOW DOSE
               A HIGH DOSE
-»—
0
I
10
I
20
T
30
I
40
I
50
I
60
I
70
1
80
1
90
1
100
110
                             TIME ON STUDY (WEEKS)
                                      8-6

-------
                                PROBABILITY OF SURVIVAL
PROBABILITY OF SURVIVAL
CO
I
c
g
o •
oo <-+ -n
o ~s -••
r- -i.tQ
^ 0 C -
0 3- -5 <"
m — ' n>
o
-s oo
O 1
s  ro
— 1 r+ •
=0 3- W
„ << O'

H^ (t) C
VD 3 ~S
oo n> < _,
rs: _, ~1' ^
13 QJ m -^
.Ij m en
o 0
o o z
3 ^ «
< c
on) n _
_, /rt *— ' 03
^ <" -< 0'
-h <
0-0 s
o- -* m
^ m
"=> ~5 r^
pi pj ** '
^ . ^J
S CT w
Cd
fD BJ
• Q.
3
— '•
"^ co
• o
CO
fD
-s
ro
Q.
O_
•
01
ro
0
c
c






































> c
' c



>0

0 0
o o
































J C
1 0
3 C



n ^
m ^
I >
PS
m 33
S >

Z CO
—\
•J}
o
1—


























3 C
i f
3 C






































) C
> U
3 C






































' C
3 S
































t?

t^



c
S































^•*'






3 C
31 ?

























^
[>'
t*'
1^*


!
^^ II
i^l^
ml



3 C
3 C£
3 C




















1
t>




>*" [

rf^
fV«

1
iP






3
1 <











9
D-l


C

fc. •

T'
>
l~l
M
1
|
JL
?
j










c
D C
D C



_ ^


CO




H
^ -^

0
CO
~H
o ^
-< °
g
m
m
7t
^— ' >j






CO






en


) C
>
) C






































3 C
' N
3 C



>0
II
8S
m m































3 C
} O
3 C



D 2
3-
± m
P 3D
m J>
r^ ^

~P
~H
X
O
"~


























3 C
3 *
3 C






































> C
> CJ
) C































t?






> C
: g

























i
1^"
>.•



o6"5
'





3 C
I C


















J^
J
t^ *

P"
' :
^**


f
,

•*C"f^^



rfl
B


3 C
J g















^i "
:
^^*





c
C>-.
/-v ^
IT
J




-rfP
CH
r



3 C
3 C£
3 C











>^'
i

»:





> O
> 0
o









91



9>
'T
i
QJ
1
y
ol




i_i
j












-------
              TABLE 8-2.  ANALYSIS OF PRIMARY TUMORS IN MALE RATS3
KIDNEY:  TUBULAR-CELL ADENOCARCINOMA
Tumor rates
vehicle-
control
Low
dose
High
dose
     Overall13
     Adjusted0
     Terminal
Statistical  tests6

     Life table
     Incidental tumor test
     Cochran-Armitage Trend,
       Fisher Exact Tests
0/48 (0%)
0.0%
0/33 (0%)
p = 0.009
p = 0.009

p = 0.038
0/49 (0%)
0.0%
0/20 (0%)
(f)
(f)

(f)
3/49 (6%)
18.8%
3/16 (19%)
p = 0.028
p = 0.028

p = 0.125
KIDNEY:  TUBULAR-CELL ADENOMA OR ADENOCARCINOMA
Tumor rates
Vehicle-
control
Low
dose
High
dose
     Overall13
     Adjusted0
     Terminal
Stati sti ca I  tests6^

     Life table
     Incidental tumor test
     Cochran-Armitage Trend,
       Fisher Exact Tests
0/48 (0%)
0.0%
0/33 (0%)
p = 0.019
p = 0.030

p = 0.084
2/49 (4%)
5.6%
0/20 (0%)
p = 0.194
p = 0.327

p = 0.253
3/49 (6%)
18.8%
3/16 (19%)
p = 0.028
p = 0.028

p = 0.125
       groups received doses of 500 or 1,000 mg/kg of trichlorethylene by
 gavage.
^Number of tumor-bearing animals/number of animals examined at  the site.
cKaplan-Meier estimated lifetime tumor incidence after adjusting  for
 intercurrent mortality.
^Observed tumor incidence at terminal  kill.
eBeneath the control  incidence are the p-values associated with the trend test.
 Beneath the dosed group incidence are the p-values corresponding to pairwise
 comparisons between  that dosed group  and the controls.  The life table analysis
 regards tumors in animals dying prior to terminal kill as being  (directly or
 indirectly) the cause of death.  The  incidental tumor test regards these
 lesions as non-fatal.  The Cochran-Armitage Test and the Fisher  Exact Test
 compare the overall  incidence rates directly.
^The configuration of tumor incidences precludes the use of this  statistic.

SOURCE:  NTP, 1982.
                                     8-8

-------
sacrifice by life table (p = 0.028) and incidental  tumor (p = 0.028) tests.
Each test for linear trend was significant (p < 0.05), as shown in Table 8-2.
Historical corn oil control data for male Fischer 344 rats evaluated in the NTP
bioassay program as of June 15, 1981 for studies of at least 104 weeks showed a
renal tumor incidence of 3/748 (0.4%) adenocarcinomas; thus, this tumor type is
rare in this strain of rats.  However, the comparison for kidney adenocarcinomas
between high-dose and vehicle-control male rats was not significant (p < 0.05)  by
the Fisher Exact Test, and the statistical significance achieved with the other
tests above reflects the significantly (p _< 0.005)  higher mortality in high-dose
males as compared to vehicle-control males.  Other  observed kidney tumors in rats
include a transitional cell carcinoma of the renal  pelvis in a low-dose male,
tubular adenomas in two low-dose males, a carcinoma (NOS) of the renal pelvis in
a high-dose male, a transitional cell papilloma of  the renal pelvis in an
untreated control male, and a tubular adenocarcinoma in a high-dose female.
     Renal tubular adenocarcinomas had a general  histopathologic appearance of
solid sheets of cells, often atypical, and small  areas of necrosis.  These
tumors commonly invaded surrounding normal tissue.   Small renal  tubular adenomas
consisted of solid collections of tubular epithelial  cells filling contiguous
tubules.  In larger adenomas there was greater evidence of cellular atypia, and
there was no sharp demarcation between larger adenomas and adenocarcinomas.
Renal adenocarcinomas were distinguished from renal  adenomas by basement membrane
invasion.
     Toxic nephrosis characterized as cytomegaly was found in 98 treated male
rats and 100 treated female rats, but not in control  rats.   This lesion was
found in early deaths, and it progressed in severity during the course of the
study.  Enlargement of tubular cells was detected early and progressed to the
extent that some tubules were dilated and hard to identify.  With longer sur-

                                      8-9

-------
vival  the number of dilated cells  decreased,  corresponding tubules became di-
lated, and the basement  membrane had  a  striped  appearance.  The most advanced
stage of cytomegaly extended into  the renal cortex.   Cytomegaly was graded as:
1 - slight, a subtle change in  a limited  area of  the  kidney; 4 - severe, an
obvious lesion which involved a substantial part  of the  kidney, and which could
alter its function; 2 -  moderate;  3 - well marked.  Average grades for each
group of rats were:  0 for male and female controls,  2.8 for low-dose males,
3.1 for high-dose males, 1.9 for low-dose females, and 2.7 for high-dose
females.  In contrast, renal nephropathy, a common spontaneous lesion in aging
rats, was of lower incidence in treated males (88% in untreated controls, 85%
in vehicle controls, 59% in low-dose, and 37% in  high-dose) which could relate
to the higher mortality  in treated males.
     Increases of tumor  incidences at other sites in  treated rats were not
apparent.  Malignant mesothelioma  incidence was higher in low-dose males
[vehicle-controls, 1/50; low-dose, 5/50 (p =  0.042, life table test); high-
dose, 0/48], but high-dose animals did  not develop this  tumor type, and the
low-dose incidence was not significantly  (p < 0.05) increased according to the
other statistical tests  used in this  study.
     The TCI doses used  in this study produced  statistically significant de-
creases in the survival  of male rats, and a decreased survival in high-dose
female rats compared to matched vehicle-controls  was  apparent, although this
difference was not found to be statistically  significant. By definition  (NCI,
1976) the maximum tolerated dose  in carcinogenicity bioassays is the highest
dose that can be administered during  the  chronic  study which will not be expec-
ted to alter the animals' survival rate from  effects  other than carcinogenicity,
Thus, TCI doses beyond that maximally tolerated,  particularly in male rats,  ap-
pear to have been used in this study.  The toxicity of TCI in the kidney tubule
                                     8-10

-------
region, as cytomegaly, approximated the "well-marked"  grade on  the average in
both groups of treated male rats and the high-dose group of female rats,  and it
has been noted that the severity of this lesion increased with  longer survival.
Although a small  but statistically significant excess  of renal  tubular adeno-
carcinomas in high-dose males compared to vehicle-control  males was calculated
by statistical methods which adjusted for mortality, the NTP and its Board of
Scientific Counselors have concluded that this study in Fischer 344 rats  is
inadequate for a judgment on the carcinogenicity of TCI.  The following is
quoted from the latest version (received May 24, 1983) of the abstract on this
study from the NTP:
          Under the conditions of these studies, epichlorohydrin-free
          trichloroethylene caused renal tubular-cell  tumors in male F344/N
          rats, produced nephropathy in both sexes, shortened the survival
          times of males and did not produce a carcinogenic response among
          females.  This experiment was considered to  be inadequate to
          evaluate the presence or absence of a carcinogenic response to
          trichloroethylene in male F344/N rats.

Use of additional lower doses in this study could have provided a broader
evaluation of dose-response for the observed toxic effects of TCI.

8.1.1.2  National Cancer Institute (NCI, 1976)—A carcinogenicity study of TCI
in Osborne-Mendel rats was reported by the National Cancer Institute (NCI)
in 1976.  The composition of the TCI sample used, determined by vapor-phase
chromatography-mass spectrometry, was as follows:  TCI, >L 99.0% (w/w);
1,2-epoxybutane, 0.19%; ethyl acetate, 0.04%; N-methylpyrrole,  0.02%;
diisobutylene, 0.03%; epichlorohydrin, 0.09%.  The rats were 48 days old  when
the study began.  Treatment groups included 50 animals/sex/dose, and 20 animals/
sex comprised matched vehicle-control (corn oil) groups.  Ninety-nine male and
98 female rats used as vehicle "colony controls" were  also compared with  treat-
                                     8-11

-------
ment groups in this study.  Vehicle "colony controls" included matched controls



for TCI and other chemicals being tested concurrently.



     Maximum and one-half maximum tolerated doses for the carcinogenicity study



were estimated from survival, body weight, and necropsy results from a preli-



minary 8-week subchronic test.  In the carcinogenicity study, treated male and



female rats were given time-weighted average doses, by gavage, of 549 and 1,097



mg/kg/treatment of TCI 5 days/week for 78 weeks.  Doses were presented as



time-weighted averages due to the dose level changes made during the study, as



shown in Table 8-3, on the basis of apparent toxicity from treatment with TCI.



     Rats were allowed to survive until  terminal sacrifice at 32 weeks post-



treatment.  Animals were observed daily for mortality and toxic signs, and



body weights and food consumption were recorded weekly for the first 10 weeks



and monthly thereafter.  Survivors and decedents were necropsied, and tissues



and organs, including tumors and lesions, were examined histopathologically.



     At the beginning of the study, rats were randomly assigned to treatment



and control groups according to body weight.  Body weight trends in Tables 8-4



and 8-5 show decreased'body weight gain in males and females in both treatment



groups compared to matched vehicle-controls.  Food consumption trends in Tables



8-4 and 8-5 indicate decreased food intake by both groups of treated female



rats as compared to matched vehicle-controls, whereas food intake by treated



male rats appeared comparable to that of matched vehicle-controls.  Although



decreased body weights in treated animals could reflect decreased food con-



sumption, it would appear that decreases in food consumption and body weights



ultimately were a  result of toxicity from treatment with TCI.   For example,



food consumption between matched control and low-dose males was approximately



equal, whereas body weight gain was less in the latter group.
                                      8-12

-------
 TABLE 8-3.   DOSAGE AND OBSERVATION SCHEDULE  -  TRICHLOROETHYLENE  CHRONIC  STUDY
                             ON OSBORNE-MENDEL  RATS

Dosage
group
Low-dose
males and
females


High-dose
males and
females


Dose (mg
TCI/kg
body wt)
650
750
500
500
no treatment
1300
1500
1000
1000
no treatment
Percent of
TCI in
corn oil
60.0
60.0
60.0
60.0

60.0
60.0
60.0
60.0
60.0
Age at
dosing3
(weeks)
7
14
23
37
85
7
14
23
37
85
Treatment
period
(weeks)
7C
9c
14C
48d
32
7C
9c
14c
48d
32
Time-weighted
av. dose& (mg
TCI/kg body wt)




549




1097
Matched controls received doses of corn oil  by gavage calculated  on  the basis
of the factor for the high dose animals.

aAge at initial  dose or dose change.
bTime-weighted average dose = I (dose in mg/kg x no.  of days at that dose)/
 I (no. of days receiving any dose).   In calculating  the time-weighted average
 dose, only the days an animal  received a dose are considered.
cDosing 5 days per week each week.
dDosing 5 days per week, cycle of 1 week of  no treatment followed by 4 weeks  of
 treatment.  (Animals were treated for 38 weeks of the 48-week  period.)

SOURCE:  NTP, 1982.
     Survival patterns in Figures 8-3 and 8-4 indicate a rather poor survival

of animals in this study, particularly in treated animals where a decline in

survival was evident early in the study.  A dose-related decrease in survival

of treated male rats was observed; however, survival  of low-dose females was

less than that of high-dose females early in the study.  Statistical analysis

of survival patterns in the NCI (1976) report showed  a significant (p < 0.05)

decrease in the survival of high-dose males (p = 0.001) vs.  matched control
                                      8-13

-------
                  TABLE 8-4.  MEAN BODY WEIGHTS,  FOOD  CONSUMPTION,  AND SURVIVAL  -  TRICHLOROETHYLENE
                                              CHRONIC  STUDY  -  MALE  RATS
Vehicle-control
Time
interval
(weeks)
0
00
£ 14
26
54
78
110
Body
Mean
(g)
193
504
570
616
559
382
weight3
Std. dev.
(g)
15.0
39.4
43.3
50.2
76.7
26.9
Foodb
(g)
0
153
156
149
153
91
No. of
animals
weighed
20
20
20
20
16
2
Body
Mean
(g)
193
474
533
573
523
383
Low dose
weight
Std. dev.
(g)
15.8
40.4
47.1
47.6
67.3
101.5
Food
(g)
0
150
158
147
158
114
No. of
animals
weighed
50
50
47
40
31
8
Body
Mean
(g)
194
446
500
513
462
423
High dose
weight
Std. dev.
(g)
16.7
45.3
47.2
48.3
47.2
45.3
Food
(g)
0
140
150
134
148
255
No. of
animals
weighed
50
45
43
30
12
3
Calculated using individual  animal  weight.
bAverage weight per animal  per week.

SOURCE:  NCI, 1976.

-------
                  TABLE 8-5.   MEAN BODY WEIGHTS, FOOD CONSUMPTION, AND SURVIVAL - TRICHLOROETHYLENE
                                            CHRONIC STUDY - FEMALE RATS
Vehicle-control
00
1
1— >
in




Time
interval
(weeks)
0
14
26
54
78
110
Body
Mean
(g)
146
302
351
388
373
326
weight3
Std. dev.
(g)
11.4
33.8
37.6
51.3
58.2
80.1
Foodb
(g)
0
107
124
128
146
119
No. of
animals
weighed
20
20
19
17
16
8
Body
Mean
(g)
144
271
293
315
317
311
Low
weight
Std. dev.
(g)
11.0
24.4
28.8
29.8
39.9
86.0
dose
Food
(g)
0
98
110
120
145
114

No. of
animals
weighed
50
44
34
28
20
13

Body
Mean
(g)
144
262
286
307
317
311
High
wei ght
Std. dev.
(g)
9.5
26.3
30.8
38.6
43.5
67.7
dose
Food
(g)
0
104
106
113
135
136

No. of
animals
weighed
50
47
45
37
23
13
Calculated using individual  animal  weight.
bAverage weight per animal  per week.

SOURCE:  NCI, 1976.

-------
                                                             PROBABILITY OF SURVIVAL
                                                                                                              oo
                                                                                                                          CO
t/1
o
cr
vo

en
oo
 i
cr>
c

o -n
-h -•• ...
to c °
C -5
-S fD
-"• CO
< 1 CO
-a
r+ -5 ±»
X' °- -1
DC —
3- 0 ^
— ' t"1" m
° Lo w
o -i o-
JH.CO
=T c-t- H
<< C
fD C/1 .< O
fD — '• "^
i 3 <
el- a) m
fD fD ^ O
OJ to (/)
fD O
O- -h
00.
3-0 O
a> -s
— • o
ro a"
a>
-S CT CD
0) -J. OH
t/i — '•
• H-

_.


t>
I
I
0
o
in










>.



1
6 A
o 8
O 33
O O
C/0 l~








>.,
^...


u








>1"
f
rs...
,.,'r


) T j
.-1








>'1?>J
..[>•'














>....!




o-P~° i
i









»-^"
' *



V'











•>-> 	

i
r>-
O- J
r>- —^
Y

^J—









£>--: '"
»* * *
(
n_j
^— ~^










1
>..;
....
•
Y
_J











>.>.'•-
* * v
o- J
O- J
1
I«J



"P —









-------
                                                                      PROBABILITY OF SURVIVAL
                                                                                                                              00
CO
o
o
10

CTl
 00
tO  TJ
C  -J.
-s ua
<  c
-"• -S

OJ
—• oo
   I
 I  4S.

d-
-s

o  -s
3- o
—' Q.
O  C
-i  O
O  <-h
n>  i

3" -••
     O> c+
     3
     ro n>
n>  3
QJ  OJ

fD  fD


-h O
§H,

0> X3
—' -S
n>  o
   CT
-s  a>
OJ  CT
   .. \      >   9


             |>   o
m

9 «.
z o
            3  «.
            -<  o

            m
            en
               CO
               o
               «>,
               o
               o
               o
                         1   r-   o
                         n   9   O
                         O   O   33
                         o   o   o
                         en   en   i-
                         m   rn
                                                   -6M-
                                                                           J,*"'
                                                                                     ?J
                                                                                                            >.l
                                                                                                                                      >•
                                                                                                                                   •H-i-

-------
males and of both high-dose females (p = 0.028)  and low-dose females  (p  =
0.049) vs. matched control  females.  Survival  at terminal  sacrifice was  as
follows: 3/20 matched control  males,  8/50 low-dose males,  3/50 high-dose males,
8/20 matched control  females,  15/48 low-dose females (2 missing females  were
excluded from the denominator),  and 13/50 high-dose females.
     A carcinogenic effect  of  TCI  was not discerned in this  study.  Total tumor
incidences expressed as the number of rats with  tumors were  as follows:  5/20
matched control  males, 7/50 low-dose  males, 5/50 high-dose males,  7/20 matched
control females, 2/48 low-dose males, and 12/50  high-dose  females.  No statis-
tically significant (p < 0.05) differences in  tumor incidence between  matched
control and treatment groups were  found.
     Chronic nephropathy was common in treated animals of  both sexes  and both
doses.  This nephropathy was described as slight to moderate degenerative and
regenerative changes in tubular epithelium.  This nephropathy was  stated to be
unlike the chronic nephropathy encountered as  a  commonly occurring  lesion in
aging rats and recognized frequently  in treated  and control  rats in this study.
The occurrence of unilateral malignant mixed tumors in kidneys of  two  matched
control males, a hamartoma  in  the  kidney of one  low-dose male and  one  matched
control male, and an adenocarcinoma in the kidney of one low-dose  male were
noted.  High incidences of  chronic respiratory disease were  found  in  each
treated group as well as in matched controls.
     Under the conditions of this  study, TCI was not carcinogenic  in  Osborne-
Mendel rats.  However, animal  survival was rather low, particularly in treated
animals, and greater survival  during  the course  of the study might  have  per-
mitted a stronger evaluation for carcinogenicity.  Reduction in survival and
body weight gain indicate that the TCI doses used were toxic to the animals,
and a broader evaluation of dose response could  have been  possible if  addi-
                                     8-18

-------
tional lower doses had been tested and if dose levels had not been adjusted
during the study.  Rats treated with TCI, and their controls, were housed in a
room with other rats treated with dibromochloropropane, ethylene dichloride,
1,1-dichloroethane, or carbon disulfide.  There were 10 to 15 air exchanges
each hour in the room, but ambient air was not tested for the presence of
volatile materials.

8.1.1.3  Maltoni (1979)—Maltoni (1979) summarized the results of a study done
to estimate the carcinogenic potential of TCI in Sprague-Dawley rats.  Thirty
males and 30 females, 13 weeks old, were assigned to each group.  Two treatment
groups received 50 or 250 mg/kg of TCI by gavage daily, 4 to 5 days weekly for
52 weeks, and a vehicle-control group was given olive oil.  The TCI sample was
at least 99.9% pure and was contaminated with £ 50 ppm each of 1,2-dichloro-
ethylene, chloroform, carbon tetrachloride, and 1,1,2-trichloroethane.  No de-
tectable epoxides were found in this test material, and 10 ppm diisopropylamine
was added as a stabilizer.  Surviving animals were sacrificed at 140 weeks.
All animals were given a necropsy examination, and tissues and organs were
evaluated microscopically.
     A carcinogenic effect was not demonstrated in this study.  However,  al-
though the experiment lasted 140 weeks, the animals were dosed for 52 weeks,
which is a duration below potential  lifetime exposures.  Tumor findings were
presented in the study report; however, mortality patterns and toxic signs, if
any, were not described to indicate the sensitivity of the animals to the TCI
treatment used.  Use of animals younger than 13 weeks of age at the beginning
of the study could have given a broader assessment of the carcinogenic poten-
tial  of TCI through an evaluation of treatment with TCI during the early  growth
and development of the rats.
                                     8-19

-------
8.1.2  Oral Administration (Gavage):   Mouse
8.1.2.1  National  Toxicology Program (NTP, 1982)--A carcinogenicity bioassay on
TCI in B6C3F1 mice was recently completed for the National  Toxicology Program,
and a 1982 draft report of the study has been reviewed by the NTP's Board of
Scientific Counselors.
     The TCI test  sample and the experimental design were those used in the
NTP (1982) carcinogenicity bioassay in Fischer 344 rats previously described
in this document.   Evaluation of mortality, body weights, and pathology in a
13-week preliminary dose-finding study led to the selection of 1,000 mg/kg/day,
5 days per week, for 103 weeks as the single dose level used in the bioassay.
Fifty males and 50 females per group were assigned to untreated control groups,
corn oil vehicle-control groups, and treatment groups, when 8 weeks old.
Vehicle-control and treated animals received 0.5 ml corn oil with each gavage
administration.  Mice were caged in groups of 10 for the first 8 months of the
study, and in groups of 5 for the rest of the study.  The animals were evaluated
for mortality, body weight, and pathology until  terminal sacrifice at 112 to
115 weeks.
     Results of this study were analyzed by the  statistical methods used for
the NTP (1982) carcinogenicity bioassay in rats.  Mean body weights of treated
males were lower than those of vehicle control males, whereas mean body weights
between vehicle control and treated females were comparable (Figure 8-5).  Sur-
vival was significantly (p = 0.004) lower in treated males compared to vehicle-
controls, and although survival in treated females was lower after 95 weeks, the
overall difference in survival between vehicle-control and treated females was
not significant (p < 0.05) (Figure 8-6).  Deaths of two vehicle-control males,
three treated males, one vehicle-control female, and three treated females were
attributed to "gavage error," and these animals  were censored from the survival
                                      8-20

-------
   50
   40'
I. 30.
01
Q
O
m
z  20-





gib0





t
§o c

MALE
ODOSE


D
0 0


MICE
M c PONI!
DGROU

D
DD

O O
O



p
D
D
rt
O




3 D
[
C
j




D D
0
0






D '
O Q(






' D D

°0





D
D C
0 C
O











   10'
      0     10     20    30    40    50    60    70    80    90    100    110

                             TIME ON STUDY (WEEKS)
40-
5
5 30-
UJ
0
§ M1
z
111



,£10
n^
F1





TD a (

FEMAL
D VEHK
ODOSE


n D


E MICE
:LE CONI
D GROU

D
°o °



3

,e






i
D 8 c






1






BB'






'og






o n












10     20     30    40    50    60    70

                 TIME ON STUDY (WEEKS)
                                                      80    90    100    110
             Figure 8-5.  Growth  curves for mice administered

             trichloroethylene  in corn oil by gavage.
             SOURCE:  NTP, 1982.   8-21

-------
                                     PROBABILITY OF SURVIVAL
                                                                                                     PROBABILITY OF SURVIVAL
CO
o
o
m
 CO
 ro
co
ro
ro
      o
      n>
   c
   -i
  1 n>

   co
      fD
   01
   c
   -5
O
O  O
-i  C
3  1


o  n>
      cr o
      «< -s

      to 3
      ai _i.
      < o
      a> ro

      (D OJ
      •  Q.
         n>
         -s
         ro
         Q.
0 0 0 0 C
8 — k) CO *
0 0 0 C






CO

-1 4.,
m
O
Z
en


m






o

OD-
1
1
R
to x >
m o r~
O f-: m
30 O _
O O O
C Z m













O








































) C
' C
) C

















(

> c
3 g

















:

3 C
3 S
















, 0<
c
H
c
J I















f
P
IP

> c
3 c














8'
"-b



— •
) O
) O










D
cP


UP
p





                                                                                                o

                                                                                                N>
                                                                                                                 O
                                                                                                                 bn
                                                                                                                 O
p
^
o

ID







8-

H
m
O
cn

o o"
_
m
m




o


Ol
O




O P §



o < >
o cp i-
cn ^ m
rn o -,
D r- 2


Ss?


c z
















o
r~




















































0
o
























o























rt
cP
























) O


F




















d^
o
_
p
p
^


















0°l
0
p

D


















O
n
P
P
O
p










-------
patterns shown in Figure 8-6.  At terminal  sacrifice,  33/50 vehicle-control
males, 16/50 treated males, 32/50 vehicle-control  females,  and  23/50 treated
females were alive.
     A carcinogenic effect of TCI was not apparent in  this  study except  for a
significant (p _£ 0.002) increase in the incidence  of hepatocellular carcinomas
in treated male and female mice compared to corresponding vehicle-controls
(Tables 8-6 and 8-7).  Hepatocellular carcinomas metastasized to the lung  in
five treated males and one vehicle-control  male.  Hepatocellular adenoma inci-
dence was significantly (p < 0.05) increased in treated female  mice by each
statistical method indicated in Table 8-7,  but in  treated male  mice only by
life table analysis (Table 8-6).  Spontaneous hepatocellular tumor formation  is
a characteristic of the B6C3F1 mouse strain, and incidences of  hepatocellular
carcinoma in historical corn oil control  B6C3F1 mice as of June 15, 1981 for
carcinogenicity studies of 104 weeks duration in the NTP bioassay program  were
18% (120/256) in males and 2.9% (22/751)  in females.
     Hepatocellular carcinomas had a general histopathologic appearance  of
markedly abnormal cytology and architecture, with  diffuse patterns of tumor
tissue infiltrating normal tissue, whereas  hepatocellular adenomas consisted of
circumscribed areas of distinct parenchyma  cells.   A treatment-related effect
on nonneoplastic pathology in liver was not evident in the study report.
     Cytomegaly in the kidney was found in  90% of  treated male  and 98% of  treat-
ed female mice, but not in corresponding vehicle-controls.   The average  grades
of severity for cytomegaly, according to the scheme used for grading cytomegaly
in kidneys of treated Fischer 344 rats in the previously described NTP (1982)
study, were 1.5 in dosed males and 1.8 in dosed females.  Comparison with  the
average grades for male and female Fischer  344 rats given 1,000 mg/kg/day  of
TCI in the NTP (1982) bioassay indicates that toxic nephrosis of the kidney
                                     8-23

-------
              TABLE 8-6.   ANALYSIS  OF  PRIMARY  TUMORS  IN  MALE  MICE3
LIVER:   ADENOMA
Tumor rates
Overallb
Adjustedc
Termi nal^
Statistical tests6
Life table
Incidental tumor test
Cochran-Armitage Trend,
Fisher Exact Tests
Vehicle-
control
3/48 (6%)
9.1%
3/33 (9%)
Dosed
8/50 (16%)
27 .5%
1/16 (6%)
p = 0.022
p = 0.191
p = 0.113

LIVER:  CARCINOMA
Tumor rates
Overall0
Adjustedc
Terminald
Statistical tests6
Vehicle-
control
8/48 (17%)
22.1%
6/33 (18%)
Dosed
30/50 (60%)
92.5%
14/16 (87.5%)
     Life table
     Incidental tumor test
     Cochran-Armitage Trend,
       Fisher Exact Tests
p < 0.001
p < 0.001

p < 0.001
LIVER:  ADENOMA OR CARCINOMA
Tumor rates
Overall
Adjusted0
Terminal^
Statistical tests6
Vehicle-
control
11/48 (23%)
30.7%
9/33 (27%)
Dosed
38/50 (76%)
97.1%
15/16 (94%)
     Life table
     Incidental tumor test
     Cochran-Armitage Trend,
       Fisher "xact Tests
p < 0.001
p < 0.001

p < 0.001
 aThe dosed group  received doses of 1,000 mg/kg of trichlorethylene by gavage.
 ''Number of tumor-bearing animals/number of animals examined at the site.
 cKaplan-Meier estimated lifetime tumor incidence after adjusting for intercurrent
 mortality.
 ^Observed tumor incidence at terminal kill.
 eBeneath the dosed  group incidence are the P-values corresponding to pairwise
 comparisons between that dosed group and the controls.  The life table analysis
 regards tumors in  animals  dying prior to terminal kill as being (directly or
 indirectly) the  cause of death.  The incidental tumor test regards these
 lesions as non-fatal.  The Fisher Exact Test compares the overall incidence
 rates directly.

 SOURCE:  NTP,  1982.
                                    8-24

-------
             TABLE 8-7.  ANALYSIS Oc PRIMARY TUMORS IN FEMALE MICE3
LIVER:  ADENOMA
Tumor rates
Vehicle-
control
Dosed
     Overall*3
     Adjusted0
     Terminal
2/48 (4%)
6.2%
2/32 (6%)
Statistical tests6

     Life table
     Incidental tumor test
     Cochran-Armitage Trend,
       Fisher Exact Tests
8/49 (16%)
30.2%
6/23 (26%)
                         p = 0.016
                         p * 0.023

                         p = 0.049
LIVER:  CARCINOMA
Tumor rates
Overall6
Adjusted0
Terminal^
Statistical tests6
Vehicle-
control
2/48 (4%)
6.2%
2/32 (6%)
Dose
13/49 (27%)
43.9%
8/23 (35%)
     Life table
     Incidental tumor test
     Cochran-Armitage Trend,
       Fisher Exact Tests
                         p < 0.001
                         p = 0.002

                         p = 0.002
LIVER:  ADENOMA OR CARCINOMA
Tumor rates
Overallb
Adjusted0.
Terminal^
Statistical tests6
Vehicle-
control
4/48 (8%)
12.5%
4/32 (13%)
Dosed
19/49 (39%)
63.8%
13/23 (57%)
     Life table
     Incidental tumor test
     Cochran-Armitage Trend,
       Fisher Exact Tests
                         p < 0.001
                         p < 0.001

                         p < 0.001
aThe dosed group received doses of 1,000 mg/kg of trichlorethylene by gavage.
Number of tumor-bearing animals/number of animals examined at  the site.
°Kaplan-Meier estimated lifetime tumor incidence after adjusting for intercurrent
 mortality,
''Observed tumor incidence at terminal  kill.
6Beneath the dosed group incidence are the P-values corresponding to pairwise
 comparisons between that dosed group  and the controls.  The life table analysis
 regards tumors in animals dying prior to terminal kill as being (directly or
 indirectly) the cause of death.  The  incidental tumor test regards these
 lesions as non-fatal.  The Fisher Exact Test compares the overall incidence
 rates directly.

SOURCE:  NTP, 1982.
                                    8-25

-------
was less severe in the mice in this study at an equivalent  dose.   Only one dosed
mouse had a cytomegaly grade above 2.   One vehicle-control  male and  one treated
male each had a renal  tubular cell adenocarcinoma.
     Under the conditions of this bioassay, TCI treatment  produced statistically
significant increases  of hepatocellular carcinomas  in male  and female B6C3F1
mice.  These results repeated those of the NCI  (1976) carcinogenicity study
discussed herein, showing statistically significant increases  in  the incidence
of hepatocellular carcinomas in male and female B6C3F1 mice given TCI stabilized
with epoxides.  In effect, the repeat  study excludes the epoxides as necessary
factors in the increased incidence of  hepatocellular carcinomas in male and
female B6C3F1 mice treated with TCI under the conditions of the NTP (1982) and
NCI (1976) bioassays.   Use of additional doses  in the NTP  (1982)  study could
have provided an evaluation of dose response for hepatocellular carcinomas in
mice.

8.1.2.2  National Cancer Institute (NCI, 1976)--A carcinogenicity study of TCI
in B6C3F1 mice was reported by the National Cancer Institute (NCI) in 1976.
The TCI sample was that used in the NCI (1976)  study in Osborne-Mendel rats
described herein.  Treatment groups included 50 animals/sex/dose, and 20
animals/sex comprised  matched vehicle-control (corn oil) groups.   Seventy-seven
male and 80 female mice used as vehicle "colony controls"  and  70 male and 76
female mice serving as untreated "colony controls"  were also compared with the
treatment groups in this study; "colony controls" included  matched controls for
TCI and other chemicals being tested concurrently.   Mice were  35 days old when
the study began.
     Maximum and one-half maximum tolerated doses for the carcinogenicity study
were estimated from survival, body weight, and necropsy results from a preliminary
                                     8-26

-------
8-week subchronic test.   In the carcinogenicity study, time-weighted average doses

of TCI were given by gavage, 5 days/week, for 78 weeks, to the mice at levels

of 1,169 and 2,339 mg/kg/treatment in males and 869 and 1,739 mg/kg/treatment

in females.  Doses were  presented as time-weighted averages due to the dose

level changes made during the study, as shown in Table 8-8, in an attempt to

maintain maximum and one-half maximum tolerated doses.
        TABLE 8-8.  DOSAGE AND OBSERVATION SCHEDULE - TRICHLOROETHYLENE
                          CHRONIC STUDY ON B6C3F1 MICE
Dosage
group
Low-dose
males


High-dose
males


Low-dose
females

High-dose
females


Dose (mg Percent of
TCI/kg TCI in
body wt) corn oil
1000
1000
1200
no treatment
2000
2000
2400
no treatment
700
900
no treatment
1400
1400
1800
no treatment
15.0
10.0
24.0

15.0
20.0
24.0

10.0
18.0

10.0
20.0
18.0

Age at
dosing3
(weeks)
5
11
17
83
5
11
17
83
5
17
83
5
11
17
83
Treatment
period
(weeks)
6C
6C
66C
12
6C
6C
66C
12
12C
66C
12
6^
6C
66C
12
Time-weighted
av. dose" (mg
TCI/ kg body wt)



1169



2339


869



1739
Matched controls received doses of corn oil by gavage calculated on the basis
of the factor for the high dose animals.

aAge at initial dose or dose change.
bTime-weighted average dose = I (dose in mg/kg x no. of days at that dose)/
 I (no. of days receiving any dose).  In calculating the time-weighted average
 dose, only the days an animal received a dose are considered.
cDosing 5 days per week each week.
 SOURCE:   NCI, 1976.
                                      8-27

-------
     Mice were permitted to survive until  terminal  sacrifice at 12 weeks fol-



lowing treatment.  Animals were observed daily for mortality and toxic signs,



and body weights and food consumption were recorded weekly for the first 10



weeks and monthly thereafter.  Survivors and decedents were necropsied, and



tissues and organs, including tumors and lesions, were examined histopatho-



logically.



     At the beginning of the study, mice were randomly assigned to treatment



and control groups according to body weight.  Body weight trends in Tables 8-9



and 8-10 show comparable weight gain between matched vehicle-controls  and both



treatment groups for both males and females.  Food consumption trends  in Tables



8-9 and 8-10 indicate similar consumption  between the female groups and slight-



ly higher consumption in treated males as  compared to matched controls.  There-



fore, it is possible that, if food consumption between matched control  and



treated groups of males had been equivalent, body weight gain in treated males



might have been less than that in matched  control males.



     Survival patterns are presented in Figures 8-7 and 8-8.  Differences in



survival reported as statistically significant (p < 0.05) were those between



high-dose and low-dose males (p = 0.001),  matched control males and low-dose



males (p = 0.004), and both treatment groups of females and matched control



females (p = 0.035).  Survival at terminal  sacrifice was as follows:  8/20



matched control males, 36/50 low-dose males, 22/50 high-dose males, 20/20



matched control females, 42/50 low-dose females, and 42/47 high-dose females  (3



high-dose females were missing and not included in the denominator).



     Prior to terminal sacrifice, approximately 50% of the treated males and  "a



few" treated females exhibited bloating or abdominal distension.  Results of



necropsy and histopathologic examinations  revealed a statistically significant



(p < 0.05) increase in the incidence of hepatocellular carcinomas in male and
                                     8-28

-------
                TABLE 8-9.  MEAN BODY  WEIGHTS,  FOOD  CONSUMPTION,  AND  SURVIVAL  -  TRICHLOROETHYLENE
                                            CHRONIC  STUDY  -  MALE  MICE
Vehicle-controls
Time
interval
(weeks)
!° o
o
D
14
26
54
78
90
Body
Mean
(g)
17
28
31
32
34
34
weight3
Std. dev.
(g)
0.5
0.1
0.2
0.4
0.6
0.7
Foodb
(g)
0
23
24
21
24
34
No. of
animals
weighed
20
20
20
18
8
8
Body
Mean
(g)
17
28
31
33
34
33
Low dose
wei ght
Std. dev.
(g)
2.0
0.8
0.9
1.0
0.8
0.7
Food
(g)
0
27
28
26
32
32
No. of
animals
wei ghed
50
50
48
45
40
35
Body
Mean
(g)
17
29
32
34
35
34
High dose
weight
Std. dev.
(g)
1.1
0.4
0.7
0.4
1.0
1.3
Food
(g)
0
26
27
30
39
38
No. of
animals
weighed
50
49
42
33
24
20
Calculated using individual  animal  weight.
^Average weight per animal  per week.

SOURCE:  NCI, 1976.

-------
                 TABLE 8-10.  MEAN BODY WEIGHTS, FOOD CONSUMPTION, AND SURVIVAL - TRICHLOROETHYLENE
                                             CHRONIC STUDY - FEMALE MICE
Vehicle-controls
00
1
CO
0




Time
interval
(weeks)
0
14
26
54
78
90
Body
Mean
(g)
14
23
25
28
29
28
weight3
Std. dev.
(g)
0.0
0.4
0.2
1.1
0.6
0.6
Foodb
(g)
0
27
22
19
23
28
No. of
animals
wei ghed
20
20
19
18
18
17
Body
Mean
(9)
14
23
26
27
29
30
Low dose
weight
Std. dev.
(9)
0.6
0.3
0.7
0.5
0.7
0.6
Food
(g)
0
23
23
21
25
27
No. of Body
animals Mean
weighed (g)
50 14
49 24
49 25
45 26
41 28
40 28
High dose
weight
Std. dev.
(g)
0.7
0.5
0.5
0.3
0.4
0.7
Food
(g)
0
24
24
22
26
26
No. of
animals
weighed
50
49
47
41
40
39
Calculated using individual animal  weight.
^Average weight per animal per week.

SOURCE:  NCI, 1976.

-------
                                                                             PROBABILITY OF SURVIVAL
I/O
o
o
uo

CT>
00
I
CO
      -S CQ
      < C
      ->• -s

      £D
      —' C»
           -s
           ->• -a
           o  -s
           zr o
           —' a.
           o  c:
           -s  o
           O  r+
           CO  I
           r+ —'
           ^" —'•
          «<  3
           —i —i.
           n>  H-
       1
         fD
         LO
          -s  -"•
          ro  3
          QJ  Q,
          r+  H-
          rD  fD
          a.  co

          3  o
          *  -h

          fD ~O

          3  o
          ->.  cr
          O  fu
          fD  cr
         o
         -h
                  s-
m

O
z
V)

c
o
               m
                  §
                  8'
                  8
                               X   r-

                               Q   <
               §OO
               Ui   v>   r-
               m   m
ff

-------
PROBABILITY OF SURVIVAL
O —
CO tO T|
o c -••
c: -s ua ^_
PC < c °
O -'• -S
m < ro
cu
— ' 00
1 ,sj
"Z. 1 00 O~
o
i— i r+
• -s
-j. -a
1— » O -S
^O 13" O t*>
CT> O C
-s o
O t-t- H
ro i 5
<-*• — ' s ^
?i! o °
co ro «-h 2
i 3 cn
co ro ro — i
ro i to C $5-
c-l- rt- 0 °
-$-••-<
ro 3 —
CU QJ ^
ST SJ" m
ro ro m o>
Q- */> ^ O~
-+,0 —
ft> -a
— • -s ?J_
ro o o
-j. cr
o ->•
m i: oo.
^ °
^<
o
-h
CO.
o
o.
o


f *!
:
> -.^^ '
V
i
j

i



-------
female mice treated with TCI  (Table 8-11).   A dose-related  response  in  females
is evident.  Metastasis of hepatocellular carcinoma to the  lung was  observed in
four low-dose males and in three high-dose  males.   The time to observation  of
the first hepatocellular carcinoma in each  group was as follows:   72 weeks  in
matched control  males, 81 weeks in low-dose males,  27 weeks in high-dose  males,
and 90 weeks in females.  With regard to "colony control" mice, hepatocellular
carcinomas were diagnosed in  5/77 (6.5%) vehicle-control  males, 5/70 (7.1%)
untreated control  males, 1/80 (1.3%) vehicle-control females,  and 2/76  (2.6%)
untreated control  females.  Hepatocellular  carcinoma incidences in "colony
control" mice were comparable to those in matched vehicle-control males (1/20
or 5%) and females (0/20 or 0%).  Statistically significant (p < 0.05)  diffe-
rences in the incidence of other tumor types between control and treatment
groups were not found.
     The general pathology of hepatocellular carcinomas consisted of various
characteristics, including an orderly cord-like arrangement of neoplastic cells,
a pseudoglandular pattern resembling adenocarcinoma, or sheets of highly  ana-
plastic cells with minimal cord- or gland-like arrangement.
     Under the conditions of  this study, TCI induced a statistically signi-
ficant (p < 0.05)  increase in hepatocellular carcinoma incidence in  male  and
female B6C3F1 mice.  Although the number of matched vehicle controls was  low,
the use of pooled colony controls gives additional  support  for treatment-related
effects.
     A more precise estimate  of dose response perhaps could have been obtained
if additional lower doses had been used and if constant doses  rather than
time-weighted averages had been used.  Treated animals were housed in the
                                      8-33

-------
     TABLE 8-11.   INCIDENCE OF HEPATOCELLULAR  CARCINOMAS  IN  B6C3F1  MICE  RECEIVING  TRICHLOROETHYLENE BY ORAL GAVAGE
CO

Dose 0 mg/kg
Number with
tumors/number
of animals 1/20
Incidence 0.05
p-value
Number with
metastases 0/20
Males
1169 mg/kg
26/50
0.52
a
1.39 x 10-4
4/50

2339 mg/kg 0 mg/kg
31/48 0/20
0.646 0
a
3.44 x 10~6
3/48 0/20
Females
869 mg/kg 1739 mg/kg
4/50 11/47
0.08 0.234
a
1.75 x 10-1 1.35 x 10-2
0/50 0/47
  Statistically  significant  (p  <  0.05)  increase compared to matched vehicle-controls  (0 mg/kg) by  Fisher  Exact
   Test.

  SOURCE:   NCI, 1976.

-------
same room as mice treated with other volatile compounds*;  however,  since 1)

controls were in the same room as treated animals,  2)  oral  TCI  doses probably

would have been much higher than ambient levels of  other volatiles, 3)  the

cages had filters to limit the amount of chemical  released into the ambient

air, 4) the total room air was exchanged 10 to 15  times per hour,  and 5) dosing

was done in another room under a large hood, the likelihood that the other

volatile compounds were responsible for the observed results is considered low.

Ambient levels of volatiles in the animal quarters  were not measured.  Further

evidence that the induction of hepatocellular carcinomas in B6C3F1 mice was  due

to treatment with TCI in this study is the repeat  study (NTP, 1982) with a

similar experimental design in which the mice were  not housed with mice treated

with other volatile chemicals.  Additionally, the  repeat NTP (1982) study of

purified TCI in B6C3F1 mice devoid of detectable levels of epoxides indicates

that the presence of epoxides was not necessary for the induction of hepatocel-

lular carcinomas in B6C3F1 mice in the NCI (1976)  study.


8.1.2.3  Van Duuren et al. (1979)t--A carcinogenicity study of purified TCI  in

ICR/Ha Swiss mice was described by Van Duuren et al. (1979).  Experimental

details provided below include those obtained through written communication

with B.L. Van Duuren, New York University.  The TCI sample was > 99% pure, and,

although contaminants were not identified, epoxides were considered not to be

present in the test material.  It was stated that  although range-finding studies
*l,l,2,2-tetrachloroethane, 3-chloropropene, chloropicrin, 1,1-dichloroethane,
 chloroform, sulfolene, iodoform, ethylene dichloride, methyl chloroform, 1,1,2-
 trichloroethane, tetrachloroethylene, hexachloroethane, carbon disulfide,
 trichlorofluoromethane, carbon tetrachloride, ethylene dibromide, dibromo-
 chloropropane.
tAlthough discussed under the heading of Oral Administration;  Gavage (Mouse),
 this study in mice also includes subcutaneous injection, repeated skin applica-
 tion, and initiation-promotion experiments with TCI and an initiation-promotion
 experiment with TCI oxide.


                                      8-35

-------
were done to estimate maximum tolerated doses (doses which did  not  affect  body
weight gain or induce clinical  signs of toxicity in a 6- to 8-week  test  which
included histopathologic examination of treatment site sections)  for all test
groups in this study, TCI doses used in the main topical  application experi-
ments were below those which could have been tolerated by the animals.   TCI
doses used in the skin treatment experiments were selected to permit a compa-
rison with TCI oxide, with respect to carcinogen!city, on an equimolar basis.
This study incorporated the following experiments:   1) 30 females received
topical applications of 1 mg TCI three times weekly for approximately 581  days;
2) 30 females were treated topically with a single application  of 1.0 mg TCI
followed 14 days later by applications of 2.5 yg phorbol  myristate  acetate
(PMA) to the skin three times weekly until  termination of the study at 452 days.
In addition, a different group of 30 females was similarly treated  with  an ap-
plication of 1.0 mg TCI epoxide, a metabolite of TCI, followed  by applications
of 2.5 ng PMA; 3) 30 females were given subcutaneous injections of  0.5 mg  TCI
once weekly for 622 days; and 4) 30 males and 30 females were administered 0.5
mg TCI by gavage once Weekly for 622 days.   A vehicle-control group of 30  mice
and an untreated control group of 100 mice were included in each experiment.
Dosing vehicles were acetone in the skin treatment tests and trioctanoin in the
other tests.  In the initiation-promotion experiment, 210 mice  treated with PMA
alone were also on test.  The mice were 6 to 8 weeks old at the beginning  of
the study; six animals were housed in each cage.  Test sites on the skin were
shaved as necessary and were not covered, and test samples were administered  in
ventilated hoods; however, it was the authors' impression that  TCI  was  immedi-
ately absorbed and that evaporation from test sites was minimal.  The animals
were weighed monthly, and each animal was examined by necropsy.  Tumors  and
lesions were examined histopathologically.  Histopathologic evaluation was

                                      8-36

-------
routinely done on skin, liver, stomach, and kidney in the skin  application
experiments, injection site tissue and liver in  the subcutaneous  injection
experiment, and stomach, liver, and kidney in the gavage  experiment.
     Treatment with TCI and TCI oxide did not produce statistically  significant
(p < 0.05) increases in tumor incidence in this  study (Tables 8-12,  8-13,  and
8-14).  In the initiation-promotion experiment,  the numbers  of  mice  with skin
papillomas (squamous cell  carcinomas) were:  4 (1) treated with TCI,  15  (3)
treated with PMA alone, and none in the control  groups.   Repeated application
to the skin produced lung and stomach tumors in  9 and 1 TCI-treated  mice,
respectively, 11 and 2 vehicle-treated controls, respectively,  and 30 and  5  un-
treated controls, respectively.  Lung tumors were diagnosed  as  benign papillo-
mas, and stomach tumors were described as papillomas of the  forestomach.
Tumors were not found in TCI-treated and control animals  in  the experiment with
subcutaneous doses.  In the experiment with gavage dosing, forestomach tumors
were observed in one female and two male mice given TCI,  no  vehicle-control
animals, and five female and eight male (one male had a squamous  cell  carcinoma
in the forestomach) untreated control mice.  Three mice given repeated applica-
tions of TCI oxide onto the skin had skin papillomas. Results  in Tables 8-12,
8-13, and 8-14 show that the mice were sensitive to positive control  agents
under the design of this study.   Median survival  times approximated  the number
of days the animals were on test, and body weights were comparable between
treatment and control  groups in each experiment.
     Use of higher TCI doses could have more strongly challenged  the  mice  for
carcinogenicity in the skin treatment experiments, but Van Duuren et  al. (1983)
noted that low activity or lack of activity of chloroolefins tested  for carci-
nogenicity in skin treatment experiments in their laboratory was  expected be-
cause of evidence for the relatively low level of epoxidizing enzymes in mouse
                                      8-37

-------
                 TABLE 8-12.  MOUSE SKIN BIOASSAYS OF TRICHLOROETHYLENE AND TRICHLOROETHYLENE OXIDE
                                       (30 FEMALE HA:ICR SWISS MICE PER GROUP)
                        Initiation-promotion3
                                                          Repeated application'3
    Compound
Dose, mg/     Days to   Mice with        Dose, mg/    Days to   Mice with         No. of mice
application/   first    papillomas/     application/   first    papillomas/          with
   mouse       tumor  total  papillomas     mouse       tumor  total papillomas  distant tumors^
Trichloroethylene   1.0
Trichloroethylene
 oxide              1.0

7,12-Dimethyl-
                264    4/9 (l)c
371
                       3/3
                           1.0
NT
                                       9 lung
                                       1 stomach
benz[a]anthracenee 0.02
00
1
co
00 PMA controls
(120 mice) 0.0025
( 90 mice) 0.005
Acetone (0.1 mL)

No treatment
(100 mice)
54 29/317 (18) NT
p < 0.0005f


141 9/10 (1)
449 6/7 (2)
0.1 mL

_ « _ —

-





0 11 lung
2 stomach
0 30 lung
5 stomach
aCompounds were applied to dorsal skin in 0.1 mL acetone once followed 14 days later with 2.5 mg PMA
 in 0.1 mL acetone 3 times weekly.
^Compounds applied in 0.1 mL acetone 3 times weekly.   NT = Not tested.
cNumber of mice with squamous cell carcinomas are in  parenthesis.
dLung tumors were benign papillomas, and stomach tumors were papillomas of the forestomach.
ePositive control agent.
fStatistically significant by chi-square test compared to controls.
SOURCE:  Van Duuren et al., 1979.

-------
                              TABLE 8-13.   SUBCUTANEOUS INJECTION OF TRICHLOROETHYLENE
                                        (30 FEMALE HA:ICR SWISS MICE PER GROUP)

Compound and
dose3
Trichloroethylene, 0.5
B-propiolactone,d 0.3
•J0 Trioctanoin, 0.05 mL
CO
No treatment (100 mice)

Days on
test
622
378
631

649
No. of mice
with local
sarcomas'3
0
24
0

0

P value0
NS
<0.0005
_

-
aCompound subcutaneously injected in  left  flank  once weekly in 0.05 ml  trioctanoin.   Dose is mg per
 injection per mouse.
bHistopathologically characterized as fibrosarcomas.
CNS = Not significant.  Statistical  significance calculated by chi-square test compared to controls.
dPositive control  agent.

SOURCE:  Van Duuren et al., 1979.

-------
                         TABLE 8-14.  CARCINOGENICITY OF TRICHLOROETHYLENE BY INTRAGASTRIC FEEDING  IN

                                     MALE AND FEMALE HA:ICR SWISS MICE (30 MICE PER GROUP)
oo
I

Compound and
dosea
Trichloroethylene, 0.5

B-Propiolactone, 1.0

Trioctanoin, 0.1 mL

No treatment
(100 mice)
(60 mice)

Sex
M
F
M
F
M
F

M
F

Days on
test
622
622
539
582
631
635

649
636
No. of mice with
forestomach
tumors'5
1
2
24(23)
25(23)
0
0

5
8(1)

P value0
NS
NS
<0.0005
<0.0005
_
-

_
-
      aDose in mg per intubation per mouse.   Compounds were given once weekly in 0.1 mL trictanoin.

      ^Number of mice with squamous cell carcinoma of the forestomach is given in parentheses.

      CNS = Not significant.  Statistical significance calculated by chi-square test compared to controls.


      SOURCE:  Van Duuren et al., 1979.

-------
skin relative to other organs such as liver.   Administration of TCI more than
once each week in the experiments with oral and subcutaneous treatment also
might have more strongly challenged the animals for carcinogenicity.  Dr. Van
Duuren noted (written communication) that once-weekly subcutaneous injection
would allow time for the injection site to heal before subsequent injections;
however, more than 1 injection per week might have more effectively tested the
carcinogenic potential of TCI at sites distant from the injection site.  Routine
histopathologic examination of a greater range of tissues and organs, both from
the mice dosed orally as well as from the mice dosed by subcutaneous injection,
might have provided a broader evaluation for carcinogenicity.

8.1.3  Inhalation Exposure;  Rats
8.1.3.1  Henschler et al. (1980)—Henschler et al. (1980) described a study on
the carcinogenic potential  of TCI in Han:Wist rats.  The initial  age of the
animals was not given.  Thirty male and 30 female rats were assigned to each of
three dose groups which were exposed to 0 (controls exposed to inhalation
chamber air only), 100 (0.55 mg/L air), or 500 (2.8 mg/L air) ppm TCI vapor 6
hours daily, 5 days/week, for 18 months.  The animals were individually caged
in this study.  Surviving rats were sacrificed at 36 months.  All  animals were
subjected to necropsy and histopathologic examination.
     The TCI sample, stabilized with 0.0015% triethanolamine, contained the
following impurities by gas chromatography-mass spectrometry analysis:   chloro-
form, carbon tetrachloride, 1,1,2-trichloroethane, 1,1,1-trichloroethane, and
1,2,2-tetrachloroethane, each present at less than 0.0000025% (w/w).   Epoxides
were not detected in this test material.  Exposure levels of TCI  in the inhala-
tion chambers were monitored with direct ultraviolet spectrophotometry, and the
flow rate of the TCI air mixture was adjusted to yield a threefold  turnover of
                                     8-41

-------
the total volume each hour.  A description of the analytical  method was provi-
ded by D. Henschler, University of Wurzburg, as follows:   Exposure levels were
automatically controlled to 100 or 500 ppm, as appropriate,  and were continu-
ously monitored in a gas-flow cell in an ultraviolet beam of 205 mm.  The
accuracy of this method was determined by the noise of the electronic equip-
ment, and the registered band indicated +_ 7% of the median at 100 ppm.
     No effect of TCI on body weight gain was observed.  A statistically signifi-
cant (p < 0.05) decrease in survival of treated rats was  not found (Table 8-15).
     A carcinogenic effect of TCI was not evident in this study.  Use of a
maximum tolerated dose is not indicated in the study report, and the com-
parability of survival and body weight between control and treatment groups
indicates that higher exposure levels of TCI possibly could  have been used to
provide a broader evaluation of TCI carcinogenicity.

8.1.3.2  Bell et al. (1978; audit of Industrial Bio-Test  Laboratories, Inc.
study)--A carcinogenicity study of TCI vapor in Charles River rats was con-
ducted at Industrial Bio-Test Laboratories, Inc. (IBT) from  1975 to 1977, and
audit findings by the Manufacturing Chemists Association  (MCA)  have been repor-
ted (Bell et al., 1978).  The TCI product used in this study contained, on a
weight percent basis, TCI, > 99%; diisobutylene, 0.023%;  butylene oxide, 0.24%;
ethyl acetate, 0.052%; N-methylpyrrole, 0.008%; and epichlorohydrin, 0.148%.
Each of the three treatment groups plus an untreated control  group consisted of
120 rats/sex.  Treated animals were exposed to nominal concentrations of 100,
300, or 600 ppm (0.55, 0.165, or 0.330 mg/L air) TCI vapor 6 hours/day, 5
days/week, for 24 months.  Surviving animals were sacrificed on termination of
treatment.  Thirty rats/sex scheduled for interim sacrifice  to evaluate hema-
tology and clinical chemistry were not given pathologic examinations.
                                      8-42

-------
        TABLE 8-15.  ESTIMATION OF PROBABILITY OF SURVIVAL OF RATS (%)a
Group
Control
Week
1
50
75
100
115
126
156
M
100.0
96.6
93.3
86.7
83.3
66.7
46.7
F
100.0
100.0
96.6
86.7
70.0
56.6
16.7
100
M
100.0
96.6
96.6
90.0
86.6
80.0
23.3
ppm
F
100.0
96.6
93.3
86.6
63.3
50.0
13.3
500
M
100.0
100.0
100.0
90.0
76.6
66.6
36.7
ppm
F
100.0
93.3
90.0
76.6
50.0
50.0
16.7
aBy the Kaplan and Meier (1958) method.
SOURCE:  Henschler et al., 1980.

     According to the audit report by the MCA, there were noticeable deficien-
cies in the performance of this study.  Examination of analytical  data on the
TCI vapor concentrations in the inhalation chambers revealed wide  deviations
from the nominal  concentrations and high variability in the number of measure-
ments made daily which do not allow a precise determination of the actual
exposure levels.   For example, histograms of measurements (Figures 8-9, 8-10,
and 8-11) made during the initial  120 days of exposure show a clustering of
data around the nominal  concentrations; however, only 259/565, 393/623, and
389/598 readings  made during the first 6 months of the study fell  within _+ 30%
of the 100, 300,  and 600 ppm nominal  concentrations, respectively.  Exposure
data in the report by Bell  et al.  (1978) cover the initial  6 months of this
                                      8-43

-------
00
I
                                                                                                    I   I   I   I
          0  20 40 60 80 100 120 140 160 180200 220240 260280 300320 340360 380400 420440 460 480 500520 540560 580600 620640 660680 700

                                                            PPM                                          MAX. VALUE

                                                                                                         2040 ppm
                   Figure 8-9.   Trichloroethylene histogram of  chamber concentration measurements

                   T-l (100  ppm).


                   SOURCE:   Bell et  al., 1978.

-------
oo
I
en
         0 20 40  60  80 100 120140 160 180 200220240 260280 300320 340360 380 400420440 460480 500520 540560 580600 620 640660 680700 +

                                                         PPM                                           MAX. VALUE

                                                                                                        1676 opm
                 Figure  8-10.  Trichloroethylene histogram  of chamber concentration  measurements

                 T-l (300 ppra).



                 SOURCE:   Bell et  al., 1978.

-------
                                        917-8
                                                      FREQUENCY
GO
O
O
CO
rt>
t-t-
00
    c

 
         (T)
         to
         r+
         O
         O)
         O
         3-
         CU
         3
         CT
         ro
         -s

         O
         O
         3
         O
         (D
         CU
         fD
         CU
         fD

-------
study, and additional information from the Chemical Manufacturers Association
has been requested regarding exposure levels obtained during the final 18
months of this study.
     Histopathologic reexamination of livers did not show a carcinogenic effect
from exposure to TCI.  Hepatocellular carcinomas were found in 3/99 control
male and 3/99 low-dose male rats.  Histopathologic reexamination of other
tissues was not evident.

8.1.4  Inhalation Exposure:  Hamsters
8.1.4.1  Henschler et al. (1980)—Henschler et al. (1980) conducted a carci-
nogenicity study on TCI in Syrian hamsters.  The TCI sample and the experimen-
tal conditions were those used in the carcinogenicity study of TCI in Han:Wist
rats by Henschler et al. (1980) discussed herein.  Thirty male and 30 female
hamsters per exposure group were exposed to 0 (controls exposed to inhalation
chamber air only), 100, or 500 ppm TCI vapor 6 hours daily, 5 days per week,
for 18 months.  The initial age of the animals was not given.  Terminal  sacri-
fice was at 30 months.  All animals were subjected to necropsy and histopatho-
logic examination.
     No effect of TCI on body weight gain was observed.  A treatment-related
effect on survival of hamsters was not found (Table 8-16); however, a marked
increase in the death rate for all groups ensued after 75 weeks.
     A carcinogenic effect of TCI was not evident in this study.  Use of a
maximum tolerated dose is not indicated in the study report, and the similar-
ity of survival  and body weight between control  and treatment groups indicates
that higher exposure levels of TCI possibly could have been used to provide a
broader evaluation of TCI carcinogenicity.
                                   8-47

-------
8.1.5  Inhalation Exposure:  Mice



8.1.5.1  Henschler et al. (1980)--Hensch1er et al.  (1980)  conducted a carcino-



genicity study on TCI in Han: NMRI mice.  The TCI  sample and the experimental



conditions were those used in the carcinogenicity  study of TCI in Han:Wist rats



by Henschler et al. (1980) discussed herein.  Thirty male and 30 female mice



per exposure group were exposed to 0 (controls exposed to inhalation chamber



air only), 100, or 500 ppm TCI vapor 6 hours daily, 5 days per week, for 18



months.  The initial  age of the animals was not given.  Terminal sacrifice was



at 30 months.  All animals were subjected to necropsy and histopathologic



examination.







      TABLE 8-16.  ESTIMATION OF PROBABILITY OF SURVIVAL OF HAMSTERS (%)a
Week
1
50
75
100
115
126
156

Control
M F
100.0 100.0
100.0 100.0
100.0 83.6
56.7 23.3
40.0 6.7
13.3
-
Group
100 ppm
M F
100.0 100.0
100.0 100.0
100.0 89.7
63.3 20.7
26.7
6.7
-

500
M
100.0
100.0
100.0
70.0
40.0
26.7
-

ppm
F
100.0
90.0
76.6
30.0
-
-
-
aBy the Kaplan and Meier (1958) method.



SOURCE:  Henschler et al., 1980.
                                     8-48

-------
No effect on body weight gain was observed.  Survival patterns described in
Table 8-17 show a greater mortality in treated mice compared to controls, par-
ticularly after 50 weeks in males and 75 weeks in females.  A statistically
significant (p < 0.05) decrease in survival rates of males and females in both
treatment groups compared to controls was reported.  The survival  patterns
indicate that the TCI exposure levels used were toxic to male mice; increased
mortality in treated females was particularly evident during the period when
lymphoma incidence was increasing in these groups.
        TABLE 8-17.  ESTIMATION OF PROBABILITY OF SURVIVAL OF MICE (%)a
Group
Control
Week
1
50
75
100
115
126
156
M
100.0
93.3
83.3
66.7
33.3
23.3
-
F
100.0
93.3
83.3
46.7
33.3
26.7
-
100
M
100.0
86.7
63.3
36.6
16.6
10.0
-
ppm
F
100.0
96.6
73.3
30.0
13.3
3.3
-
500
M
100.0
83.3
56.6
26.7
13.3
3.3
-
ppm
F
100.0
100.0
82.8
37.9
10.3
3.4
-
aBy the Kaplan and Meier (195.8)  method.
SOURCE:  Henschler et al.t  1980.
                                     8-49

-------
     A carcinogenic effect of TCI was not evident in male mice in  this  study;
however, the incidence (number affected/number examined)  of malignant lymphomas
in female mice was as follows:  9/29 (control), 17/30 (100 ppm), and  18/28
(500 ppm).  The response in the low (p = 0.042) and  high  (p =  0.012)  dose
groups represents a statistically significant  (p < 0.05)  increase  over  the
control incidence by the Fisher Exact Test.  The time-to-tumor occurrence was
shorter in treated females; however, the majority of lymphomas were apparent
after treatment was terminated at 78 weeks,  as illustrated in  Figure  8-12.
Induction of hepatocellular carcinomas by exposure to TCI was  not  observed  in
this study.  Hepatocellular adenoma was diagnosed in one  control and  two low-
dose male mice, and hepatocellular carcinoma was discovered in one control  male
mouse.
     According to the authors, the increased incidence of malignant lymphoma in
female mice could have been due to an occurrence of  immunosuppression capable
of enhancing the susceptibility of tumor induction by specific, mostly  inborn,
viruses.  The 30% incidence of malignant lymphoma in matched control  female
mice in the Henschler et al. (1980) study is higher  than  the 12 to 22%  (average
16%) range of lymphoma incidence determined  by Luz (1977) between  1968  and  1977
for 830 untreated female NMRI-Neuherberg mice  used as control  groups  in long-
term experiments.  It was the authors'  impression that it was  not  clear whether
an immunosuppressive effect would have been  exerted  by TCI itself  or  by other
nonspecific influences, such as stress.
     Under the conditions described in the report of their study,  Sanders et al,
(1982) indicated an ability of orally administered TCI to reduce the  immune
status in CD-I mice; therefore, enhancement  of the lymphoma incidence in the
treated female mice possibly could have been related to an immunosuppressive
effect from TCI treatment.  Kreuger (1972) presented an argument similar to
                                     8-50

-------
    20
    15
    10'
LL)
03
D  CONTROLS

A  100ppm

O  500 ppm

                                          O
                                        Y
                                                 \  i
                                                            s
                                                           s
                                                                 t
          36 40     50     60     70    80     90     100    110    120    130

                                   WEEKS



        Figure  8-12.  Incidence of malignant  lymphomas  in female mice.


        SOURCE:  Henschler et al., 1980.
                                    8-51

-------
that of Henschler et al.  (1980)  in which a coexistence of persistent  antigenic
stimulation and chronic immunosuppression is pathogenetically related to malig-
nancy in immunocompetent  tissues.   Furthermore,  Kreuger (1972)  indicated that  a
significant increase in the incidence of lymphoreticular neoplasms  corresponded
to a significantly shortened latent period of tumor formation in several experi-
ments cited in his report.  The  argument of Henschler et al.  (1980),  therefore,
provides a possible interpretation of the results of their study.

8.1.5.2  Bell et al. (1978; audit  of Industrial  Bio-Test Laboratories, Inc.
study)—A carcinogenicity study  of TCI vapor in  B6C3F1 mice was conducted at
Industrial Bio-Test Laboratories,  Inc. (IBT) from 1975 to 1977, and audit
findings by the Manufacturing Chemists Association (MCA) have been  reported
(Bell et al., 1978).  The TCI test sample and the experimental  design were
those used in the IBT study in Charles River rats discussed elsewhere herein.
Each of three treatment groups (exposed to nominal levels of 100,  300, or 600
ppm TCI vapor, 6 hours daily, 5  days/week, for 24 months), as well  as an un-
treated control group, consisted of 140 mice of  each sex.  Except  for 40 mice/
sex/group assigned to interim sacrifice for evaluation of clinical  chemistry
and hematology, a pathologic examination was scheduled for each animal.  Sur-
viving animals were sacrificed on termination of treatment.
     Serious flaws in the conduct of this study  were described in  the audit
report by the MCA.  Wide deviations in actual exposure levels compared to
nominal concentrations and high  variability in the number of vapor measurements
made daily,  as found in laboratory records made  early in the study, are des-
cribed in the discussion of the IBT carcinogenicity study in rats.   Furthermore,
although all mice evidently were received from the same supplier,  control mice
appear to have been selected from a shipment obtained approximately 3 weeks
                                     8-52

-------
earlier.  Hence, the control  group was not specifically matched  with  the  treat-
ment groups.  Histopathologic reexamination of livers  revealed 39  discrepancies
between gross and microscopic observations, which were mostly the  result  of
improper sectioning of liver  containing tumors found grossly. Twelve mice
(<_ 3 per group) were identified as originally missexed during reexami nation  of
liver specimens.  The sex of  nine mice (<_ 2 per group) could  not be determined,
and for the tabulation of liver tumor data, the sex shown  on  laboratory  records
was assumed.
     Statistical analysis of  liver tumor data from reexamination of liver
pathology was done by the Fisher Exact Test and the chi-square analysis  (Table
8-18).  Significant (p < 0.05) increases in the incidences of hepatocellular
adenoma, hepatocellular carcinoma, and hepatocellular  adenoma/hepatocellular
carcinoma combined in treated males compared to untreated  males  and the  inci-
dence of hepatocellular adenoma/hepatocellular carcinoma combined  in  treated
females compared to untreated females were calculated  as shown in  Table  8-18.
With respect to the treatment groups which were on study concurrently, statis-
tically significant (p < 0.05) increases in both the number of hepatocellular
carcinomas alone and hepatocellular adenomas/hepatocellular carcinomas combined
in high-dose vs. either low-  or mid-dose males and in  hepatocellular  carcinomas
in high-dose vs. low-dose females were found.  Increases in liver  tumor  inci-
dence in high-dose compared to low-dose mice may be related to clustering of
exposure level  frequencies around nominal  concentrations,  as  shown in Figures
8-9, 8-10, and  8-11 in the discussion of the IBT rat study, and  increased hepa-
tocellular carcinoma incidence was evident in B6C3F1 mice  treated  with TCI by
gavage in the NTP (1982) and  NCI (1976) bioassays discussed in this document.
However, this study is weakened by the deficiencies in its conduct described
above.  A recently completed  carcinogenicity study by  Dr.  C.  Maltoni  and
                                      8-53

-------
                                             TABLE 8-18.
                                                          STATISTICAL ANALYSIS OF THE INCIDENCE OF PRIMARY HEPATOCELLULAR NEOPLASMS
                                                               IN CONTROL AND TRICHLOROETHYLENE TREATED B6C3F1 MICE
CO
I



Hepatocellular adenoma
Incidence
Percent
Chi square
Fisher Exact Test
Hepatocellular carcinoma
Incidence
Percent
Chi square
Fisher Exact Test
Hepatocellular adenoma/
hepatocellular carcinoma
combi ned
Incidence
Percent
Chi square
Fisher Exact Test



Hepatocellular adenoma
Incidence
Percent
Chi square
Fisher Exact Test
Hepatocellular carcinoma
Incidence
Percent
Chi square
Fisher Exact Test
Hepatocellular adenoma/
hepatocellular carcinoma
combined
Incidence
Percent
Chi square
Fisher Exact Test

Untreated
controls

2/99
2.02%



18/99
18.18%





20/99
20.20%



Untreated
controls

2/99
2.02%



6/99
6.06%





8/99
8.08%




T-ia

7/95
7.37%
2.042
0.0752

28/95
29.47%
2.821
0.0463*



35/95
36.84%
5.815*
0.00778*


T-l

5/100
5.00%
0.572
0.2265

4/100
4.00%
0.116
0.8382



9/100
9.00%
0.000
0.5083
Males

T-2b

7/100
7.00%
1.820
0.0873

31/100
31.00%
3.741
0.0262*



38/100
38.00%
6.794*
0.004407*
Females

T-2

1/94
1.06%
0.002
0.8672

9/94
9.57%
0.1418
0.2604



10/94
10.64%
0.132
0.3579


T-3C T-l vs. T-2 T-l vs. T-3 T-2 vs. T

10/97
10.31%
4.504* 0.032 0.214 0.329
0.0150* 0.6465 0.3223 0.2834

43/97
44.33%
14.431* 0.006 3.930* 3.184
0.000063* 0.4698 0.0235* 0.0371*



53/97
54.64%
23.408* 0.000 5.427* 4.836*
0.00000049* 0.4933 0.0098* 0.0138*


T-3 T-l vs. T-2 T-l vs. T-3 T-2 vs. T-3

4/99
4.04%
0.172 1.364 0.000 0.719
0.3413 0.9824 0.7461 0.2007

13/99
13.13%
2.096 1.599 4.205* 0.303
0.0730 0.1026 0.0188* 0.2916



17/99
17.17%
2.930 0.020 2.250 1.211
0.0427* 0.4427 0.0662 0.1353
                         *Significant at 95% level.
                         a T-l: 100 ppm
                         b T-2: 300 ppm
                         c T-3: 600
                         SOURCE:  Bell et a!., 1978.

-------
associates can provide further evaluation of the carcinogenic potential  of TCI
vapor in B6C3F1 mice.
8.1.6  Other Carcinogenicity Evaluations of Trichloroethylene--Although  done
mainly to estimate general  toxic effects of TCI, as discussed elsewhere  in this
document, the studies summarized in Table 8-19, except for that by Laib  et al.
(1978), were described by Waters et al.  (1977) as studies in which carcinogenic
activity for TCI was not demonstrated.   However, these studies involved  small
groups of animals and durations of treatment and observation which were  short
with respect to the lifetime of the animals or, in the study by Rudali  (1967),
a duration which was unspecified.
     Laib et al. (1979, 1978) histochemically evaluated livers from Wistar rats
exposed to either vinyl chloride (VC) or TCI, by the protocol shown in  Table
8-19, for focal deficiencies of nucleoside triphosphatase (ATPase) as evidence
for pre-malignant hepatocellular lesions.  The authors stated that Friedrich-
Freksa had shown that focal deficiencies of distinct hepatocellular enzymes
precede formation of hepatocellular malignancies.  There was no effect  in adult
female rats exposed to either chemical,  and ATPase-deficient foci  were  found in
livers from newborn rats exposed to VC,  but not in livers from newborn  rats
exposed to TCI.  The authors concluded  that this experiment did not indicate an
oncogenic potency for TCI in hepatocytes of newborn rats.  However, it  is not
certain whether a longer duration of exposure of newborn rats to TCI could
ultimately have resulted in enzyme-deficient foci in liver.

8.1.7  Trichloroethylene Oxide
8.1.7.1  Repeated Skin Application and  Subcutaneous Administration:  Mice
8.1.7.1.1  Van Duuren et al. (1983).  Van Duuren et al. (1983) evaluated the
carcinogenicity of six epoxides of structurally related chloroalkenes by

                                      8-55

-------
                       TABLE 8-19.  SUMMARY OF NEGATIVE CARCINOGENIC ITY DATA FOR TRICHLOROETHYLENE
Species

Dogs


Rats
Guinea pigs
Monkeys
Rabbits

Cats

00
en
en Mice



Rats
Guinea pigs
Monkeys
Rabbits
Dogs

Rats


Number

16


12
11
2
4

8



28



15
15
3
3
2

37 newborn and 2
adults per group

Exposure3
Inhalation
150-750 ppm in air 20-48
hr/wk for 7-16 wk
Inhalation
3,000 ppm, 27 exposures, 36 days
100 ppm, 132 exposures, 85 days
200 ppm, 148 exposures, 212 days
200 ppm, 178 exposures, 248 days
Inhalation
200 ppm, 178 exposures, 248 days


Intragastric
0.1 mL in 40% oil solution
2/wk, unspecified treatment
period
Inhalation
3,825 mg/m3, 8 hr/day, 5 days/wk,
6 months
189 mg/m3 continuous for 90 days


Inhalation
2,000 ppm, 8 hr/day, 5 days/wk, _
10 weeks followed by 2 weeks
without treatment
Results Reference

No tumors; no deaths Seifter, 1944


Three rats died; Adams et al.,
no tumors 1951



No tumors; no deaths Monsinger and
Fiorentini, 1955


No tumors; no deaths Rudali, 1967



No tumors; no deaths Prendergast
et al., 1967




No tumorigenic potential Laib et al., 1979,
and Laib et al .,
1978b
Surviving animals were sacrificed on approximately the final  day of exposure or, in the studies by Laib et al.,
 observation; however, the observation period is not specified in the report by Rudali  (1967).
''Addition to summary by Waters et al. (1977).

SOURCE:  Waters et al., 1977.

-------
repeated skin application and subcutaneous injection in chronic studies with
female ICR/Ha mice.  The epoxides were cjj^-1-chl oropropene oxide (cis-CPO);
trans-1-chloropropene oxide (trans-CPO); cis-1,3-dichloropropene oxide (cis-
DCPO); trans-l,3-dichloropropene oxide (trans-DCPO), trichloroethylene oxide
(TCIO); and tetrachloroethylene oxide (PCEO).   The epoxides were prepared by
a modification of the method of Kline et al.  (1978)  to obtain pure samples.
However, cj_s_-DCPO and trans-DCPO contained 10  to 15% rn-dichlorobenzene during
the initial  14 months of the study, and TCIO  contained 10% dichloroacetyl
chloride early in the study.  Pure chemicals  were stored at -70°C.
     Mice were 6 to 8 weeks old when the study began.  Six animals were housed
per cage.  There were 30 mice in each experimental  group, except for 100 mice
in each untreated control group.  Maximum tolerated doses were estimated for
the chronic studies from doses not showing clinically apparent toxicity, inclu-
ding histopathologic examination of treatment  sites and liver, in preliminary
toxicity tests of 6 to 8 weeks duration.
     For the chronic skin application study,  test sites were shaved initially
and about every 2 weeks thereafter.  The epoxides were dissolved in 0.1 ml ace-
tone for application to the dorsal skin 3 times each week for the duration of
the study at the doses described in Table 8-20, except for PCEO, which, due  to
its instability in acetone, was applied undiluted to skin followed immediately
by application of acetone.  In the subcutaneous injection study, epoxides in
0.05 ml pure tricaprylin were administered at  the doses shown in Table 8-21
into the left flank once weekly for the duration of the study.
     Animals were observed daily and weighed  bimonthly.  Papillomas were cha-
racterized as skin lesions of about 1 mm in diameter which persisted more than
30 days.  Each animal was necropsied, and each lesion and tumor was histopatho-
logically examined.  Routine sections examined microscopically included skin,
                                     8-57

-------
                                      TABLE 8-20.  CHRONIC SKIN APPLICATION IN MICE
Thirty female  ICR/Ha Swiss mice per group were painted on the dorsal skin 3 times weekly with halo-epoxides at the doses
indicated in 0.1 mL acetone by micropipet except where noted. _
      Compound
Time (days) to
 fi rst tumor
  Median survival
time/test duration
      in days
No. of mice with
papilloma/total
   papillomas
No.  of mice with
  squamous cell
    carcinoma
P value3







00
en
OO





cis-CPO, 10 mg
trans-CPO, 10 mg
cis-DCPO, 10 mg
trans-DCPO, 10 mg
TCIO, 2.5 mg
PCEO, 7.5 mgf
m-Di chl orobenzene ,
3.5 mg
Dichloroacetyl chloride
250 yg
dl-Diepoxybutane, 3.3 mg
(positive control )
Acetone, 0.1 mL
No treatment (100
animals)
225
345
371
255

268




147




>432/432
513/574
522/574
511/573
526/575
>459/459
520/575

>576/576

462/573

>576/576
551/580

14/20b
15/20
16/19
20/27d
0
3/39
0

0

23/63

0
0

10
10C
10
176

1




21"




<0.0005
<0.0005
<0.0005
<0.0005

0.014




<0.0005




aP values based on number of animals with benign and malignant tumors:  X^ test with Yates'  correction was used.
bThree were keratoacanthomas.
C0ne animal developed a fibrosarcoma and a squamous cell  carcinoma.
dOne was a histiocytoma, and one was a keratoacanthoma.
eOne was a fibrosarcoma in the treatment area.
fpCEO was applied at 5 yL (7.5 mg) by Eppendorf pipet followed immediately by 0.1 mL acetone by micropipet.
90ne was a keratoacanthoma.
nTwo were fibrosarcomas in the treatment area.
SOURCE:  Van Duuren et al., 1983.

-------
                TABLE 8-21.  CHRONIC SUBCUTANEOUS INJECTION IN MICE
 Thirty female ICR/Ha Swiss mice per group were given injections once weekly of
 halo-epoxides s.c. in the left flank at the doses indicated in 0.05 ml tricaprylin
     Compound
                        Median survival
                       time/test duration
                            in days
  No. of animals
with local tumors
P value3
ci§_-CPO, 1 mg


trans-CPO. 1 mg

cjs_-DCPO, 500 yg
                             482/551


                             418/556

                             500/557
  1 fibrosarcoma            0.052
  1 carcinoma13
  1 leukemia
  4 fibrosarcomas           0.003
  1 carcinoma
  3 fibrosarcomas           0.003
  1 anaplastic sarcoma
  1 carcinoma and
    mammary tumor
trans-DCPO, 500 yg
TCIO, 500 yg
PCEO, 500 yg

m-Dichlorobenzene,
165 yg
Dichloroacetyl chloride,
50 yg
dJ^-Diepoxybutane, 330
yg (positive control)

Tricaprylin, 0.05 ml
No treatment
(100 animals)
498/557
547/558
491/558

525/560

>560/560

475/553


>524/561

551/580
5 fibrosarcomas0
1 fibrosarcoma
1 anaplastic sarcoma
1 leukemia
0

1 fibrosarcoma

9 fibrosarcomase
2 carcinomas
1 mammary tumor
0

1 fibrosarcoma
0.003
NS^
NS



NS

<0.0005





 aAll  mice with local  tumors included in calculation of p values:   X^ test with Yates'
  correction was used.  Each treated group was compared with the untreated control
  group.  C-CPO was marginally significant using the 0.050 level as cutoff point.
 bSquamous cell carcinomas.
 C0ne  with a focus of  angiosarcoma.
 dNS = not significant, and  p > 0.20.
 eOne  with a focus of  anaplastic sarcoma.

SOURCE:   Van Duuren et al.,  1983.
                                          8-59

-------
liver, stomach, and kidney in the skin application study,  and injection site
tissue and liver in the subcutaneous  injection study.
     As shown in Tables 8-20 and 8-21, vehicle and untreated control  groups
were concurrently on study.  Positive control  groups were  treated  with d^l-
diepoxybutane.  The above-mentioned impurities jn-dichlorobenzene and  dichloro-
acetylchloride were tested at the levels in which they were present in the
epoxides.
     Results described in Table 8-20  show no induction of  skin tumors from
repeated application with TCIO, and no significant (p < 0.05) increase in local
tumors resulted from subcutaneous injection of TCIO compared to controls (Table
8-21).  The positive control animals  responded with development of highly sig-
nificant (p < 0.0005, chi-square test) tumor increases in  both studies (Tables
8-20 and 8-21).  The investigators stated that no significant (p < 0.05) inci-
dences of malignant distant tumors were observed.
     As discussed by Van Duuren et al. (1983), chloroepoxides have been con-
sidered as activated carcinogenic metabolites of chloroalkenes.  The negative
results for carcinogenicity with TCIO in the study by Van  Duuren et al. (1983)
reflect the negative results for carcinogenicity with TCI  reported by Van Duuren
et al. (1979) using the same experimental design.  Presentation of data for each
epoxide tested by Van Duuren et al. (1983) is made herein  to provide evidence
that the relative carcinogenic activity of the six epoxides was not specifically
dependent on stereochemistry (cis or trans) or rate of hydrolysis (see Figure
8-14 herein).  The authors further noted that vinyl chloride epoxide has a
half-life in aqueous solution similar to that of TCIO, but that the former
compound has demonstrated carcinogenicity by subcutaneous  injection and initi-
ating  activity in an initiation-promotion study.  As indicated by the authors,
the epoxides tested would be direct-acting agents which would not have to rely

                                      8-60

-------
on metabolic activation at the sites of carcinogenic action in their study.
Under the conditions of the study by Van Duuren et al. (1983), TCIO was not
carcinogenic in female ICR/Ha mice.  Routine histopathologic examination
of additional tissues and organs might have given a broader evaluation of
pathology.

8.1.8  Cell Transformation Studies
8.1.8.1  Trichloroethylene
8.1.8.1.1  Price et al. (1978).  Price et al. (1978) evaluated the potential of
TCI to transform Fischer rat embryo cells (F 1706, subculture 108) in vitro.
The TCI sample, obtained from Fisher Scientific Co., was >^ 99% pure and would
have been stabilized with 20 ppm triethyl amine or 25 ppm diisopropylamine
(personal communication with Fisher Scientific Co.).  Cells were grown in
Eagle's minimum essential medium in Earle's salts supplemented with 10% fetal
bovine serum, 2 mM L-glutamine, 0.1 mM nonessential  amino acids, 100 U penicil-
lin, and 100 yg streptomycin per ml.  Liquid TCI was diluted 1:1000 in growth
medium before exposure of cells to desired TCI doses.
     Doses for the transformation assays were selected as those producing
minimal reduction in plating efficiency relative to medium-only control  cells
in preliminary tests.  In the preliminary tests, triplicate cultures were
exposed to 10-fold dilutions of TCI for 4 days before fixing and staining of
cells and counting of colonies.  Minimal reduction in plating efficiency (3 to
16%) compared to controls was produced over the range of TCI doses tested;
hence, the two highest doses (1.1 x 104 and 1.1 x 103 yM) of TCI were selected
for the transformation assay.
     In the transformation assay, quadruplicate cultures of F1706 cells  were
treated at 50% confluency with the two selected concentrations of TCI.  Cells
                                      8-61

-------
were exposed to TCI for 48 hours and then washed and refed with growth medium
alone.  Each culture was considered a separate series.   Additional  cultures
were exposed to the positive control agent 3-methylcholanthrene diluted in
acetone (1:1000) and growth medium to a dose of 0.37 pM.  Negative  control
cultures were grown in the presence of acetone (1:1000) in growth medium and in
growth medium alone.  At each subculture, one set of flasks was retained for 2
weeks without subdivision, and the other set of flasks  was subdivided 1:2 each
week to yield two new sets of cultures, one for the holding series  without
subdivision, and one for subdivision.  Cell transformation was characterized by
progressively growing foci composed of cells lacking contact inhibition and
orientation, and by the ability of foci to grow in semi-solid agar  four subcul-
tures after treatment.  The ability of transformed cells to induce  local fibro-
sarcomas after subcutaneous injection of 1 x 10^ cells  into newborn Fischer 344
rats was also evaluated.
     Transformation of cells exposed to TCI was observed, as well as the abil-
ity of transformed cells to grow in semi-solid agar and to induce local fibro-
sarcomas after subcutaneous injection into rats (Table  8-22).  Transformation
of cells exposed to the positive control agent 3-methylcholanthrene was
evident, as shown in Table 8-22.  Spontaneous cell transformation was not found
in negative control cells  (Table 8-22).
     Although transformation of F1706 cells in the presence of TCI  was observed
by Price et al.  (1978), the potency of TCI was low relative to that of the
positive control chemical, 3-methylcholanthrene.  The  response to TCI may
reflect an ability of this cell line to convert TCI to active metabolites, and
further evaluation of the  effect of TCI through active metabolites possibly
could have been  made by additional  testing with an exogenous metabolic acti-
vation system and by testing the ability of the F1706  cells to produce TCI

                                       8-62

-------
                  TABLE  8-22.   IN VITRO TRANSFORMATION OF FISCHER RAT EMBRYO CELL CULTURES BY TRICHLOROETHYLENE
00
Compound
Trichloroethylene

3-Methy 1 chol anth rene
Acetone
Medium
Dose
(uM)
1.1 x 10*
1.1 x 103
3.7 x 10'2
1:1,000

Morphological
transformation3
+5 (44)
+5 (44)
+4 (37)
(NA)
(NA)
Growth in
agar at P + 4&
9
8
124
0
0
Tumor incidence0
No. of animals with tumors/
No. of animals inoculated
13/13 (48)
12/12 (55)
12/12 (27)
0/13 (82)
NO
    aNumbers  represent  subcultures after treatment when transformed foci were first observed.  The acetone and
    medium control  cultures were still negative at the termination of the experiment nine subcultures after treatment.
    Cultures were  held for 2 weeks at each subculture and stained prior to screening.  Numbers in parentheses represent
    number of days  in  vitro prior to observation of morphological transformation.  NA = not applicable, as there was
    no  observed transformation by subculture 9 (81 days).

    bNumbers  represent  the average number of microscopic foci per three dishes.  Each dish was inoculated with 50,000
    cells and held  for 4 weeks at 37°C in a humidified C0£ incubator prior to staining with 2-(p-iodophenyl)-3-(p-
    nitrophenyl)-5-phenyl-2H-tetrazolium chloride.

    cAnimals  inoculated subcutaneously with cells that had been treated with chemical five subcultures earlier.
    Numbers  in parentheses represent days postinoculation to 100% tumor incidence.  ND = not done.

    SOURCE:  Adapted from Price et al., 1978.

-------
metabolites.  The authors stated that the F1706 cell  line,  at  the subculture
used (108), did not require addition of murine type "C" virus  and was  free of
any known infectious virus; however, these cells did  contain  genetic information
of the Rauscher leukemia virus.   It is not clear whether cell  transformation
was induced through action of the test chemicals on the host  cells,  their
integrated oncornavirus genome,  or both.  On the other hand,  the cell  transfor-
mation test is simply an in vitro screening system, designed  to provide at
best only suggestive evidence for carcinogenicity.  Additionally, cells were
exposed to liquid TCI diluted in growth medium; therefore,  since TCI is a
volatile substance, exposure of  cells to TCI as a vapor could have indicated
how different exposure conditions possibly could have influenced its ability  to
transform F1706 cells.

8.1.8.1.2  Styles (1979).  Styles (1979) reported an  investigation on  TCI in  a
cell transformation system with  BHK cells, using growth in  semisolid agar as  an
endpoint, as part of a larger study (Purchase et al., 1978) done to screen che-
micals for carcinogenic potential.  The BHK-agar transformation assay technique
used has been described by Styles (1977) and Purchase et al.  (1978).  In the
study reported by Styles  (1979), baby Syrian hamster kidney (BHK-211C1 13) cells
were exposed to five different doses of test substance in vitro in serum-free
liquid tissue culture medium in the presence of rat liver microsomal fraction
and cofactors (S-9 mix; Ames et  al., 1975).  The liver microsomal fraction was
obtained from Sprague-Dawley rats induced with Arochlor 1254.
     Cells were grown and maintained in Dulbecco's modification of Eagle's
medium in  an atmosphere  of 20% C02  in air.  Cells were maintained at 37°C until
confluent;  subsequently,  the cells were trypsinized and resuspended in fresh
growth medium.  Resuspended cells were  grown until 90% confluent for transfor-
                                      8-64

-------
mation assays or 100% confluent for stock.  Only cells with normal morphology
were used for assays.  To minimize spontaneous transformation frequency, cells
were obtained at low passage, grown to 90% confluency, and frozen in liquid
nitrogen.  Cells were thawed at 37°C in growth medium for further use.
     Test compounds were dissolved in DMSO or water as appropriate.  Each dose
was tested in replicate assays.  Cells incubated until 90% confluency were
trypsinized and resuspended in Medium 199 at a concentration of 10^ cells/ml.
Resuspended cells (106) were incubated with test chemical and S-9 mix
at 37°C for 4 hours.  To evaluate compound vapors, cells (2.5 x 10^) were
plated and incubated until 90% confluent, at which time growth medium was
replaced with medium 199 and S-9 mix.  Duplicate plates were exposed to test
chemical vapor at 37°C for 3 hours.  After treatment, cells were centrifuged
and resuspended in growth medium containing 0.3% agar.  Survival after treat-
ment was estimated by incubating 1,000 cells at 37°C for 6 to 8 days before
counting colonies.  Transformation was evaluated by counting colonies after
cells were plated and incubated for 21 days at 37°C.  The dose-response for
transformation was compared to that for survival.  Styles (1977) accepted a
fivefold increase in transformation frequency above control values at the LCso
as a positive result.  The spontaneous transformation frequency of BHK cells
(72 experiments) in this study was 50 +_ 16 per 106 survivors.  Suitability of
the soft agar medium for colony growth was checked by assays with polyomatrans-
formed BHK-21/C1 13 cells or Hela cells.
     Data reported by Styles (1979) on unstabilized TCI (source and purity not
reported) are compared with those found with vinyl chloride (VC) (ICI Mond
Division; purity not reported) by Styles (1977, 1979) in Figure 8-13.  Results
were negative with exposure to TCI solution in DMSO added to culture medium in
a dose range including levels in which toxicity was observed.  Vinyl  chloride
                                      8-65

-------
99-8




oo
o
c
m



OO
c-h
n>
00
w
1— 1
—4
««

1— >
UD

U3

































-"• -r\
3 -"•
03 C 3<
I -S O
^ w
1 o
INS 00 _
t— > 1 ^ 1
-~~. (— > 1 "
o oo 1 3
1— » • . I
l\
CO C""> — "2
(D ;
— '• — ' :
3 — • S
-•• ~i t
r+ Q) E

O 00 |
• — b
O
3
2
OJ
— i.
0
3
OJ
00
oo
OJ
oo
o
3
<
3
O
° i
— J» o
a* f
("D 5
o
^ 2
3 Vl
Q. O
.
Cr
-s

O
3-
O
-s
o
n>
3-
fD
3
ft)




O "i

>
5 < _ O — o
« u D X Q rsj -
aj ~ O n ^
« 3 2 m
1 Q.Q. n z
' S ~ z 3 o
lll 1 | °"
-3 o r
- n i z S
|Jj ll"""
' Si 7^ Z A ro
f — ' O W"

!

rsj
0 0








O "

O
C 0 (/) M
lif 8W
E ^ » o
i'i» 3 "-
*«• 5 3

- 2 i ±
g- | 3 O
s1? * s-
' s 3 ^.
5 =••* ~
i =•*- 3
S i
) *j
3-


NJ
Jl
8










•



i
1
1


1


1
|i
















A



£
~









TRANSFORMANTS
PER 10*
SURVIVORS
8Q O o Q O
Q O o O O
i i I I i i
1
|
1
N.
\.
x. !
^i^
, X«
V 	
V
\
• s
n

1
' •







TRANSFORMANTS
PER 10«
SURVIVORS
888883
1 i I i i T
1
O
114
n
n
n




1



• L
•








% SURVIVORS
tfi O
o o o
1 1
1
1
1 9
1 y
1 /
/
I /
r
A
/;


/
/
A ^








% SURVIVORS
o § 3







0
1

/
0
/
9^







-------
was effective as a gas in producing transformation and toxicity;  however,  VC
solution in DMSO added to culture medium was ineffective to suggest that cells
were exposed to doses too low to produce an effect by this approach.  Although
TCI doses high enough to produce toxicity did not induce transformation, expo-
sure of cells to TCI as a vapor could have provided a comparison  of the trans-
formation potential of TCI as a vapor and TCI in liquid solution, as was done
for VC.  Thus, the negative results reported in this study may,  in fact, be due
to its incompleteness.
     The study by Purchase et al. (1978), which was done on 120  chemicals  of
various classes, showed that the BHK-agar transformation assay system was
about 90% accurate in discriminating between compounds with demonstrated car-
cinogenic or noncarcinogenic activity, and was in approximately  83% agreement
with results from assays done by the authors with S. typhimurium (TA 1535, TA
1538, TA 98, TA 100).  Styles (1979) and Purchase et al. (1978)  indicated,
without presenting numerical data, that results were obtained in  the Salmonella
assays on VC in liquid solution, VC gas, and TCI in liquid solution that were
similar to results found in the transformation assays.  Purchase  et al. (1978)
also observed that metabolically activated agents transformed BHK cells more
strongly in the presence of S-9 mix, thus suggesting that BHK cells have limi-
ted intrinsic metabolic capability.
8.1.8.2  Trichloroethylene Oxide
8.1.8.2.1  DiPaolo and Doniger (1982).  DiPaolo and Doniger (1982) assessed the
capacity of putative epoxide metabolites of six chloroalkenes, including TCI,
to induce morphologic transformation in cultured Syrian hamster  cells.   The
tested epoxides were cis-1-chloropropene oxide (c-CPO); trans-1-chloropropene
oxide (t-CPO); cis-l,3-dich1oropropene oxide (c-DCPO); trans-1,3-dichloropropene
                                      8-67

-------
oxide (t-CPO), trichloroethylene oxide (TCIO); and tetrachloroethylene oxide
(PCEO).  TCIO and PCEO were prepared by auto-oxygenation of TCI and PCE with
removal of chloride by-products by extraction with 0.5N sodium hydroxide.   The
other epoxides were synthesized by oxidation of the parent compounds in m-chlo-
roperbenzoic acid.  Each epoxide was provided pure, as determined by nuclear
magnetic resonance analysis, by Drs. Kline and Van Duuren of New York Univer-
sity.
     Cells for culture were obtained from Syrian hamster embryos removed from
their mothers on the twelfth and thirteenth days of gestation.  Primary cul-
tures were derived by seeding 1 x 107 cells/dish from trypsinized embryos.
Secondary cultures were made 2 to 3 days  later by seeding 5 x 106 cells/10  ml
complete medium/plate.  Mass cultures were grown in modified Dulbecco's minimum
essential medium supplemented with 10% fetal bovine serum (FBS) under an atmos-
phere of 11% carbon dioxide.
     For each plate, 300 cells from secondary cultures were plated in complete
medium with 20% FBS along with 6 x 104 feeder cells previously irradiated as
confluent monolayer cultures.  After incubation for 2 days, the medium was
removed, the cells were washed with phosphate-buffered saline (PBS), and the
desired concentration of epoxide in PBS was added to the cells.  After incuba-
tion of the cells in the presence of epoxide for 30 minutes at 37°C, the PBS
was removed, the cells were washed, and fresh complete medium was added. After
further incubation for 5 days, the cells  were washed with PBS, fixed, and
stained with Giemsa.  Cell colonies ^ 2 mm in diameter were counted for cloning
efficiency and were examined for morphologic transformation, exhibited as criss-
crossing and piled-up cells not seen in controls.  Transformation frequency was
calculated relative to surviving colonies and initial number of cells plated.
Twelve dishes of cells were evaluated per dose.

                                      8-68

-------
     Results of the cell transformation assay with TCIO show a weak transfor-
mation effect, with high concentrations (>1 mM) capable of producing cytotox-
icity in the form of a dose-related reduction in cell survival (Table 8-23).
The weak effectiveness of TCIO at high concentrations may relate to its short
stability half-life of 1.3 minutes in aqueous solution (Figure 8-14).  However,
chemical stability in aqueous solution alone might not have been directly cor-
related with the ability of the six epoxides to transform cells in this study.
The most stable of the six epoxides in aqueous solution, c-DCPO and t-DCPO
(Figure 8-14), were the most potent inducers of cell transformation (Table
8-23), but the trans isomer was a stronger inducer than the cis.  Additionally,
although PCEO and c-CPO had similar stability half-lives in aqueous solutions
(Figure 8-14), c-CPO was the stronger inducer of cell transformation, which,
as suggested by the authors, may reflect a more limited ability of PCEO to
interact with DNA.  Although use of a positive control agent was not indicated
for this assay, the authors mentioned that transformed colonies found in this
study were similar to those induced by benzo[a]pyrene in this cell  line in
their laboratory.

8.1.9  Summary of Animal and Cell  Transformation Studies
     Increased incidences of malignant tumors in animals treated with TCI were
found in carcinogenicity studies on male and female B6C3F1 mice with gavage or
inhalation treatment, female Han:NMRI mice with inhalation treatment, and male
Fischer 344 rats with gavage treatment.  Thus, there are at least four studies
which indicate that TCI-exposed rodents show elevated cancer incidence.   The
NCI/NTP studies show an unequivocal  increase in mouse liver tumors.   The other
two studies just cited provide additional  support to the overall  body of evi-
dence, particularly since they were carried out in different animal  species or
                                      8-69

-------
          TABLE  8-23.   TRANSFORMATION AND  SURVIVAL OF SYRIAN HAMSTER EMBRYO
                    CELLS  TREATED WITH  DIVERSE  CHLOROALKENE OXIDESa
Compound Dose, mM
Acetone
c-CPO 0.11
0.27
0.55
1.1
t-CPO 0.22
0.55
1.1
2.2
c-DCPO 0.005
0.01
0.02
t-DCPO 0.01
0.025
0.05
TCEO 1.1
2.5
5.0
PCEO 0.87
4.3
8.7
17.1
Percent
survival
100 (29)b
110
94
90
50
96
78
54
12
92
93
81
89
85
44
91
86
73
98
96
100
70
Total
transformed
colonies
0
4
5
6
12
0
2
6
1
1
3
1
8
8
4
1
1
4
0
3
2
5
Average
transformations/
cells seeded, x 103

1.1
1.4
1.7
3.3
-
0.56
1.7
0.26
0.26
0.83
0.26
2.2
2.2
1.1
0.26
0.26
1.1
-
0.83
0.56
1.4
aPlastic dishes (60 mm)  were seeded with  300 cells  from a  secondary  culture with
 irradiated hamster cells (6 x 104) for a feeder,  grown for 2 days,  and  incubated  with
 the test chemical  in PBS in the absence  of growth  medium  for 30  min at  37°C.   Cells
 were washed, fed again, and incubated for 7 days  after which colonies were stained,
 counted for survival data,  and scored for morphologic  transformation.   Twelve
 dishes were used for each treatment.

^Number in parenthesis is the absolute cloning efficiency  of sol vent-treated
 cultures.

SOURCE:  DiPaolo and Doniger, 1982.
                                        8-70

-------
                            Compound
                                                                             Structure
                                                                                                  Half-Life (mm)
                           c/s-1 -Chloropropene oxide
                           
-------
strains.  Incidences of malignant tumors were not observed to be increased in
animals treated with TCI in carcinogenicity studies on female Fischer 344 rats,
male and female Osborne-Mendel  rats, male and female Sprague-Dawley rats, and
male and female ICR/Ha Swiss mice, all  with gavage treatment; male and female
Charles River rats, male and female Han:Wist rats, male and female Syrian
hamsters, and male HanrNMRI mice, all  with inhalation treatment; female ICR/Ha
Swiss mice with subcutaneous injection; and female ICR/Ha Swiss mice with
topical application.  Increased incidences of malignant tumors were not found
in carcinogenicity studies of TCI oxide in female ICR/Ha Swiss mice with subcu-
taneous injection or topical application.  Tumor-initiating activity was not
observed in studies where either TCI or TCI oxide was applied to the skin of
female ICR/Ha Swiss mice as the initiating agent.  Cell transformation activity
by TCI in cultured Fischer rat embryo cells and by TCI oxide in Syrian hamster
embryo cells was reported.  Cell transformation activity for TCI was not found
in cultured baby Syrian hamster kidney cells; however, when vinyl chloride was
tested under the same conditions as TCI, no transformation occurred with either
compound (see complete section 8.1 for details).

8.2  DESCRIPTION AND ANALYSIS OF EPIDEMIOLOGIC STUDIES
     There are currently six epidemiologic studies which relate to the issue of
TCI exposure and cancer.

8.2.1  Axelson et al.  (1978)
     Axelson et al.  (1978) found no excess mortality due to cancer in a cohort
study  of 518 Swedish male TCI production workers.  Workers with >_ 10 years of
observation were identified as a subcohort.  The authors did not indicate the
number of workers  in this subcohort, but did indicate that the subcohort inclu-
ded 3,643 of the 7,688 total person-years of observation in the study.  Since

                                      8-72

-------
the 1950s the TCI manufacturer had provided,  free of charge,  two urinary  ana-
lyses per year per worker of the TCI metabolite trichloroacetic acid (TCA).
Additional analyses were reported to have been provided at  low cost.  Using
this information, the subcohort with _>. 10 years of observation was divided into
low (average reading of TCA in the urine of less than 100 mg/L) and high  (ave-
rage reading of TCA in the urine of more than 100 mg/L).  A level  of 100  mg/L
of TCA in urine was reported to correspond roughly to an 8-hour time-weighted
average exposure of 30 ppm TCI in the air, which is the Swedish occupational
standard for TCI.  The expected number of deaths was calculated by multiplying
the person-years of observation by cause- and age-specific  national  death rates
from the respective calendar years and summing the fractional  age and calendar-
year contributions over the entire cohort.  Table 8-24 compares the expected
and observed numbers of deaths for the cohort and subcohort.
     There were no significant differences between the observed and the expec-
ted numbers of cancer deaths in the total cohort or in any  of the subcohorts.
The high exposure group, with >^ 10 years of observation, had  a risk ratio of
1.7 for cancer deaths, but this was not statistically significant.  Two deaths
from stomach cancer and two from leukemia occurred in the cohort;  the authors
did not indicate what the expected number of  deaths for these sites would be,
however.
     Although the data presented in this study do not demonstrate that TCI is
a human carcinogen, there are several limitations in the study design, or at
least in the manner in which the data are reported, that would restrict any
conclusions regarding this study.  For example, little information is available
concerning the duration of exposure of the workers.  No definition of minimum
exposure period seems to have been made by the authors, so  that workers who
were exposed for a very short period of time  could have been  included in  the
                                      8-73

-------
TABLE 8-24.  A COHORT STUDY  OF  MORTALITY  AMONG  MEN  EXPOSED  TO TRICHLOROETHYLENE
           IN THE 1950s AND  1960s  AND  FOLLOWED  THROUGH  DECEMBER  19753
Cohort
Total cohort

Subcohort
(_> 10
years of
latency
time)
- high exp.

- low exp.

Cause Person-yrs.
of at
death observation
All deaths 7688
Tumor deaths
All deaths 3643
Tumor deaths



All deaths 548
Tumor deaths
All deaths 3095
Tumor deaths
Observed
49
11
37
9



8
3
29
6
Expected
62.0
14.5
39.8
9.5



7.6
1.8
32.2
7.7
Risk
ratio
0.8
0.8
0.9
0.9



1.1
1.7
0.9
0.8
Confidence
95% limits
risk ratio
0.6 -
0.4 -
0.7 -
0.4 -



0.5 -
0.3 -
0.6 -
0.3 -
1.0
1.4
1.3
1.8



2.1
4.9
1.3
1.7
Exposure categorization is based on trichloroacetic acid (TCA)  in the urine,
 with low exposure referring to those not exceeding a concentration of 100 mg/L
 TCA in urine on average.
SOURCE: Adapted from Axel son et al., 1978.
study cohort.  Secondly, the definition of high and low exposure for the sub-
cohort was based on the average of TCA levels in the urine with no regard to
how long the individual may have worked in the high or low exposure area.
Thus, an individual may have had very low TCA levels over a short period of
time, despite the fact that most of his working time was in a high-exposure
area where few samples had been taken.  Thirdly, the number of workers in the
study (518) was relatively small for the detection of a significant risk of
cancer from TCI exposure.  Furthermore, it should be noted in this regard that
only half of the total person-years of observation were for the group with >^ 10
years of observation, indicating that most of the persons in the cohort had

                                      8-74

-------
less than 10 years of observation.   Thus,  less  than 259  people  had  been  fol-
lowed for as long as 10 years,  which itself may not be a long enough  observa-
tion period in which to observe an  increase in  cancer mortality,  considering
that a latency period for cancer may be as long as  20 years.  Finally, no
analysis was made with regard to specific  tumor sites.  If TCI  is a carcino-
gen, it is more likely to act on one particular organ or tissue site  than  on
all organ or tissue sites uniformly.  It is interesting  that 2  of the 11 deaths
from cancer in this cohort were from leukemia  (18%), whereas approximately 4%
of cancer deaths among males in Sweden are leukemia deaths.
     In summary, this study does not indicate that  TCI is a human carcinogen;
however, there are severe limitations in interpreting the data.  It should also
be noted that the primary author has reported in personal  communications (Axel-
son 1980 and 1983), that the study  is being updated through 1980  and  that  the
cohort will be expanded to include  1100 workers who were exposed  to TCI  between
1970 and 1975.

8.2.2  Tola et al. (1980)
     An ongoing cohort study on workers in Finland  exposed to TCI was reported
by Tola et al. (1980).  According to the authors, the cohort is to  be followed
and analyzed every fifth year.   The cohort consisted of  1,148 men and 969  women
known to have been occupationally exposed  to TCI at some time between 1963 and
1976.  Names of workers were obtained from the  biochemical  laboratory of the
Institute of Occupational. Health in Finland, where  most  of the  analyses  for
urinary trichloroacetic acid (TCA), a TCI  metabolite, were done as  part  of the
routine medical  examinations for workers.   Urinalyses were done by  the Fujiwara
reaction, which was reported to be  sensitive to levels as low as  1  mg TCA/L
urine.  The list of names of persons whose urinary  TCA levels had been measured
                                     8-75

-------
was supplemented with the names of  persons  who  had  been  reported to  have TCI
poisoning, according to the Occupational  Disease  Register  of  Finland, and with
the names of persons who had been reported  by employees  to have been exposed to
TCI.  After confirming worker identities, the vital  status and the causes of
death of those who died as of November 30,  1976 were obtained from the  Popula-
tion Data Register of the Central Statistical Office in  Finland, and cancer
death information was made available by the Finnish Cancer Registry.
     Urinary TCA levels for 2,004 workers were  reported.  The majority  (91%)
of the workers had TCA levels below 100 yg/L urine,  which, according to Axelson
et al. (1978), corresponds to an ambient exposure level  of 30 ppm TCI.  Further-
more, 78% of the workers had urinary TCA levels below 50 mg/L urine. An addi-
tional 80 workers registered with clinical  poisoning from  exposure to TCI were
also part of the cohort; however, urinary TCA measurements were not  available
for these workers.
     The observed numbers of all deaths and of  cancer deaths  were compared with
those expected based on national mortality  statistics for  1971.  A similar com-
parison was done for a subcohort of workers exposed before 1970.
     Results on the total cohort and subcohort  are summarized in Table  8-25.
No statistically significant (p < 0.05) differences between the observed and
the expected numbers of deaths from cancer  were found for  either the total
cohort or the subcohort.  Tumor sites for those dying from cancer  included the
gall bladder, lung, breast, uterus, testis, and multiple myeloma.  Two  workers
with more than 100 ug TCA/L urine and one worker with clinical  poisoning had
cancer; however, the difference of three cases  from the expected  number was  not
statistically significant  (p < 0.05).
     An excess cancer incidence among workers exposed to TCI  was  not found  in
this study.  The study, however, had a number of limitations.  The  duration  of

                                      8-76

-------
follow-up was short.  The age structure of the cohort was such that 60% of
the males and 40% of the females were less than 40 years old at the end of
follow-up.  The possibility exists, as indicated by the authors, that some
workers with low urinary TCA levels actually could have been exposed to another
degreasing solvent but were included due to the use of an improper exposure
test by the attending physician.  The actual durations of exposure for the
cohort were unknown.
 TABLE 8-25.  COMPARISON OF WORKERS EXPOSED TO TRICHLOROETHYLENE WITH THE TOTAL
    FINNISH POPULATION FOR MORTALITY FROM ALL CAUSES AND FOR CANCER MORTALITY
                                  Total  cohort
                              All  deaths                   Neoplasms
     Person-years        Expected      Observed       Expected     Observed
                                     Males
       7,041              52.1          40.0           6.6          5.0
                                    Females
       6,892              32.2          18.0           7.7          6.0
                    Subcohort (workers exposed before 1970)
                                     Males
       4,269              36.1          33.0           6.3          5.0
                                    Females
       4,822              25.8          13.0           6.0          4.0

SOURCE:  Adapted from Tola et al.,  1980.
                                     8-77

-------
8.2.3  Malek et al.  (1979)



     Malek et al.  (1979)  studied the incidence of  cancer  among  a  group  of  57



dry-cleaners in Prague,  Czechoslovakia,  who  used TCI  as a cleaning  solvent.



Exposure to TCI was  verified by analysis of  the urine for trichloroacetic  acid.



This group was reported  to  represent 86% of  all men  in Prague who had spent at



least one year working in this  type of  service since  the  1950s.   The follow-up



was from 5 to 50 years with 50% of  the  workers reported to have been followed



for more than 20 years.   Six cases  of cancer were  reported for  the  group,  which



was reported to be  not significantly (p  < 0.05) different from  what would  be



expected.  The expected  number  of cancer cases was not included in  the  authors'



report, however.   The primary weakness  of this study  is,  of course, the small



sample size, with  the resulting weakness in  the power of  the study  to detect a



significant increase in  cancer  incidence above that  expected,.





8.2.4  Novotna et  al. (1979)



     Novotna et al.  (1979)  identified 63 histologically confirmed cases of



liver cancer in the  city of Prague, Czechoslovakia,  for the years 1972  and 1974,



and reviewed their  employment histories  for  evidence  of exposure  to TCI.   The



study used no controls.   On the basis of employment  information dating  back to



January 1, 1957, 56  of the  cases could  be identified  as never having been  occu-



pationally exposed  to TCI.   For seven cases  there  was no  occuptional history.



The fact that none  of the liver cancer  cases had any  occupational exposure



to TCI is not particularly  surprising considering  that in the City  of Prague,



which has a population of approximately one  million,  there were,  at the time



of the study, only  544 workers  exposed  to TCI at 63  work  places.  The authors



indicated, however,  that at one time there had been  1,000 work  places where



there was TCI exposure,  but they did not indicate  when this was.
                                      8-78

-------
8.2.5  Harden  et al.  (1981)
     Hardell  et al.  (1981)  studied the chemical  exposure  histories  of  169 men
who had malignant lymphomas (Hodgkins  and  non-Hodgkins  disease)  and 338  con-
trols matched by age,  sex,  and place of residence.   The men  were aged  25 to 85,
and were identified  from the  records of the Department  of Oncology, University
Hospital, Umea, Sweden for  the period  1974-1978.  Controls werre extracted from
the National  Population Register.   The study was specifically  undertaken to
examine the association between malignant  lymphoma  and  exposure  to  phenoxy
acids or chlorophenols, but exposure to other chemicals,  including  TCI,  was
also evaluated.  Exposures  were ascertained by means of a self-administered
questionnaire which  included  questions on  employment, leisure  time  activities,
exposures to various chemicals, and drug and smoking habits.
     For the analysis  of solvent exposures, including TCI, the authors elimi-
nated the matching.   Seven  cases and three controls were  reported as having had
"high-grade" exposure  to TCI  (exposed  continuously  for  one week  or  more  or ex-
posed repeatedly to  brief exposures for at least one month).  When  compared
to 60 cases and 222  controls  with  no exposure to solvents, chlorophenols, or
phenoxy acids, the unadjusted odds ratio,  calculated by the  Carcinogen Assess-
ment Group, of being a case and having an  exposure  to TCI is 7.88 (95% CL,
2.30, 27.04).  Without adjusting for age,  this result is  somewhat question-
able, however.  Also,  the ascertainment of chemical  exposure by  self-admini-
stered questionnaire is of  doubtful value  because of the  varying recall  abil-
ities of individuals.   Also,  it is possible that the liver cancer cases  may
have been exposed to TCI prior to  January  1, 1957,  the  earliest  date for which
the authors had employment  data.  Fifteen  to 17 years is  not particularly long
when considering the latency  period for a  carcinogen.  Thus, the limitations
of this study preclude its  usefulness  in evaluating the carcinogenicity  of TCI,
                                     8-79

-------
although further investigation of the results  is  suggested.

8.2.6  Paddle (1983)
     Paddle (1983)  obtained from the Mersey Regional  Cancer  Registry  in  England
a list of all  the liver cancer cases for the period  1951-77.   Information  on  the
cases included name,  address at registration,  and dates  of birth  and  registration,
This list was then  reduced to those who had both  primary liver cancer and  an
address near Runcorn, which is the location of an Imperial Chemical  Industries
(ICI) plant that has  manufactured trichloroethylene  since 1909.   This resulting
list of 95 was then compared with a list of the "tens of thousands"  of employees
of the Runcorn plant  who had worked during the period 1934-76.  None  of  the 95
primary liver cancer  cases who had lived near  Runcorn were found  to  have worked
in the ICI plant.  As a further check, the files  in  the  company's computerized
medical records system were searched for records  of  past employees dying with
liver cancer as the underlying cause.  One liver  cancer  death  was found  for the
Runcorn plant and one liver cancer was found for  a neighboring plant. The
cancer registry records, however, showed that  the first  of these  had  suffered a
primary cancer of the esophagus with a secondary  liver tumor;  the other  patient
had had multiple secondary deposits from an unknown  primary  tumor.
     There are several deficiencies in this study.  In addition,  there are
limitations in the author's description which  make the study difficult to
evaluate.  One problem is that the author chose only those individuals from the
cancer registry who had addresses near Runcorn.  Obviously,  had a worker lived
farther away or moved from Runcorn following employment  at the ICI plant,  he
would not have been included in the study.  Another problem  is that  the  author
provides no indication as to the likelihood of finding an ICI  worker with  TCI
exposure among the 95 primary liver cancer cases.  Although  the author reported
                                     8-80

-------
that the list of 95 liver cancer cases was compared with a list of "tens of
thousands" of employees at the ICI plant at Runcorn, he also indicated in his
report that only about "1,000 or so" had TCI exposure.   No controls were used
in the study; thus there is not a good basis for a statistical  comparison.
Lastly, for those workers with TCI exposure, it is not  known how long they  had
been exposed, or how recent their exposures were.

8.3  RISK ESTIMATES FROM ANIMAL DATA
     The evidence for the carcinogenicity of TCI reviewed in this document  con-
sists primarily of positive mouse studies (NCI, 1976;  NTP, 1982; Henschler  et
al., 1980) and one marginally positive rat study (NTP,  1982).   These positive
animal studies reported an incidence of excess tumors  of several types.   Addi-
tional supporting evidence for the carcinogenicity of  TCI includes the fact
that TCI has been reported to be weakly mutagenic in numerous test systems  and
to have a low order of DNA binding.  A metabolic intermediate,  trichloroethy-
lene oxide, has been reported as positive in cell transformation studies; there
are no known differences among species with regard to metabolic pathways or
metabolic profiles.  The NCI and NTP studies, which show an unequivocal  increase
in liver tumors in B6C3F1 mice, serve as the basis for  cancer risk quantifica-
tion in this document.
     The evidence for the carcinogenicity of TCI has an uncertain classifica-
tion depending upon the differing scientific interpretations of the relevance
of mouse liver tumors (hepatocellular carcinomas) to human cancer risk.   How-
ever, the more conservative public health view would regard TCI as a probable
human carcinogen.  It is important to note that the quantitative estimations  of
the impact of TCI as a carcinogen are made independently of the overall  weight
of evidence for the carcinogenicity of TCI.  The calculations are made only
under the presumption that TCI is a human carcinogen.

                                      8-81

-------
8.3.1  Selection of Animal  Data  Sets
     For some chemicals, several  studies  in  different  animal  species,  strains,
and sexes, each run at several doses  and  different  routes  of  exposure,  may  be
available.  A choice can be made, when  appropriate,  as to  which  of  these  data
sets to use in the mathematical  extrapolation  model.   It may  also be appropri-
ate to correct for differences in metabolism between species  and for different
routes of administration.
     For TCI, the data from the  NTP (1982) and NCI  (1976)  oral gavage  lifetime
studies in male and female  mice,  showing  a significant dose-related excess  of
tumors (hepatocellular carcinomas), have  been  used.  Tumors were found  at both
the "high" dose (approximately the maximum tolerable dose) and the  "low"  dose
(one-half the maximum tolerable  dose).  Table  8-26  gives the  time-weighted
average dosage and tumor incidence in male and female  mice for the  NCI  and  NTP
studies.  TCI in corn oil  vehicle was given  once a  day by  gavage (vehicle only
was given to controls); in  the NTP study, 0.5  ml of  oil was used to dose  B6C3F1
mice by gastric intubation  5 days/week, starting at  8  weeks of age, continuing
for 103 weeks, and sacrificing at 112 and 115  weeks.   In the  NCI study, about
0.3 mL of corn oil was used, varying  with body weight, to  dose B6C3F1  mice  by
gavage 5 days/week, starting at  5 weeks of age, continuing for 78 weeks,  and
sacrificing after 90 weeks  (see  Section 8.2.1  for details  of  these  studies).
The time-weighted average gavage doses  for both studies are given in Table  8-26.
     A dose-response relationship for both sexes is  apparent  for each  of  the
two doses used in the NCI study.  It  should  be noted that  the chronic  daily
gavage doses used in the NTP and NCI  studies are in the range reported for
acute oral toxicity of TCI, LDjo to 1059, in other  strains of mice  (see Chap-
ter 5).  However, animal survival in  the  NCI lifetime  study  (1.5 year)  was  44%
for high-dosed males, 72% for low-dosed males, and  84% to  89% for females.
                                     8-82

-------
         TABLE 8-26.  INCIDENCE OF HEPATOCELLULAR CARCINOMAS IN MALE AND FEMALE MICE (B6C3F1)  IN THE NTP (1982)
                                              AND NCI (1976) GAVA6E STUDIES
oo
00
CO
Average Time-weighted
terminal average gavage dose3
Study
NTP



NCI





aSum (dose
Sex weight (g) mg/kg/day
M 40 0
1,000
F 35 0
1,000
M 33 0
1,169
2,339
F 26 0
869
1,739
in mg/kg x no. of days at that dose)
mg/animal/day
0
40

35
0
38.6
77.2

22.6
45.2

Incidence
8/48
30/50
2/48
13/49
1/20
26/50
31/48
0/20
4/50
11/47

(17%)
(60%)
(4%)
(27%)
(5%)
(52%)
(65%)
(0%)
(8%)
(23%)

    Sum (no. of days  receiving any dose)

-------
Similarly, in the 2-year NTP study, 32% of the males and 46% of the females
survived chronic administration of the single daily dose given, an amount which
approximated the low dose of the NCI study.  Presumably, TCI is subjected to
elimination processes during the 24-hour interval  between doses and on week-
ends, and hence the chronic LD may be much higher than the acute LD, parti-
cularly when the half-life is very short.  Prout et al. (1984)  estimated the
T 1/2 of TCI in B6C3F1 mice as only 1.75 hour, while Green and  Prout (1984)
demonstrated that chronic daily oral dosing of TCI (1,000 mg/kg/day) in mice
(180 days) produces a consistent daily amount of urinary metabolites,  indica-
ting that accumulation of TCI with chronic daily dosing does not occur.  Prout
et al. (1984) and Green and Prout (1984)  studied the metabolism and kinetics of
TCI in B6C3F1 mice given TCI by gavage in corn oil at doses of  10 to 2,000
mg/kg, i.e., including the conditions of the NTP and NCI bioassays.

8.3.2  Interspecies Dose Conversion
8.3.2.1  General Considerations—Carcinogenesis is a complex process,  not en-
tirely understood, and no observational  evidence exists to validate satisfac-
torily the extrapolated predictions of carcinogenesis from animal  models to
humans.  Thus, there is no scientific basis per se for  choosing one extrapola-
tion method over another in extrapolation of the carcinogenic response.  How-
ever, in extrapolating the dose-carcinogenic response relationships of labora-
tory animals to humans, the doses used in the bioassays must be adjusted in
some way to allow for such differences as size and metabolic rates. Therefore,
other biological information, particularly interspecies data, is examined for
TCI, since physiologic, biochemical, and  toxicologic responses  other than
cancer may provide information as to what factors  might be considered  and what
method is generally appropriate for the  extrapolation from laboratory  animals
                                      8-84

-------
to humans for this chemical.
     The major components requiring consideration in determining an appropriate
extrapolation base for scaling carcinogenicity data from laboratory animals to
humans are:  1) toxicologic data, 2) metabolism and kinetics, and 3) covalent
binding.  The biological basis for extrapolation of dose-carcinogenic response
relationships has been outlined and discussed previously by Davidson (1984) and
Parker and Davidson (1984).
8.3.2.1.1  Toxicologic Data.  The acute lethal concentrations or LDsgs for
inhalation and oral exposures to TCI are remarkably similar across species
(mice, rats, dogs, and humans), presumably due to an equally effective con-
centration leading to central nervous system depression, resulting in death
(Smith, 1966).  Klassen and Plaa (1966, 1967) and Tucker et al .  (1982) have
investigated other toxicologic end points (liver and kidney enzyme activity
changes and other modalities of liver and kidney damage), but these observa-
tions have not been extended across more than one to two species (see Chapter
5).  Hepatotoxicity of TCI has been shown to be directly related to the extent
of the metabolism of the parent compound (Buben and 0' Flaherty,  1985).

8.3.2.1.2  Metabolism and Kinetics.  The carcinogenic potential  of TCI is
generally considered to reside in cellular "reactive" intermediate metabolites
of TCI (e.g., TCI oxide, chloral, dichloroacetyl chloride, formyl  chloride)
rather than in the intact molecule itself.  These metabolites have been pos-
tulated to be responsible, for example, for the hepatotoxicity of  TCI (as
expressed by cellular morphological and functional  changes) and  its cellular
genotoxicity (as expressed by DNA binding), although the primary mechanism
(genotoxic or nongenotoxic) for TCI's carcinogenic potential  in  mice has not
been satisfactorily established.  The metabolism of TCI  has been relatively

                                      8-85

-------
extensively studied in three species:   mice,  rats,  and  humans.   Limited  phar-
macokinetic data are also available.   (See Chapter  4 for  a  detailed  review of
the metabolism and pharmacokinetics  of TCI.)   The pertinent metabolism and
pharmacokinetic information which  might be applied  to extrapolation  of TCI
dose-carcinogenic response relationships from laboratory  animals to  humans
is discussed in the sections that  follow.
     In mice, rats, and humans,  the  principal  end metabolites of TCI are common:
trichloroacetic acid (TCA), trichloroethanol  (TCE)  and  TCE-glucuronide,  and the
precursor metabolite, chloral.   In addition,  TCI  oxide  has  been  demonstrated to
be formed by both mouse and rat  hepatic microsomes  in vitro (Miller  and  Guenge-
rich, 1982).  Minor metabolites, including carbon dioxide,  dichloroacetic acid,
oxalic acid, and N-(hydroxyacetyl)-aminoethanol  (HAEE), have also been identi-
fied for each of these species  (see  Chapter 4).   Green  and  Prout (1984)  and
Dekant et al. (1984) made comparative  studies of  the biotransformation of TCI
in rats and mice and failed to  detect  any major  species or  strain difference in
metabolism.  Thus, there is persuasive experimental  evidence that the metabolic
pathways for TCI are qualitatively similar in mice, rats  and humans.
     The relative magnitude of  metabolism of TCI  across species, as  an index of
cellular damage from "reactive"  metabolites, provides information for determi-
ning a "mouse to man" extrapolation  scaling factor.
     Four types of studies have been done to investigate  dose-metabolism rela-
tionships:  1) single-dose studies across species,  using  14C-TCI and urinary
radioactivity and/or other methods as  measures of metabolism,  2) kinetic models
based on inhalation kinetics, 3) comparative balance studies between rat and
mouse with  l^C-TCI at one or two dose  levels, and 4) multiple  dose level stu-
dies in rats and/or mice.  Comparative balance studies, which  determine  the
metabolic fate and disposition of TCI  (amount metabolized,  amount excreted
                                     8-86

-------
unchanged, etc.), have been carried out after oral  dosing and after inhalation
exposures in both rats and mice.

     8.3.2.1.2.1  Oral Studies.   The balance studies  in which TCI  dosing was  by
the oral  route are the most pertinent to the conditions of the NTP and NCI  bio-
assays.  Prout et al.  (1984) administered 14C-TCI  in  corn oil  in the form of
single intragastric doses of 10,  500, 1,000, and 2,000 mg/kg to male Osborne-
Mendel and Wistar-derived rats,  and to male B6C3F1  and Swiss mice.  Exhaled
breath, urine, feces,  and carcass were analyzed for 14C-TCI radioactivity
(and unchanged TCI) for up to 72  hours after dosing,  and expressed as a per-
cent of dose administered.  The  results of these studies, as calculated and
re-expressed from the  data of Prout et al., are given in Table 8-27.  Regarding
these data, several general observations can be made  comparing the metabolism
of TCI in the rat and  mouse.
     As indicated in Table 8-27,  the recovery, from 91% to 98%, of 14C-radio-
activity in this experiment indicates that virtually  complete gastrointestinal
absorption of TCI occurred for all doses in both rats and mice. Peak blood
levels occurred at about 1 hour  in mice and at 3.5  hours in rats.   Some 80% to
90% of the TCI was excreted, either in the form of  metabolites or  unchanged in
exhaled breath, within the first  24 hours.
     There appears to  have been  no rat strain difference or mouse  strain dif-
ference in the disposition of TCI.
     For mice, oral doses of less than 1,000 mg/kg  were virtually  completely
metabolized.  Significant amounts of unchanged TCI  appeared in exhaled air  at
doses of 1,000 mg/kg (18%) and 2,000 mg/kg (14%),  indicating saturation of
metabolism beginning to occur at  doses approaching  1,000 mg/kg. For rats,
oral doses of less than 500 mg/kg were virtually completely metabolized, with
                                      8-87

-------
          TABLE 8-27.  DISPOSITION OF 14C-TCI 72 HOURS AFTER SINGLE ORAL DOSES TO MALE OSBORNE-MENDEL AND
                               WISTAR-DERIVED RATS AND TO MALE B6C3F1 AND SWISS MICE
Mice (mean of 4)

Dose
(mg/kg)
no
i
00
00 10
500
1000
2000
Dose/
animal3
(mg)



0.30
15.0
30.0
60.0

Metabolized
(mg equivalent) -



0.28
13.73
23.28
46.90

Exhaled
(mg equivalent)



0.01
0.90
5.25
8.16

Dose
(mg/kg)



10
500
1000
2000
Dose/
animal a
(mg)



2.0
100.0
200.0
400.0
Rats (mean

Metabolized
(mg equivalent)



1.93
55.40
79.00
80.40
of 4)

Exhaled
(mg equivalent)



0.03
42.70
112.40
311.20
aBased on experimental weight of animals:   rats,  average 200 g; mice, 30 g.

SOURCE:  Calculated from the data of Prout et al.,  1984.

-------
saturation occurring at about this dosage level.  These observations of satura-

tion of metabolism are in accord with those of Filser and Bolt (1979) and of

Andersen et al. (1980) who also observed saturation of metabolism in the rat

during inhalation exposure.

     For oral doses of 10 to 2,000 mg/kg (covering the range used in the NTP

and NCI mouse bioassays, Table 8-26), the ratio of the amount metabolized in

the rat versus the amount rfietabolized in the mouse was as follows:



                 Dose                     Ratio:  Rat/mouse
               (mg/kg)                 (mg metabolized/animal)

                  10                   (1.93/0.28)   = 6.89
                 500                   (55.40/13.73) = 4.03
                1000                   (79.00/23.28) = 3.39
                2000                   (80.4/46.90)  = 1.71


     At the highest dose, the ratio of the amount metabolized in the rat ver-

sus the mouse is smaller than it is at lower doses, probably due to saturation

of metabolism in the rat.  At the lowest dose, closer to 100% of the assimila-

ted dose is metabolized in both rats and mice.  These data indicate that, for

the dose range used in the cancer bioassays, the metabolism and, hence, the

fractional TCI doses metabolized for rats and mice are more consistent with

a V|2/3 base (surface area) than a V|l*0 base (mg/kg).

     Green and Prout (1984) orally dosed rats and mice subchronically for 180

days with TCI in corn oil at a level  of 1,000 mg/kg/day.  The relative propor-

tions of urinary metabolites (TCA and TCE-glucuronide) were unaffected by

strain differences.  The amount of the daily dose metabolized stayed constant

(Table 8-28), although the ratio of TCA to TCE-glucuronide increased.  These

observations indicate that 1) even at a high dosage level, TCI does not accu-

mulate significantly with daily dosing, i.e., the total  body half-life is suf-
                                     8-89

-------
                 TABLE  8-28.   METABOLISM OF TCI  IN B6C3F1 MICE:
                  EFFECT  OF CHRONIC DOSING (1,000 mg/kg/daya)
Day
1
10
180
1
10
180
Metabolized
(mg equivalent)
Chronic
18.27
22.50
15.75
Single-dose controls
18.27
19.35
16.86
Expired unchanged
(mg equivalent)
5.28
4.08
7.11
5.28
4.62
3.75
aBased on experimental  weight  of mice  averaging  30 g, the daily dose per mouse
 equals 30 mg in 0.5 corn  oil.

SOURCE:  Recalculated from the data  of Green  and Prout,  1984, based on an
         after-dosing collection period of  24 hours.
                                     8-90

-------
ficiently short (1 to 2 hours)  so that each daily dose is  cleared (metabolized,
exhaled, etc.) within the 24-hour dosage interval; and 2)  the increased ratio
o'f TCA to TCE-glucuronide suggests a possible adaptation of metabolic pathways,
which may occur simply because  TCA has a longer half-time  in the body than  TCE.
Prout et al. (1984) estimated the T 1/2 for TCI in mice at 1.75 hours;  there-
fore, 98% elimination occurs in 10.5 hours (T 1/2 x 6), and for some 14 hours
of each dose interval (24 hours), exposure to the highly reactive intermedi-
ates formed from TCI is practically nil.  For the rat or human with  longer
half-times of elimination, proportionally less time free from exposure to the
active chemical species is available (T 1/2 = 4 to 5 hours for humans).
     Dekant et al. (1984), using balance studies, compared the metabolism of
TCI in female rats (Wistar) and female mice (NMRI).  14C-TCI was administered as
a single dose by gavage in corn oil  at 200 mg/kg.  Radioactivity was determined
in urine, feces, carcass, and exhaled air for 72 hours post-administration.
Total recovery of radioactivity was 93% to 98%, and a virtually complete oral
absorption is indicated by the  results shown in Table 8-29.  For this single
dose, the rats excreted 52% of  the TCI dose through the lungs unchanged, while
the mice excreted 11% of the dose unchanged.  The ratio of the amount of TCI
metabolized by the rat versus that metabolized by the mouse was as follows:

                  Dose                            Ratio:  Rat/mouse
                 (mg/kg)                       (mg metabolized/animal)
                  200                           (23.04/4.56) = 5.05

The comparative surface area ratio of the two species, as  calculated from
their experimental body weights, is (240/24.4)°-67 = 4.46.  At this  dose,
the amounts of TCI metabolized  by the rat and the mouse are closely  propor-
tional to their relative surface areas.
                                     8-91

-------
         TABLE 8-29.  DISPOSITION OF 14C-TCI RADIOACTIVITY FOR 72 HOURS
           AFTER SINGLE ORAL DOSE (200 mg/kg) TO RATS AND MICE (NMRI)
                                     Mice (av.  of 3)

                                     Absolute dose
                                     5.1 mg/animal

                                     mg equivalent
                                       per animal
                    Rats (av. of 2)a

                    Absolute dose
                    48 mg/animal

                    mg equivalent
                      per animal
Expired TCI
        Unchanged

Metabolized
0.56 (11.0%)
24.96 (52.0%)
COo
Urine
Feces
Carcass
Washes

Total
0.31
3.89
0.25
0.10
0.01
4.56 (89.4%)
5.12
0.91
19.78
0.86
1.39
0.10
23.04
48.0





(48.0%)

aBased on experimental weight of animals:  female rats, 240 g; female mice,
 25.5 g.

SOURCE:  Recalculated from the data of Dekant et al., 1984.
                                      8-92

-------
     Buben and 0'Flaherty (1985)  recently examined the extent  of TCI  metabo-
lism in male mice during subchronic administration by gavage in  corn  oil  vehi-
cle given for 6 weeks (5 days/week; conditions similar to those  of the TCI  car-
cinogenicity bioassays).  While the strain was Swiss-Cox and not B6C3F1,  which
was used in the NTP and NCI bioassays, the results in this study are  probably
relevant and generally applicable.  A dose-amount metabolized  curve was esta-
blished with doses of 0, 100, 200, 400, 800, 1,600, 2,400, and 3,200  mg/kg/day.
Metabolism was measured by urinary metabolites excreted per day  (24-hour  urine
collections).  TCA (15% to 30%),  TCE, and TCE-glucuronide (70% to 85%), as
determined by gas chromatography, were found to be the principal  metabolites;
oxalic acid was not found to be a significant metabolite.  Metabolites in feces
represented less than 5% of the amount in urine.  No significant day-to-day
variability nor week-to-week trends were found in the urinary  metabolites
excreted.  Presumably, because of the relatively short half-life of TCI (1.75
hours), urinary metabolites did not accumulate over several  days with chronic
dosage.
     Figure 4-8 (in Chapter 4) shows the biphasic-appearing relationship  ob-
served between metabolism (amount of urinary metabolites) and  TCI dose.  The
initial portion of the curve, from 0 to 1,600 mg/kg TCI, is linear; the second
portion shows relatively abrupt plateauing, with an evident saturation of meta-
bolism.  In the linear portion of the curve, from 0 to 1,600 mg/kg dose,  27.5%
of the dose was converted into urinary metabolites; above 1,600  mg/kg, the
fraction of each dose metabolized decreased.  For comparison,  Green and Prout
(1984) orally dosed mice daily for 6 months with TCI in corn oil  at a level of
1,000 mg/kg/day, and found that 525-mg equivalents of TCI/kg body weight  were
metabolized per day (calculated from data in Table 8-28).  Buben and  0'Flaher-
ty (Figure 4-8) observed, for this dose of TCI, a metabolism of  318 mg urinary

                                      8-93

-------
metabolites/kg body weight.   The data  of Green  and  Prout  were  obtained  with
l^C-TCI and provide a measure of total  metabolism.   It  is likely  that Buben
and 0'Flaherty's measurements of the urinary  metabolites  underestimated the
total  metabolism, since the  measurements did  not  include  TCI metabolism to
carbon dioxide, metabolites  in feces and bound  to cellular macromolecules, and
other metabolites in urine in addition to TCA,  TCE, and TCE-glucuronide. The
molecular weight of TCI is 131.4, that of TCA is  163.4, and that  of TCE is
149.4.  If an adjustment is  made on  the basis of  1  ymol TCI being metabolized
to 1 pmol of TCA or TCE, and the mg  equivalent  represented by  urinary metabo-
lites is 272 mg/kg body weight.  In  comparison  to the findings of Green and
Prout, this represents an underestimation of  about  50%.  Such  a comparison is
in accord with the observations of Green and  Prout  that about  53% to 75% of  the
orally administered TCI at 1,000 mg/kg/day was  metabolized (Table 8-28). Buben
and 0'Flaherty, on the other hand, observed a TCI metabolism  fraction of 27.8%.
     An important observation can be made on  the  basis  of the  data of Buben  and
O'Flaherty (Figure 4-8) and Prout et al. (Table 8-27),  seen in relation to the
NTP and NCI bioassay data.  The NTP average gavage  assay  dose  (1,000 mg/kg/day)
is within the linear portion of the dose-amount metabolized curve of Buben and
O'Flaherty (Figure 4-8).  However, the saturation of metabolism is approached
in the "high" doses (1,739 and 2,339 mg/kg/day) of  the NCI bioassay and a lower
percentage of the administered dose is metabolized  at these doses.  Assuming
that Swiss-Cox mice accurately represent the metabolism characteristics of
B6C3F1 mice, the metabolism (or fractional dose metabolized)  contributing to
a carcinogenic response can be estimated from Figure 4-8  in the following
manner:
                                      8-94

-------
                                        Urinary
                                       metabolites             Percent  increase
           Gavage dose              from Figure  4-8             of  metabolites
           (mg/kg/day)          (mg metabolites/kg/day)            with dose

     NTP  Males:    1000                   318
          Females: 1000                   318

     NCI  Males:    1169                   372
                   2339                   625                        68

          Females:  869                   276
                   1739                   540                        96
Thus, with a twofold increase of TCI dosage to male mice of the NCI  study,

metabolism increased by only 68%.  This observation is consistent  with the

failure to observe a greater dose-related tumor increase in male mice in  the

NCI bioassay (Table 8-26).

     Buben and 0'Flaherty demonstrated that certain indices of TCI hepatotoxi-

city (increased liver weight and decreased liver glucose-6-phosphatase activity)

correlated well with the amount of TCI metabolized by mice.  The degree of

hepatotoxicity (as measured by each of these toxicity parameters)  when plotted

against urinary metabolite excretion was linear, suggesting that the hepatotox-

icity of TCI in mice is directly related to the metabolism of TCI.


     8.3.2.1.2.2  Inhalation Studies.  Stott et al. (1982) exposed rats and

mice to 10 and 600 ppm of ^C-TCI for 6 hours and estimated the amount metabo-

lized by collecting the radioactivity excreted in urine, feces, expired air,

etc., and by determining TCI in expired air for 50 hours post-exposure.  Tables

8-30 and 8-31 summarize their results.  The following general observations  can

be made on the basis of these data:

     For mice exposed to TCI for 6 hours at 10 ppm and 600 ppm, virtually all

of the total TCI uptake was metabolized.  At 600 ppm exposure for  6  hours,  all
                                     8-95

-------
    TABLE 8-30.  DISPOSITION OF   C-TCI 50 HOURS AFTER INHALATION EXPOSURE
                              OF MALE B6C3F1 MICE
10 ppm for 6 hr

TCI expired
Total body dose
Total metabolized
mg-eq/kg
0.08 (0.8%)
10.31
10.24 (99.2%)
mg/animala
0.003
0.36
0.36
600 ppm for
6 hr
mg-eq/kg nig/animal
10.01 (2.4%)
412.30
402.32 (97.6%)
0.35
14.43
14.0
aBased on 35 g mouse; calculated from data.

SOURCE:  Stott et al., 1982.
     TABLE 8-31.  DISPOSITION OF 14C-TCI 50 HOURS AFTER INHALATION EXPOSURE
                          OF MALE OSBORNE-MENDEL RATS


TCI expired
Total body dose
Total metabolized
10 ppm
mg-eq/kg
0.10 (2.1%)
4.70
4.60 (97.8%)
for 6 hr
mg/animala
0.03
1.18
1.15
600 ppm for 6 hr
mg-eq/kg mg/animal
29.83 (21.1%) 7.46
141.24 35.31
111.29 (78.9%) 27.82
aBased on 250 g rat; calculated from data.

SOURCE:  Stott et al., 1982.
                                     8-96

-------
of the assimilated dose metabolized was equivalent to that of an oral  dose to
mice of 500 mg/kg, and was approximately two- to fourfold less than the body
metabolic load in the NTP and NCI mouse bioassays.  For rats as well,  exposure
to 10 ppm resulted in the metabolism of virtually all of the TCI pulmonary
uptake; however, at 600 ppm some of the total assimilated dose was exhaled
unchanged (21%), indicating nonlinear kinetics and suggesting the approach of
saturation of metabolism at this exposure concentration.  At 600 ppm exposure
of rats for 6 hours, the total body metabolic load roughly approximated that
from an oral dose of 500 mg/kg when 42% of the dose was exhaled unchanged.
These observations of approaching saturation of metabolism in rats at  500 ppm
are in accord with those of Andersen et al. (1980), who estimated saturation
at about 1,000 ppm and a Vmax (maximum rate of the saturable process)  of 28
mg/kg/hour, and are considerably higher than the observations of Filser and
Bolt (1979), who estimated saturation at about 65 ppm but a comparable Vmax of
28 mg/kg/hour.  From the data of Stott et al. (1982), the Vmax can be  estimated
as 19 mg/kg/hour.
     The ratios of the amount of TCI metabolized by the rat versus the mouse at
10 and 600 ppm (Tables 8-30 and 8-31) are as follows:

                  Exposure
                ppm, 6 hours                 Ratio; Rat/mouse
                   10                      (1.15/0.36)  = 3.19
                  600                      (27.82/14.0) = 1.99

Therefore, on the basis of comparative surface area (250 g/35 g)0-67 = 3.7, the
ratio of the amounts metabolized by the rat versus the mouse approximates their
relative surface areas at the lower dose.  At the higher dose, the ratio of
the amount metabolized in the rat versus the mouse is smaller, probably due to
                                     8-97

-------
saturation of metabolism in the rat.   These inhalation  results  are  comparable
to those observed after oral  administration of TCI.

8.3.2.1.3  Metabolic Profile.   The experimental  evidence  is  consistent  with
the metabolic pathways for TCI  being  qualitatively  similar in mice,  rats,  and
humans.  In these species, the  following principal  end  metabolites  are  common:
trichloroacetic acid (TCA), trichloroethanol  (TCE)  and  TCE-glucuronide, and
their precursor metabolite, chloral.   In recent studies,  Green  and  Prout
(1984) and Dekant et al. (1984) made  comparative studies  of  biotransformation
of TCI in rats and mice, and failed to detect any major species or  strain
differences in metabolism.   In addition, TCI oxide has been demonstrated  to
be formed by both mouse and rat hepatic microsomes  in vitro  (Miller and Guen-
gerich, 1982).
     Dekant et al. (1984) isolated and identified,  by rigorous  methodology,  the
metabolites in exhaled air and  urine  of mice and rats given  an  oral  dose of  200
mg/kg of l^C-TCI.  The metabolite profiles for these two  species are shown in
Table 4-11 (Chapter 4).  Previously,  TCE, its glucuronide, and  TCA  had  been
identified as urinary metabolites of  TCI for mice,  rats,  and humans, and carbon
dioxide had been identified as  an end metabolite in the exhaled air of  mice
and rats (see Chapter 4).  Hathaway (1980) detected traces of dichloroacetic
acid in mouse urine, but only after very large (2 ml TCI/kg) oral doses.  To
explain these results, Hathaway proposed a "spillover"  model, in which  an
excess of compound saturates "deactivation mechanisms"  of the TCI oxide formed,
producing an intermediate that  becomes available for reactions  that include
spontaneous hydrolytic dechlorination to dichloroacetic acid (Kline and Van
Duuren, 1977).  However, Figure 4-8 shows that, at  the  dose  of  200  mg TCI/kg,
metabolism is far from saturated in mice (or rats).  Dekant  et  al.  (1984)
                                     8-98

-------
identified dichloroacetic acid in the urine of both mice and rats at this dose.
In addition, these investigators identified two previously unrecognized metabo-
Tites, oxalic acid and N-(hydroxyacetyl)-aminoethanol  (HAAE) in the urine of
both of these species.  They also identified significant amounts of HAAE in  the
urine of human volunteers exposed to 200 ppm TCI for 6 hours.   There are no
reported human studies in which carbon dioxide and oxalic acid, both normal
endogenous compounds, have as yet been identified as specific  metabolites of
TCI, since doing so would require administration of labeled TCI.  Nonetheless,
the observations of Dekant et al., and of others, indicate that qualitatively,
the pathways of TCI metabolism are similar in mice, rats, and  humans (see
Chapter 4 for a more detailed review).
     Whether there is a quantitative difference in TCI metabolism in mice,
rats, and humans is a question that remains unresolved.  Dekant et al.  (1984,
Table 4-11) have noted that, in the case of the principal  urinary metabolites,
the ratio of TCE to TCA is much greater  in the mouse than in the rat.   However,
they suggest that the smaller relative amounts of TCA in the urine of mice may
be due to an active TCA enterohepatic recirculation, with TCA  excreted  in the
feces in a conjugate form.  In humans, the urinary ratio of TCE to TCA  varies,
but is generally found to be about 2:1.

8.3.2.1.4  Covalent Binding.  Banerjee and Van Duuren  (1978) compared covalent
binding of ^C-TCI to mouse and rat microsomes and observed that mice micro-
somes are two- to fourfold more active in metabolizing TCI  to  reactive  cova-
lent-binding metabolites than are rat microsomes.   The covalent binding  is
proportional to the metabolism of TCI in these species.  Stott  et al. (1982)
also investigated covalent binding in vivo after 6-hour exposures of mice and
rats to 10 and 600 ppm l^c-TCI.  The time of  maximum binding post-exposure was
                                     8-99

-------
found to be 0 and 3 hours, respectively,  for liver and kidney.   These  results
are given in Table 8-32.   At low exposure concentrations  of TCI  (10 ppm),  rat
and mouse covalent binding per yg of protein occurred to  the same  extent,
with the amount of liver  binding being two-  to threefold  greater than  kidney
binding.  At high TCI exposures (600 ppm), which are metabolism-saturating
concentrations for the rat (Tables 8-30 and  8-31), the amount of covalent
binding was three- to fourfold greater per yg of protein  in mice than  in
rats, but was in proportion to TCI metabolism; therefore, the ratios of total
amounts of covalent binding in the rat versus the mouse approximates the  rela-
tive metabolic rates in these species.  Stott et al. also observed DNA aklyla-
tion in mice dosed with 1.2 g/kg high-specific l^C-TCI.  Their results showed
that TCI or its metabolites are capable of binding to DNA, but that the binding
ability was low when compared to dimethylnitrosamine and  aflatoxin. Similar
results have been reported by Parchman and Magee (1982),  and Di  Renzo  et  al.
(1982).

8.3.2.2  Calculation of Human Equivalent  Doses from Animal Data—The pharmaco-
kinetic and metabolic data that most closely approximate  the conditions of the
NTP and NCI carcinogenicity bioassays are found in the following studies:   1)
Prout et al. (1984), oral administration  of  single doses  of TCI at dose levels
up to 2,000 mg/kg to rats and mice; 2) Green and Prout (1984), chronic (180
days) oral administration of TCI at a dose level of 1,000 mg/kg/day to rats and
mice; 3) Dekant et al. (1984), oral administration of a single dose of TCI at  a
dose level of 200 mg/kg to rats and mice; and 4) Buben and O1Flaherty  (1985),
chronic (6-week) oral administration of TCI  at levels up to 2,000  mg/kg to
mice.  The comparative studies of TCI metabolism in both  mice and  rats of Prout
et al.  (1984), Green and Prout (1984), and  Dekant et al. (1984),  as well  as
                                     8-100

-------
      TABLE 8-32.  COVALENT BINDING TO LIVER AND KIDNEY MACROMOLECULES OF
              14C-TCI IN MALE B6C3F1 MICE AND OSBORNE-MENDEL RATS

Exposure
ppm Tissue
10 liver

kidney

600 liver

kidney


Species
mouse
rat
mouse
rat
mouse
rat
mouse
rat
Binding
pmole-eq
per ng
protein
Av. of 4
0.318
0.268
0.168
0.153
20.4
4.72
5.6
1.77
SOURCE:  Adapted from Stott et al., 1982.
                                     8-101

-------
the inhalation studies of Stott et al.  (1982),  provide evidence  that,  for  a
given dose, such as those used in the  NCI  and NTP bioassays,  the magnitude of
TCI metabolism in these species closely approximates  their relative  surface
areas.  For the risk calculation, it  is assumed that  the amount  of TCI  metabo-
lized per m2 surface area is equivalent among species.
     The carcinogenicity data of the  NTP and NCI bioassays show, in  male and
female B6C3F1 mice, a significant dose-related  excess incidence  of hepatocell-
ular carcinomas at time-weighted average dosage levels of 1,000  mg/kg  in the
NTP study, and "low" and "high" doses  for  female and  male mice in the  NCI  study
of 869 and 1,739 mg/kg, and 1,169 and  2,339 mg/kg,  respectively  (Table  8-26).
The data of Buben and 0'Flaherty (1985, Figure  4-8) clearly indicate that, in
Swiss-Cox mice, the metabolism of TCI  is linearly proportional to oral  doses
up to at least 1,600 mg/kg.  Hence, except for  the  "high" doses  (1,730  and
2,339 mg/kg) given to female and male  mice in the NCI study,  the amount of
metabolites formed and contributing to  a carcinogenic response is directly
proportional to the dose, which in turn, may be proportional  to  the  tumori-
genic response.  Therefore, the dose-metabolism curve of Buben and O'Flaherty
can be used in calculating TCI metabolism from  the  bioassay doses and  in cor-
recting for the saturation metabolism  associated with the "high" doses  of  the
NCI bioassay.  However, since the Buben and O'Flaherty dose-metabolism  curve
was based solely on urinary metabolites (and hence  underestimates the total
amount of metabolism), a more definitive measure of metabolism,  using  l^C-TCI
as in the studies of Prout et al. (1984) and Green  and Prout  (1984)  can be
used to estimate the total TCI metabolism associated  with the bioassay  doses.
Also, urinary metabolites are assumed  to be linearly  proportional to total
metabolism.
                                    8-102

-------
8.3.2.2.1  NTP Study.  In this study, the time-weighted average gavage dose
was 1,000 mg/kg/day, or, based on the average terminal  weight for the male
and female mice, 40 mg/male mouse/day and 35 mg/female  mouse/day (Table 8-26).
In other studies, with similar doses given either singly or chronically in
corn oil to B6C3F1 mice, virtually complete absorption  occurred; a portion of
the dose was excreted unchanged in exhaled air (Prout et al., 1984;  Green and
Prout, 1984; Dekant et al., 1984).  On the basis of these data, the  fractional
dose metabolized for oral doses of between 30 to 60 mg  per animal  averages 78%.
These doses are on the linear portion of the dose-metabolism curve for mice of
Buben and 0'Flaherty (1984, Figure 4-8).  Hence, the daily body metabolic load
(78% of the administered dose) for the mice in the NTP  study is 780  mg/kg.
     The relationship between the experimental administered dose (mg/animal)
and the amount metabolized is estimated on the basis of data from Prout et al.
(1984) (see Table 8-27), using the "Michaelis-Menten" type of equation, M = a x
d/(b + d), where d is the experimental administered dose (mg/animal) and M is
the corresponding amount metabolized.  The parameters a and b are determined  by
least-square estimates and are equal to 594.1 mg and 702.97 mg, respectively,
with a coefficient of determination r^ = 0.99.  This equation is used to con-
vert a bioassay dose in mice to the metabolized dose.  Assuming a lifetime of
104 weeks, the daily lifetime average effective exposure (LAE)  of the mice is
given by:

                 104 weeks     5 days
        LAEf/i  =  104 weeks  x  7 days  x metabolized dose, mg/animal/day

and the equivalent dose for humans,

                     = LAEf/| x scaling factor
                                     8-103

-------
The scaling factor for mouse to humans is:   (wH/wAf  where WH and  WA are,
respectively, the human and animal  body weights;

     male mouse (70/0.040)2/3 = 145.21; female mouse (70/0.035)2/3 =  158.74

     The lifetime average human equivalent  doses  for the NTP  mouse carcinogeni-
city bioassay of TCI, expressed in  units of mg/day,  mg/m2 surface area/day,  and
mg/kg/day, are presented in Table 8-33.
     However, a total daily dose administered as  a single bolus, such as in  a
bioassay, may not be biologically equivalent to the  same total  daily  dose
resulting from a continuous 24-hour exposure (e.g.,  air  pollutant) or from
several intermittent smaller doses  (e.g., drinking water). As  has been pointed
out by Beliles (1975), a 24-hour continuous exposure to  a given concentration
of an atmospheric pollutant is not  necessarily the simple equivalent  of one-
fourth the 8-hour workplace exposure.

8.3.2.2.2  NCI Study.  For this study, the time-weighted average gavage doses
("high" and "low" doses) in mg/kg/day  given to male  and  female B6C3F1 mice are
given in Table 8-26, and the corresponding dose per  single animal  per day is
given as calculated from the average terminal weight for the  male and female
mice of 33 g and 26 g, respectively.  As noted previously, doses in corn oil  to
B6C3F1 mice in this range have been observed experimentally,  when given as a
single oral dose or chronically (daily for 180 days), to be completely absorbed
from the GI tract, although a portion  of the dose effects a first-pass of the
liver and is excreted in exhaled air unchanged (Prout et al., 1984; Green and
Prout, 1984; Dekant et al., 1984).   The data of these investigators indicate
that the percentage of TCI dose metabolized for oral doses of between 30 and  60
mg per animal averages 78% (Table 8-27).  Furthermore, these  doses are on the
                                     8-104

-------
                        TABLE 8-33.  DOSE CONVERSION FOR THE NTP BIOASSAY (1982) IN B6C3F1 MICE
00
I



Oral
experimental dose
(time-weighted)
Sex

Male

Female
mg/kg/day
0
1,000
0
1,000
mg/day
0
40
0
35

Animal
metabolized dose
mg/day
0
31.98
0
28.17

Human


equivalent dose3
(lifetime
average exposure)
mg/day
0
3,317.23
0
3,194.06
mg/m2/dayb
0
1,793.10
0
1,726.52
mg/kg/day
0
47.39
0
45.62
Tumor
incidence
8/48
30/50
2/49
13/49
                                                                        human equivalent dose (mg/day) = animal
aAssuming the animals were exposed to TCI  for their entire lifetimes,
 metabolized dose (mg/day) x 5/7 x (WH/WA)2'3

 where     W/\ = 0.040 kg; terminal average weight  for male mice
                0.035 kg; terminal .average weight  for female mice

           WH = 70 kg; weight for standard man

The factor 5/7 reflects the fraction of days  per week that the animals  received treatment.

bSurface area for standard man = 1.85 m2 (Diem and Lentner, 1970).

-------
linear portion of the dose-metabolism curve for  mice  of  Buben  and  0'Flaherty


(1984).  However, the higher doses  of 1,739 and  2,339 are  on the nonlinear,


"saturating" portion of the curve and, as  estimated from this  curve  (Figure


4-8), further adjustments of 96% and  68%,  respectively,  are appropriate,


assuming the same curve for the B6C3F1 strain mouse.   The  daily amounts of dose


metabolized for the mice in the NCI study  are estimated  below  in the same


manner as was done for the NTP study  (see  Table  8-34), using the data from the


Prout et al. (1984) study, and by also including the  data  from Buben and 0'Fla-


herty, which affects the higher dose  only, for comparison. The data from the


Buben and 0'Flaherty study are especially  important to consider because this


study reflects a chronic dosing schedule,  making it more relevant  to the condi-


tions of the NTP and NCI carcinogenicity studies. The daily lifetime average


exposure (LAE) of the mice is given by



                78 weeks     5 days
       LAEM  =  go weeks  x  7 days  x amount metabolized, mg/animal/day




and the equivalent dose for humans,



          LAE^ = U\EM x scaling factor



The scaling factors for mouse to man  (WH/WA)2/3 are:



                      male mouse (70/0.033)2/3   = 165.09


                      female mouse (70/0.026)2/3 = 193.53



     The lifetime average human equivalent doses for  the NCI mouse carcinogeni-


city bioassay of TCI, expressed in units of mg/day,  mg2  surface  area/day,  and


mg/kg/day,  are presented  in Table 8-34.  The adjustment  for  partial  lifetime in


the NCI study, 90 weeks versus 104 weeks in the NTP  study, is  made later.
                                    8-106

-------
                      TABLE 8-34.  DOSE CONVERSION FOR THE NCI  BIOASSAY (1976)  IN B6C3F1  MICE
male
                   Oral
             experimental dose
              (time-weighted)
                           Animal
                      metabolized  dose
                                               Human
                                           equivalent dose3
                                           dose (lifetime
                                        average exposure)
Sex
mg/kg/day
mg/day
mg/day
mg/day
mg/m2/dayb
mg/kg/day
Tumor
incidence
    0
1,169
2,339
  0
38.58
77.19
  0
 30.90
 58.77
(39.96)c
    0
3,157.93
6,006.21
    0
1,706.98
3,246.60
 0
45.11
85.80
 1/20
26/50
31/48
CO
0
female



0
869
1,739


0
22.59
45.21


0
18.49
35.89
(34.45)c

0
2,215.19
4,299.79


0
1,197.40
2,324.21


0
31.65
61.43


0/20
4/50
11/47

aHuman equivalent dose (mg/day) = animal  metabolized dose (mg/day) x 5/7 x 78/90 x (WH/WA)2'3

 where      WA = 0.033 kg for male mice
               = 0.026 kg for female mice

            WH = 70 kg; weight for standard man.

 The factors 5/7 and 78/90 reflect the treatment schedule of 5 days per week for 78 weeks of a 90-week lifetime.
bSurface area for standard man = 1.85 m2 (Diem and Lentner, 1970).
cAssuming urinary metabolites account for about 50% of total metabolism, the additional information provided by
 the Buben and 0'Flaherty (1985) study would predict 40 mg/day (male) and 34.45 mg/day (females), a very small
 difference, as the amount metabolized at the high doses (this is comparable to the amounts predicted by fitting
 a curve to the Prout et al. [1984] data only, considering the differences in experimental conditions and mouse
 strains).

-------
8.3.2.3  Summary—The parent chemical, TCI,  is  metabolized  to  electrophilic
intermediate products that can react  readily with  cellular  macromolecules,
resulting in a spectrum of toxic responses  that includes  carcinogenesis.   For
example, the epoxide metabolite is  considered sufficiently  capable  of  produ-
cing the toxic end point of carcinogenicity.
     Thus, the experimental  data base for TCI provides  some support for deter-
mining an extrapolation base for laboratory  animals  to  man  of  W2/3, and allows
estimates of daily animal  exposure  (fractional  dose  metabolized)  contributing
to the tumorigenic response for the conditions  of  the NCI and  NTP bioassays.
     Estimates of equivalent human  exposure  can be calculated  using the data
from the experimental animal studies.  The  appropriate  interspecies scaling
is thus assumed to be on a surface  area basis,  i.e., W2/3.   it is of some
interest also to view the estimates of equivalent  human exposure  that  produced
hepatocellular carcinomas in mice in  perspective with human body  metabolic
loads from air exposure.  Humans appear to  metabolize inhaled  TCI extensively,
as reported by various investigators  (see Chapter  4).  The  studies  in  human
subjects indicate that at least 60% and possibly as  much  as 90% of  the retained
TCI from vapor exposures of less than 500 ppm (1,690 mg/m^) is metabolized in
the body.  Thus, human subjects undergoing  light physical activity, exposed
for 8 hours to 150 ppm TCI in air,  develop  a body  burden  of about 2,500 mg
(Monster et al., 1976, 1979; Nomiyama and Nomiyama,  1977; Fernandez et al.,
1977).  Nomiyama and Nomiyama (1977)  observed that,  at  least for  up to 300
ppm inhalation exposure, the metabolism of  TCI  is  not saturated in  humans.
(See Chapter 4 for a detailed discussion of these  studies).

8.3.3  Choice of Risk Model
8.3.3.1  General Considerations—The data used by the CAG for quantitative
                                    8-108

-------
estimation are of two types:   1)  lifetime animal  studies,  and 2)  human studies
in which excess cancer risk has been associated  with exposure to  the agent.   In
animal  studies it is assumed, unless evidence exists to the contrary, that  if
a carcinogenic response occurs at the dose levels used in  the study, then
responses will also occur at  all  lower doses, with an incidence determined  by
the extrapolation model.
     There is no solidly acceptable scientific basis for any mathematical
extrapolation model that relates  chemical carcinogen exposure to  cancer risks
at the extremely low concentrations that must be dealt with in evaluating envi-
ironmental hazards.  For practical  reasons, such low levels of risk cannot  be
measured directly either by animal  experiments or by epidemiologic studies.   We
must, therefore, depend on our current understanding of the mechanisms of car-
cinogenesis for guidance as to which risk model  to use.  At the present time
the dominant view of the carcinogenic process involves the concept that most
agents that cause cancer also cause irreversible damage to DNA.  This position
is reflected by the fact that a very large proportion of agents that cause
cancer are also mutagenic.  There is reason to expect that the quantal type of
biological response, -which is characteristic of  mutagenesis, is associated  with
a low-dose linearity and nonthreshold dose-response relationship.  Indeed,
there is substantial evidence from mutagenicity  studies with both ionizing
radiation and a wide variety  of chemicals that this type of dose-response
model is the appropriate one  to use.  At higher  doses there can be an upward
curvature, probably reflecting the effects of multistage processes on the
mutagenic response.  The low-dose linearity and  nonthreshold dose-response
relationship is also consistent with the relatively few epidemiologic studies
of cancer responses to specific agents that contain enough information to
make the evaluation possible  (e.g., radiation-induced leukemia, breast and
                                   8-109

-------
thyroid cancer, skin cancer induced  by arsenic  in  drinking water,  and  liver
cancer induced by aflatoxins in  the  diet).   There  is  also some  evidence  from
animal experiments that is consistent  with  the  linear nonthreshold model  (e.g.,
liver tumors induced in mice by  2-acetylaminofluorene in the  large scale  EDgi
study at the National  Center for lexicological  Research and the initiation
stage of the two-stage carcinogenesis  model  in  rat liver and  mouse skin).
     Because its scientific basis, although  limited,  is the best of any  of the
current mathematical extrapolation models,  the  nonthreshold model, which  is
linear at low doses, has been adopted  as  the primary  basis for  risk extrapola-
tion to low levels of the dose-response relationship. The risk estimates made
with this model should be regarded as  conservative,  representing a plausible
upper limit for the risk, i.e.,  the  true  risk is not  likely to  be  higher  than
the estimate, but it could be lower.
     The mathematical  formulation chosen  to  describe  the dose-risk relationship
at low doses is the linearized multistage model.   This model  employs enough
arbitrary constants to be able to fit  almost any monotonically  increasing dose-
risk response data, and it incorporates a procedure  for estimating the largest
possible linear slope (in the 95% confidence limit sense) at  low extrapolated
doses that is consistent with the data at all dose levels of  the experiment.
     The methods used by the CAG for quantitative  assessment  are consistently
conservative, i.e., tending toward high estimates  of  risk.  The most important
part of the methodology contributing to this conservatism is  the linear  non-
threshold extrapolation model.  There  are a  variety  of other  extrapolation
models that could be used, all of which would give lower risk estimates.  These
alternative models have not been used  in  the following analysis because,  in the
opinion of the CAG, the evidence of  carcinogenicity  for TCI does not warrant
that complete an analysis.  Furthermore,  with the  limited data  available from
                                    8-110

-------
these animal bioassays, especially at the high-dose levels required for test-
ing, almost nothing is known about the true shape of the dose-response curve at
low environmental levels.  The position is taken by the CAG that the risk esti-
mates obtained by use of the low-dose linear nonthreshold model  are plausible
upper limits, and that the true risk could be lower.
     In terms of the choice of animal bioassay as the basis for  extrapolation,
where more than one acceptable study is available, a general  approach is to use
the most sensitive responder, i.e., the data set that gives the  highest estimate
of lifetime cancer risk, on the assumption that humans are as sensitive as the
most sensitive animal  species tested.  For TCI, four sets of  bioassay data in
B6C3F1 mice, which produce comparable risk estimates, are used.
     Extrapolation from animals to humans can be done on the  basis of relative
body weights, surface areas, metabolic rates, or other measures.  The general
approach is to use the extrapolation base (mg/kg, surface area,  etc.) which
can be appropriately justified by the available experimental  data from animals
and humans.  However, it is usually not clear which extrapolation base is the
most appropriate for the carcinogenic response per se.  When  there is insuffi-
cient or no experimental data to determine an extrapolation base either direct-
ly or indirectly, the most generally employed and conservative method is used,
i.e., extrapolation from animal  dose to a human equivalent dose  on the basis  of
relative surface areas (W2/3).  For the TCI  gavage studies in mice, the use of
an extrapolation based on surface area (W2/3) rather than body weight (W1*0)
would increase the unit risk estimates by a  factor of 12 to 13.   However, for
TCI, experimental  data on metabolism and kinetics are used in dose extrapola-
tion from mouse to man.
                                     8-111

-------
8.3.3.2  Mathematical Description of the Low-Dose Extrapolation Model—Let P(d)
represent the lifetime risk (probability)  of cancer at dose d.   The multistage
model has the form

                P(d) = 1 - exp [-(q0 + q^ + q2d2 + ... + qkdk)]

where

                          qi  j> 0, i  = 0, 1, 2, ..., k

Equivalently,
                    . .            _  .          p            i,
                  Pt(d)  = 1 - exp [-(q^d + q2dc + ... + qkd^)]

where

                                           - P(0)
                                       1 -

is the extra risk over background rate at dose d  or the  effect  of  treatment.
     The point estimate of the coefficents qi, i  =  0,  1,  2,  ..., k,  and
consequently the extra risk function  Pt(d) at  any given  dose d,  is calculated
by maximizing the likelihood function  of the data.
     The point estimate and the 95% upper confidence limit of the  extra  risk,
Pt(d), are calculated by using the computer program GLOBAL83, developed  by
Howe (1983).  At low doses, upper 95%  confidence  limits  on the  extra risk and
lower 95% confidence limits on the dose producing a given risk  are determined
from a 95% upper confidence limit, q^, on parameter q^.   Whenever  q^ > 0, at
low doses the extra risk Pt(d) has approximately  the form Pt(d)  =  qi x d.
Therefore, q-i x d is a 95% upper confidence limit on the  extra  risk, and
is a 95% lower confidence limit on the dose,  producing  an  extra  risk  of  R.   Let
                                    8-112

-------
 *
ll »
LQ be the maximum value of the log-likelihood function.  The upper limit,
is then calculated by increasing qj to a value q1  such that when the log-
likelihood is remaximized subject to this fixed value, q^, for the linear
coefficient, the resulting maximum value of the log-likelihood LI satisfies the
equation

                         2 (LQ - LI) = 2.70554

where 2.70554 is the cumulative 90% point of the chi-square distribution  with
one degree of freedom, which corresponds to a 95% upper-limit (one-sided).
This approach of computing the upper confidence limit for the extra risk  Pt(d)
is an improvement on the Crump et al. (1977) model.   The upper confidence limit
for the extra risk calculated at low doses is always linear.  This is conceptu-
ally consistent with the linear nonthreshold concept discussed earlier.   The
slope, q^, is taken as an upper-bound of the potency of the chemical  in
inducing cancer at low doses.  (In the section calculating the risk estimates,
Pt(d) will be abbreviated as P.)
     In fitting the dose-risk response model, the number of terms in  the  poly-
nomial is chosen equal to (h-1), where h is the number of dose groups in  the
experiment including the control group.
     Whenever the multistage model  does not fit the  data sufficiently well,
data at the highest dose is  deleted, and the model is refitted to the rest of
the data.  This is continued until  an acceptable fit to the data  is obtained.
To determine whether or not  a fit is acceptable, the chi-square statistic
                        X2  .   y   <*,  -  N,P,
                              <•   NiPi  (1-Pi)
                             1 = 1
                                     8-113

-------
is calculated where N.,-  is the number of animals  in the i^   dose  group,  X^  is
the number of animals in the i™  dose group with a tumor response,  P^  is  the
probability of a response in the  ith dose group  estimated  by  fitting  the  multi-
stage model to the data, and h is the number of  remaining  groups.   The  fit is
determined to be unacceptable whenever X^ is larger than the  cumulative 99%
point of the chi-square distribution with f degrees of freedom,  where f equals
the number of dose groups minus the number of non-zero multistage  coefficients.

8.3.3.3  Adjustment for Less Than Lifespan Duration of Experiment—If the dura-
ation of experiment Le is less than the natural  lifespan of the  test  animal L,
the slope q^, or more generally the exponent g(d), is increased  by multiply-
ing a factor (L/Le) .  We assume  that if the average dose  d,  is  continued, the
age-specific rate of cancer will  continue to increase as a constant function  of
the background rate.  The age-specific rates for humans increase at least by
the 2nd power of the age and often by a considerably higher power, as demon-
strated by Doll (1971).  Thus, we would expect the cumulative tumor rate  to
increase by at least the 3rd power of age.  Using this fact,  we  assume that the
slope q^, or more generally the exponent g(d), would also  increase by at  least
the 3rd power of age.  As a result, if the slope q-^  [or g(d)] is  calculated  at
age Le, we would expect that if the experiment had been continued  for the full
lifespan L, at the given average  exposure, the slope q-±  [or  g(d)] would  have
been increased by at least (L/l_e)3.
     This adjustment is conceptually consistent  with the proportional hazard
model proposed by Cox (1972) and  the time-to-tumor model considered by Daffer et
al. (1980) where the probability  of cancer by age t and at dose d  is given by

                        P(d,t) =  1 - exp [-f(t)  x g(d)]
                                    8-114

-------
8.3.4  Calculation of Unit Risk
8.3.4.1  Definition of Unit Risk—This section deals with the unit risk for
TCI in air and water and the potency of TCI relative to other carcinogens
that the CAG has evaluated.  The unit risk estimate for an air or water pollu-
tant is defined as the increased lifetime cancer risk occurring in a hypothe-
tical population in which all individuals are exposed continuously from birth
throughout their lifetimes to a concentration of 1 ug/m^ of the agent in the
air they breathe, or to 1 pg/L in the water they drink.  This calculation is
done to estimate in quantitative terms the impact of the agent as a carcinogen.
Unit risk estimates are used for two purposes:  1) to compare the carcinogenic
potency of several agents with each other, and 2) to give a crude indication of
the population risk which might be associated with air or water exposure to
these agents, if the actual exposures are known.

8.3.4.2  Slope Calculation—Using the incidence data in Tables 8-33 and 8-34
and the corresponding human equivalent metabolized doses, the slope estimate,
q^, for TCI is calculated using the linearized multistage model, as shown in
Table 8-35.  The q^ value is used to compare relative potencies of carcino-
gens and to calculate the unit risks for drinking water and air.
     Since NCI mice were followed for only 90 weeks, rather than for 104 weeks
as in the NTP study, the risk estimates calculated on the basis of NCI  mice are
multiplied by a factor (104/90)3 = 1.54.  The rationale for such an adjustment
is given in Section 8.3.3..S.  The slope estimates in the last column of Table
8-35 are expressed in terms of the ambient concentration.  These values are
derived by using the calculated metabolized dose-response curve and the dose-
metabolism relationship  previously established.   Since both the metabolized
dose-response and dose-metabolism curves are linear at low doses,  the resultant
                                     8-115

-------
                     TABLE 8-35.   SLOPE ESTIMATES (qt)a
            BASED ON EXTRAPOLATION FROM DATA IN MALE  AND FEMALE  MICE
Data base
            K
    (animal)D
(mg metabolized
 dose/kg/day)-l
  (human)
(mg metabolized
 dose/kg/day)-1
     (human)
(mg administered
   dose/kg/day)-l
NTP:
    Male mice
    Female mice
   1.8 x ID'3
   7.5 x 10-4
  2.2 x lO-2
  9.5 x 10-3
   1.9 x 10-2
   8.0 x lO-3
NCI:

    Male mice
    Female mice

Geometric mean
   1.6 x ID'3
   5.0 x 10~4

   1.0 x ID'3
  2.1 x ID'2
  6.9 x 10-3

  1.3 x ID'2
   1.8 x 10-2
   5.8 x 10-3

   1.1 x 10-2
aq1 is defined as the 95% upper limit of the linear component (slope)  in the
 multistage model.  Since the dose-response curve is virtually linear  below
 1 nig/kg/day, the slope is numerically equal to the upper limit of the incre-
 mental lifetime risk estimates at 1 mg/kg/day.  The slope values are  used to
 compare relative carcinogenic potencies and to calculate the unit risks for
 drinking water and air.
bq1 (animal) is calculated using the daily lifetime average effective  expo-
 sure  (LAE^).
                                    8-116

-------
dose-response curve is also linear at  low  ambient  doses.  Because the slope
estimates from all  four data sets  are  comparable,  the  geometric mean of these
four numbers is used to represent  the  slope,  ql9 for TCI.  The qj values
calculated from the NCI and NTP studies  are presented  in the last two columns
of Table 8-35, expressed in terms  of metabolized dose  and also in terms of
ambient dose after adjusting for metabolism.   The  geometric mean is qj =
1.3 x 10~2/ (mg/kg/day) in terms of the metabolized dose, or q^ = 1.1 x 10~2/
(mg/kg/day) in terms of the ambient dose.  These q^ values are used to calcu-
late the unit risks for water and  air.

8.3.4.3  Risk Associated with 1 ug/L of  TCI in Drinking Water—The daily dose
(mg/kg/day) from consumption of 2  L of water  containing 1 pg/L of TCI is cal-
culated as follows:

                   d = (1 yg/L) x  (2 L/day) x (10-3 mg/yg)/70 kg
                     = 2.9 x ID'5  mg/kg/day

Here it is assumed that, on the average, humans weigh  70 kg and consume 2 L
of water each day.   Therefore, the incremental  lifetime risk associated with
       of TCI in drinking water is:
                          p = 1.1  x  10-2  x  2.9 x  lO-5
                            = 3.2  x  10-7
where 1.1 x 10-2 is the human carcinogenic  potency in terms of adjusted ambient
dose, taken from Table 8-35.

8.3.4.4  Risk Associated with 1 ug/m3  of  TCI in Air—Two animal inhalation
studies showed positive results in mice.  These were the Henschler et al .
(1980) study, showing increased malignant lymphomas in females, and the Bell
                                     8-117

-------
et al. (1978) study showing increased  hepatocellular carcinomas  in  males.   As
discussed in the qualitative section,  both studies  exhibited  serious  defects.
The Bell  et al.  (1978)  study, which  showed increased hepatocellular carcinomas
in male mice is  used to calculate the  unit risk  for air.   Because of  the defi-
ciencies of this study, the risk estimate is  used only  to  compare with  those
calculated on the basis of gavage studies.  In order to use the  potency q^
from the gavage  studies, it is necessary to estimate the amount  metabolized
when a subject is exposed to 1 yg/m3 of TCI in air.

8.3.4.4.1  Unit  Risk Estimate Calculated on the  Basis of Human Metabolized
Dose.  Although  there are many studies on metabolism of TCI in humans,  most of
the studies did  not present the relationship  between TCI uptake  and the amount
metabolized.  The study by Monster et  al. (1976)  is one which can be  used for
estimating the amount metabolized when a subject  is exposed to 1 yg/m3  of TCI
in air.  The study presents the total  uptake  and  the percentage  excreted as TCI
in the exhaled air for four subjects who were exposed to 70 ppm  and 140 ppm for
4 hours, including two half-hour 100-watt exercises. Since results from 70 ppm
and 140 ppm are  comparable, the results for the  70  ppm  group  are used for risk
calculation.  These results are as follows:
                                       Percentage
                                       of dose               Total  dose
   Subject        Uptake (mg)        exhaled as TCI           metabolized (mg)
A 500
B 450
C 470
D 660
9
17
10
8
455.0
373.5
423.0
607.2
                                     8-118

-------
The median amount metabolized for the four subjects is 439 mg.  Under the



assumption that the dose metabolized is linearly related to the intensity and



•duration of exposure, the dose metabolized corresponding to 1 yg/m3 of TCI in




air is





           d = 439 mg x  (24 hours/4 hours) / (70 ppm x 5,475 Pg/m3/ppm)



             = 6.9 x 10~3 mg





where  1 ppm = 5,475 yg/m3.  That is, under a continuous (24 hours) exposure to



1  yg/m3 of TCI in air, the body metabolic load is estimated to be





                    d =  6.9 x 10-3/70 = 9.9 x 10~5 mg/kg/day






Therefore, the unit risk for TCI in air is





                          P = 1.3 x ID'2 x 9.9 x lO-5



                            = 1.3 x ID'6





where  1.3  x 10~2  is the  human carcinogenic potency in terms of metabolized



dose,  taken from  Table 8-35.  This unit risk estimate is comparable to the



other  three estimates to be calculated by different approaches and/or on dif-



ferent data bases.  The  CAG recommends that P = 1.3 x 10'6 be used as the unit



 risk estimate for TCI in air because it is calculated on the basis of human



metabolized dose.





8.3.4.4.2   Unit Risk Estimate Calculated on the Basis of Animal Metabolized



Dose.   Stott et al.  (1982) exposed rats and mice to 10 ppm and 100 ppm of 14C-



TCI  for 6  hours and estimated the amount metabolized  by collecting the radio-



activity excreted in urine, feces, expired air, etc., and by determining TCI



in the expired air for 50 hours  post-exposure.  The results of 10 ppm expo-
                                      8-119

-------
sure, which is more relevant  to  the  risk  calculation, are reproduced below from



Tables 8-30 and 8-31.



                                     Mice                Rats





          TCI expired  (mg)           0.003               0.03



          Metabolized  (mg)           0.36  (99%)           1.15  (98%)



          Total uptake (mg)          0.36                 1.18








     These data can be used  in two different ways to  estimate  human risk due



to 1 yg/m3 of TCI in ai r.



     Approach 1:  Since the  body metabolic  loads for  mice after oral and inha-



lation exposure are known,  it is possible to estimate the unit risk by inhala-



tion using the slope q^ (animal) =  1.0 x  10~3/(mg/kg/day), calculated on the



basis of a gavage study (Table 8-35).  For  mice, 10 ppm  in air for 6 hours re-



sulted in a metabolized dose of  0.36 mg.  Under the assumption that the amount



metabolized at low doses is  linearly related to the intensity  and duration of



exposure, the daily dose metabolized corresponding to 1  yg/m3  continuous



exposure of TCI in air is:





            0.36 mg x (24 hours/6 hours)  /  (10 ppm x  5,475 Ug/m3/ppm)



                                = 2.63 x  ID'5  mg





where 1 ppm = 5,475 yg/m3.   That is, if a mouse  is exposed continuously to



1 yg/m3 of TCI in air, its  body  metabolic load is  2.63 x 10~5  mg/day or 7.5



x 10~4 mg/kg/day for a mouse weighing 0.035 kg.  Therefore,  the lifetime incre-



mental risk for mice due to exposure to 1 yg/m3  of TCI in air  is





                       P (mice)  = 1.0 x 10'3 x 7.5 x  10'4



                                = 7.5 x 10-7





                                     8-120

-------
To convert the risk estimate from mice to humans, the human concentration
C(yg/m3) that is equivalent to mouse exposure to 1 yg/m3 of TCI must be
determined.  Under the assumption that the amount metabolized per body surface
area is equivalent between mice and humans, the human concentration C(yg/m3)
must satisfy the relationship

                       C x 20 m3/day  = 1 x 0.043 m3/day
                         (70 kg)2/3        (0.035 kg)2/3
where 20 m-Vday and 0.043 m3/day are assumed to be, respectively, the volu-
metric breathing rates for a human weighing 70 kg and a mouse weighing 0.035
kg.  As a consequence, C = 0.34 pg/m3.  That is, the exposure concentration,
0.34 pg/m3, for humans is considered equivalent to that for mice at 1 y/m3.
Therefore, the unit risk estimate for humans is

                              P = 7.5 x 10-7/0.34
                                = 2.2 x ID'6

This value is comparable to that calculated on the basis of actual  human data.
     Approach 2:   In Approach 1, only mouse data are used.   As an alternative
approach, the relationship of the metabolized doses between mice and rats could
be used to infer the human metabolized dose.  As demonstrated previously, the
relative amount of dose metabolized between mice and rats is approximately
equal to two-thirds of the ratio of body weights.  Assuming that the same rela-
tionship holds for humans and mice, the human metabolized dose corresponding  to
1 pg/m3 of TCI in air would be

                  d = 2.63 x lO-5 mg/day x (70 kg/0.035 kg)2/3
                    = 4.17 x 10"3 mg/day or equivalently
                  d = 4.17 x 10-3/70 = 5.96 x 10"5 mg/kg/day
                                     8-121

-------
where 2.63 x 10"5 mg/day is the metabolized  dose  for  mice  corresponding  to

1 yg/m3 of TCI in air, as calculated  previously in  Approach  1.   Therefore,

the unit risk is estimated to be


                         P = 1.30 x  lO-2  x 5.96 x 10~5

                           = 7.7 x 10-7

       •k             Q
where q^ = 1.30 x 10  /(mg/kg/day) is the slope for humans taken from  Table

8-35.  This risk estimate is again comparable to  those  calculated previously.


8.3.4.4.3  Unit Risk Estimate Calculated  on  the Basis of an  Inhalation Study  in

Mice.  The Bell et al. (1978) study,  which showed increased  hepatocellular car-

cinomas in male mice,  can be used to  calculate the  unit risk for TCI in  air.

Because of the deficiencies in this  study, the risk calculated  from this study

should be considered crude and is to  be  used only for comparision with those

calculated on the basis of gavage studies.  For both  gavage  and inhalation

studies, the animal  species used in  the  experiment  and  the observed tumor site

and type are identical.  Animals in  the  Bell et al. (1978) study were  exposed

to TCI 6 hours/day,  5 days/week, for  24  months.   The  dose  and response data are

as follows:
Experimental
dose (ppm)
0
100
300
600
          LAD (ppm)a          0        17.86      53.57     107.14

          Hepatocellular
           carcinoma
           incidence         18/99      28/95      31/100     43/97
         aThe lifetime average dose (LAD) is calculated by
          LAD = dose (ppm) x (5 days/7 days) x 6 hours/24 hours.
                                     8-122

-------
Using the linearized multistage model,  the  carcinogenic  potency  (the  95%  upper
limit of the linear coefficient in  the  multistage  model)  for  animals  is calcu-
lated to be

                          qj  (animal) = 4.8 x  10~3/ppm
or
                          <\\  (animal) = 4.8 x  l(T3/5475
                                     = 8.8 x  10-7/(yg/m3)

using the fact that 1 ppm = 5475 yg/m3.
     To convert the potency from animals to humans,  it is assumed  that dose
relative to body surface area is equivalent among  species.  As calculated
previously (see Section 8.3.4.4.2), the exposure concentration C (yg/m3)
for humans that is equivalent to 1  pg/m3 for mice  is C =  0.34 yg/m3.
Therefore, the carcinogenic potency for humans is

                              q^ =  8.8  x 1(T7/0.34
                                 =  2.6  x 10-6/(yg/m3)

The risk due to 1 yg/m3 of TCI in air is thus  calculated  to be

                         P =  2.6 x  10-6/(yg/m3) x  1  yg/m3
                           =  2.6 x  ID'6

This risk estimate is also comparable to those calculated by  means of different
approaches and on the basis of different data  bases.

8.3.4.5  Discussion—To the extent  possible, the available metabolism and
kinetic information for TCI has been used in the risk calculations.   The  use  of
this information may or may not reduce  the  uncertainties  that are  associated
                                     8-123

-------
with the various steps involved in risk calculations.  There are three major
aspects of uncertainty in the risk assessment of TCI:

     1.  Extrapolation from high dose to low dose (i.e., low-dose
         extrapolation)
     2.  Extrapolation from animals to humans (i.e., species con-
         version), and
     3.  Extrapolation from gavage to inhalation (i.e., route-to-
         route extrapolation.

     In calculating the dose-response relationship for TCI,  the amount meta-
bolized is considered to be an effective dose.   The use of this surrogate
effective dose may not eliminate the uncertainty associated  with the low-dose
extrapolation because the dose actually reaching the target  sites may not be
linearly proportional to the total  amount metabolized, and the  shape of the
dose-response relationship is still  unknown.   However, it  seems reasonable to
expect that the uncertainty with regard to the  low-dose extrapolation would be
somewhat reduced by the use of the metabolized  dose because  the metabolized
dose better reflects the dose-response relationship, particularly within the
high-dose region.
     To extrapolate from animals to humans, the amount metabolized relative to
body surface area is assumed to be equivalent (i.e., equally potent)  among
species.  This assumption is by no means supported by the  empirical  data.  For
TCI, there is some evidence showing that, for a given dose in mg/kg  (by the
oral route) or ppm (by the inhalation route), the amounts  metabolized relative
to body surface area are approximately equal  among species.   However, there are
no observations which support the proposition that the metabolized dose relative
to body surface area is equally effective in  inducing tumors among different

                                     8-124

-------
 species.   An  alternative approach would be to assume that mg metabolized dose/
 kg/day  is  equivalent among species.   If this assumption is made, the slope
 factor  q1,  as calculated for TCI and  expressed in terms of (mg/kg/day)"1, would
 be  reduced  by approximately 10 times.
     The slope calculated from the gavage study, using the metabolized dose,
 provides a  better basis for extrapolating a risk estimate from the gavage to
 the  inhalation route.  This is so because the relationship between the ambient
 air  concentration and the amount metabolized for TCI is available for both
 humans  and  animals.  Therefore, the uncertainty due to route-to-route extrapo-
 lation  should be reduced by the use of metabolized dose.

 8.3.4.6  Interpretation of Unit Risk  Estimates—For several  reasons, the
 upper-limit unit risk estimate based  on animal bioassays is  only an approxi-
 mate indication of the real  risk in populations exposed to known carcinogen
 concentrations.  First, there may be  important differences in target site
 susceptibility, immunological  responses, hormone function., dietary factors,
 and disease.  In addition, human populations are variable with respect to
 genetic constitution and diet, living environment, activity  patterns, and other
 cultural factors.
     The unit risk estimate can give a rough indication of the relative carci-
 nogenic potency of a given agent compared  with other carcinogens.  The compara-
 tive potency of different agents is more reliable when  the comparison is based
 on studies in the same test  species,  strain, and sex,  and  by  the same route of
 exposure, although ordinarily, the risk should be independent  of route of expo-
 sure except in special  circumstances,  for  example,  nasal  or  lung carcinomas
with inhalation exposure, or  forestomach tumors  with gavage  administration.
     The quantitative aspect of carcinogen risk  assessment is  included  here
                                    8-125

-------
because it may be of use in the regulatory  decision-making  process,  e.g.,
setting regulatory priorities,  evaluating the  adequacy  of technology-based
controls, etc.  However, it should  be  recognized  that the estimation of  cancer
risks to humans at low levels of exposure is uncertain.  At best,  the linear
extrapolation model  used here provides a  rough but  plausible estimate of the
upper-limit of risk; i.e., it is not likely that  the true risk would be  much
more than the estimated risk, but it could  very well be considerably lower.
The risk estimates for TCI presented in this chapter should not  be regarded
as immutable representations of the true  cancer risks;  however,  the  estimates
presented may be factored into  regulatory decisions to  the  extent  that the
concept of upper risk limits is found  to  be useful. Table  8-35  gives the
upper-limit slope estimates, q^, from  the linearized multistage  model.
The slope estimate can be used  to compare the  relative  carcinogenic  potency of
TCI to that of other potential  human carcinogens.  The  slope is  also used to
calculate upper-bound incremental risks at  low levels of exposure.

8.4  RISK ESTIMATION FROM EPIDEMIOLOGIC DATA
8.4.1  Selection of Epidemiologic Data Sets
     Whenever possible, human data are used in preference to animal  bioassay
data.  If epidemiologic studies and sufficiently  valid  exposure  information are
available for a compound, they may be  used  in  several ways.  If  these data show
a carcinogenic effect, they are analyzed  to give  an estimate of  the linear de-
pendence of cancer rates on lifetime average  dose.   If  no carcinogenic effects
are seen in the epidemiologic studies  when  positive animal  evidence is availa-
ble, the assumption is made that a risk does  exist, but that it  is too small  to
be observable in the epidemiologic data.   An  upper limit to the  cancer incidence
is then calculated, assuming hypothetically that  the  true  incidence is below
                                     8-126

-------
the level  of detection in a human  cohort  of the  size  studied.
     Very  little information exists  that  can be  used  for  extrapolating  from
high-exposure occupational  studies to low environmental levels.   However,  if  a
number of  simplifying assumptions  are made, it is  possible  to  construct a  crude
dose-response model  whose parameters can  be estimated using vital  statistics,
epidemiologic studies, and  estimates of worker exposures.   Thus,  the  Axelson  et
al. (1978) study provides data that  may be used  to calculate 95%  upper-limit
estimates  on relative risk, although this study  does  not  provide  evidence  for
the carcinogenicity  of TCI.  Calculations may be based on the  subcohort data
from the Axelson et  al. study, with  particular focus  on the low-exposure group,
which had  a latency  period  of at  least 10 years.

8.4.2  Description of Risk  Model
     In human studies, the  response  is measured  in terms  of the relative risk
of the exposed cohort as compared  to the  control group.   The mathematical  model
employed assumes that for low exposures the lifetime  probability  of death  from
a specific cancer, PQ, may  be represented by the linear equation

                           P0 = A  +  BHx

where A is the lifetime probability  in the absence of the agent,  and  x  is  the
average lifetime exposure to environmental  levels  in  some units,  say  ppm.  The
factor, 6^, is the increased probability  of cancer associated  with each unit
increase of the agent in air.
     If we make the  assumption that  R, the relative risk  of lung  cancer for
exposed workers compared to the general population, is independent of the  length
or age of  exposure but depends only  upon  the average  lifetime  exposure, it
follows that
                                    8-127

-------
                             = P  = A + BH (xl + X2)
                               P0   A + BH X!
or
                             RP0 = A + BH (xi + x2)

where x^ = lifetime average daily exposure to the agent for the general  popu-
lation, X£ = lifetime average daily exposure to the  agent in the occupational
setting, and PQ = lifetime probability of dying of cancer with no or negligible
TCI exposure.
     Substituting PQ = A + BH x^ and rearranging gives

                               BH = P0 (R - D/x2
To use this model, estimates of R and X2 must be obtained from the epidemiologic
studies.  The value PQ is derived from the age-cause-specific death rates for
combined males found in the U.S. Vital Statistics tables using the life  table
methodology.

8.4.3  Calculation of Upper Limits of Risk
     Although the epidemiologic studies on TCI provide no evidence for its
carcinogenicity, they can still be used to provide information on an upper
limit of risk to be expected.  The CAG adopts this procedure of using negative
epidemiologic studies for providing upper limits when animal studies are
positive.  Such use of negative epidemiologic studies can modify estimates of
potency from the positive animal studies.
     Table 8-36 shows 95% upper-limit estimates on relative risk, as derived
from data in Table 8-24 obtained by Axelson et al. (1978).  These upper-limit
estimates are calculated according to the following  approach:

                                     8-128

-------
               TABLE 8-36.  NUMBER OF WORKERS WITH CANCER DEATHS
Observed
Cohort X
Total cohort 11
Subcohort with
10-year latency
(high and low
exposure) 9
High exposure 3
Low exposure 6
Upper limit Expected
u if no effect
19.7 14.5

17.2 9.5
8.7 1.8
13.3 7.7
Relative risk
95% upper limit
1.4

1.8
4.8
1.7
     The 95% upper limit = u/expected number of cancer deaths,  where  u  is  the
solution to the sum of Poisson probabilities, as follows:
                              X
                              I  e-" (u)1  -_  0.025
                             i=0    i!
One might also interpret u as the upper limit  of  the  number  of  workers  with
cancer deaths given the observed frequency,  and X = number of observed  workers
with cancer deaths.
     These upper-limit relative risks  can  be used to  provide estimates  of  the
the upper-limit probability of death from  cancer  due  to  a lifetime  exposure to
1 yg/rn^ of TCI.  This upper-limit probability  estimate is 1.7 x 10"^  per yg/m-*
of TCI in air, and it is approximately 10  times greater  than the risk estimate
based on the NCI (1976) and NTP (1982) carcinogenicity gavage bioassays.   In
showing these risk calculations below, the CA6 emphasizes that:
                                    8-129

-------
     1.  This upper-limit value is  derived  from  a  negative epidemiologic study.
         Not only was there no  specific  target organ  for cancer, but the total
         number of cancer deaths was  less than expected.
     2.  Since there was  no specific  target organ,  the  upper-limit  lifetime
         probability estimate is based on all cancer  deaths.   Since the
         lifetime probability of death from cancer  to humans  is 0.19 (based on
         1977 U.S. death  rates, race  and sex combined), any even moderately
         conservative confidence limit will  provide a large upper-limit risk.
     The calculations are based on  the subcohort data from the Axelson et  al.
(1978) study, specifically the  low  exposure group with  at least a 10-year
latency period.  Table 8-36 projects  an  upper-limit relative  risk of 1.7.
Since "low exposure" referred to exposures  not exceeding concentrations of 100
mg TCA per ml in urine on the average, which, according to Axelson  et al.
(1978) corresponds to an  ambient air  exposure of no more than 30 ppm TCI in
air, 30 ppm is used as the 8-hour time-weighted  average dose. This corresponds
to 30 ppm x 5,475 (yg/m3)/ppm - 1.64  x 105  yg/m3.   A  lower dose estimate
would provide an even higher upper-limit probability  of death.
     In the category of years exposed, in the absence of any  other  information,
10 years was chosen as a  point  assumed to be somewhere  in the middle of the
range.
     On the basis of the  above  information, the  total lifetime dose of TCI to
the low-dose exposure workers is estimated  as:

           (0 ppm + 1.64 x 10s) x 240 days  x 10  years = 7.7 x 103 yg/m3
                    2             365 days    70  years

The 1.64 x 105 yg/m3 TCI  of the working  day is  averaged with  the 0  yg/m3 of
the non-working day, assuming that  half  of  the  air breathed during  a day will
                                    8-130

-------
be during working hours.  The value 7.7 x 10~3 yg/m3 is assumed to be a con-
tinuous exposure concentration.  Thus, the upper-limit of the probability of
death from 1 yg/m3 TCI in air (continuous exposure) is estimated as:
             p =  Q    -    j = 0.19 (1.7- 1)1 , 1-7 x 1Q-5       /m3
                      d             7.7 x 103
where R = the relative risk, dj = 1 yg/m3, d2 = dose of workers, and PQ = the
background lifetime probability of death from all cancers.
     Based on a comparison of this upper-limit estimate with that from the
animal studies, the Axelson et al . study is seen to provide no evidence that
the carcinogenicity potency of TCI is less for humans than for animals.

8.5  RELATIVE CARCINOGENIC POTENCY
8.5.1  Derivation
     One of the uses of the concept of unit risk is to compare the relative
potencies of carcinogens.  To estimate relative potency on a per mole basis,
the unit risk slope factor is multiplied by the molecular weight, and the
resulting number is expressed in terms of (mmol/kg/day)"1.  This is called
the relative potency index.
     Figure 8-15 is a histogram representing the frequency distribution of
potency indices of 54 chemicals evaluated by the CAG as suspect carcinogens.
The actual  data summarized by the histogram are presented in Table 8-37.
Where human data are available for a compound, they have been used to calculate
the index.   Where no human data are available, animal  oral  studies and animal
inhalation  studies have been used in that order.   Animal  oral  studies are
selected over animal  inhalation studies  because most of the chemicals have
animal  oral  studies;  this allows potency comparisons by route.
                                   8-131

-------
oo
 i
CO
ro
            OX) -T|
            0>  O -••
            -S  rt-iQ
            O  0> C
            -.. 3 -S
                O ft)
            O
            03
            n>
        OO
        I
                Q. tn
            in  o
            i/>  O)
            fD  
                -h rt-
                   O
 c-t- -P» -S
        cu
 CD 01  3
 -s  c:
 o  (/»  -j
 c  -a  n>
-o  (i> 13
 •   o  -j
    «-*•  
        to
    o  a>
    o>  3
    -S  ft-
    o  -•-
    -••  3
    3 U3
    O
    U3  -«,
    (T)  -S
    3  CD
    l/l _Q
        C
    (T)  (D
    <  3
    QJ  O
               cu
               rt-
               n>
               Q-
               *<  cr
                   c
               <-•• rt-
               3- -"•
               ro  o
                   n>
                                                       N>
                                                                          FREQUENCY
                                                                                          _L         —i
                                                                  O>        00          O         to
O)
oo
N>
o
                                                  I     I      I     I     I     I     I     I     I     I     I     I     I      I     I     I     I     |    |     I
                                        x

                                        O
                                                                                                                                                                   O
                                                                                                                                                                     +
                                                                                                                                                                     M
                                                                                                                                                                               X


                                                                                                                                                                               O
                                                  I    '     I     '     I     I     '      I     '     I     I     '     I     '     I     '     '     I     '    '
                                                                                                                                                                                        D

                                                                                                                                                                                        >
                                                                                                                                                                                        30
                                                 m
                                                                                                                                                                                         D
                                                                                                                                                                                         c:
                                                                                                                                                                                         >  CO
                                                                                                                                                                                         3)  3
                                                                                                                                                                                         m
                                                                                                                                                                             D

                                                                                                                                                                             >  to
                                                                                                                                                                             3)  3
                                                                                                                                                                             -I  a-
                                                                                                                                                                                        D
                                                                                                                                                                                        m

-------
          TABLE 8-37.   RELATIVE CARCINOGENIC POTENCIES AMONG 54 CHEMICALS EVALUATED  BY THE CARCINOGEN  ASSESSMENT  GROUP
                                                  AS SUSPECT HUMAN  CARCINOGENS
oo
CO
Level
of evidence3
Compounds
Acrylonitrile
Aflatoxin Bj
Aldrin
Ally! chloride
Arsenic
B[a]P
Benzene
Benzidene
Beryl 1 i urn
1,3-Butadiene
Cadmi urn
Carbon tetrachloride
Chlordane
CAS Number Humans '
107-13-1
1162-65-8
309-00-2
107-05-1
7440-38-2
50-32-8
71-43-2
92-87-5
7440-41-7
106-99-0
7440-43-9
56-23-5
57-74-9
L
L
I

S
I
S
S
L
I
L
I
I
Animal s
S
S
L

I
S
S
S
S
s
s
s
L
Grouping
based on
IARC
criteria
2A
2A
2B

1
2B
1
1
2A
28
2A
2B
3
SI opeb
(mg/kg/day)'1
0.24(W)
2900
11.4
1.19xlO-2
15(H)
11.5
2.9xlO-2(W)
234(W)
2.6
1.0xlO~1(I)
6.1(W)
1.30X10-1
1.61
	 7
Molecular
weight
53.1
312.3
369.4
76.5
149.8
252.3
78
184.2
9
54.1
112.4
153.8
409.8
Potency
index0
1x10+1
9xlO+5
4x1 0+3
9x10"!
2xlO+3
3xlO+3
2x10°
4xlO+4
2xlO+1
5x10°
7xlO+2
2xlO+1
7x10+2
Order of
magnitude
(Iog10
index)
+1
+6
+4
0
+3
+3
0
+5
+1
+1
+3
+1
+3

-------
                                                       TABLE 8-37.   (continued)
CD

I—*
CA>
Level
of evidence3
Compounds
Chlorinated ethanes
1,2-Dichloroethane
hexachloroethane
CAS Number

107-06-2
67-72-1
1,1,2,2-Tetrachloroethane 79-34-5
1,1,2-Trichloroethane
Chloroform
Chromium VI
DDT
Dichlorobenzidine
1,1-Dichloroethylene
(Vinyl idene chloride)
Dichloromethane
(Methylene chloride)
Dieldrin
2,4-Dinitrotoluene
Diphenylhydrazine
Epichlorohydrin
Bis(2-chloroethyl )ether
79-00-5
67-66-3
7440-47-3
50-29-3
91-94-1
75-35-4

75-09-2

60-57-1
121-14-2
122-66-7
106-89-8
111-44-4
Humans

I
I
I
I
I
S
I
I
I

I

I
I
I
I
I
An imal s

S
L
L
L
S
S
S
S
L

L

S
S
S
s
s
Grouping
based on
IARC
criteria

2B
3
3
3
2B
1
28
2B
3

3

2B
2B
28
2B
28
SI opeb
(mg/kg/day)'1

9.1x10-2
1.42x10-2
0.20
5.73x10-2
7x10-2
41(W)
0.34
1.69
1.16(1)

6.3x10-4(1)

30.4
0.31
0.77
9.9x10-3
1.14
Molecul ar
weight

98.9
236.7
167.9
133.4
119.4
100
354.5
253.1
97

84.9

380.9
182
180
92.5
143
Potency
index0

9x10°
3x10°
3x10+1
8x10°
8x10°
4x10+3
1x10+2
4x10+2
1x10+2

5x10-2

1x10+4
6xlO+1
1x10+2
9x10-!
2x10+2
Order of
magnitude
(]og10
index)

+1
0
+1
+ 1
+1
+4
+2
v +3
+2

-1

+4
+2
+2
0
+2

-------
                                                     TABLE 8-37.   (continued)
CO
co
en
Level
of evidence3
Compounds
Bis(chloromethyl ) ether
Ethyl ene dibromide (EDB)
Ethyl ene oxide
Heptachlor
Hexachlorobenzene
Hexachl orobutadi ene
Hexachlorocyclohexane
technical grade
alpha isomer
beta isomer
gamma isomer
Hexachl orodi benzodi oxi n
Nickel
Nitros amines
Dimethyl nitrosamine
Di ethyl nitrosamine
Dibutylnitrosamine
N-nitrosopyrrol idine
N-nitroso-N-ethylurea
CAS Number
542-88-1
106-93-4
75-21-8
76-44-8
118-74-1
87-68-3


319-84-6
319-85-7
58-89-9
34465-46^8
7440-02-0

62-75-9
55-18-5
924-16-3
930-55-2
759-73-9
Humans
S
I
L
I
I
I


I
I
I
I
L

I
I
I
I
I
Animal s
S
S
S
S
S
L


S
L
L
S
S

S
S
S
S
S
Grouping
based on
IARC
criteria
1
2B
2A
2B
2B
3


2B
3
2B
2B
2A

2B
2B
2B
2B
28
SI opeb
(mg/kg/day)-l
9300(1)
41
3.5x10-1(1)
3.37
1.67
7.75x10-2

4.75
11.12
1.84
1.33
6.2xlO+3
1.15(W)

25.9(not by ql
43.5(not by qj
5.43
2.13
32.9
Molecul ar
weight
115
187.9
44.1
373.3
284.4
261

290.9
290.9
290.9
290.9
391
58.7

') 74.1
f) 102.1
' 158.2
100.2
117.1
Potency
index0
1x10+6
8x10+3
2x10+1
1x10+3
5x10+2
2x10+1

1x10+3
3xlO+3
5x10+2
4x10+2
2x10+6
7x10+1

2xlO+3
4xlO+3
i O
9xlO+2
2x10+2
4x10+3
Order of
magnitude
(Iog10
index)
+6
+4
+1
+3
+3
+1

+3
+3
+3
+3
+6
+2

+3
+4
+3
+2
+4
                                                                                          (continued on the following page)

-------
                                                    TABLE 8-37.  (continued)
CO
(—'
co
Compounds
N-nitroso-N-methylurea
N-nitroso-diphenyl amine
PCBs
Phenols
2 ,4 ,6-Tr i chl orophenol
Tetrachl orodi benzo-
p-dioxin (TCDD)
Tetrachl oroethyl ene
Toxaphene
Tri chl oroethyl ene
Vinyl chloride
CAS Number
684-93-5
86-30-6
1336-36-3
88-06-2
1746-01-6
127-18-4
8001-35-2
79-01-6
75-01-4
of
Level
evidence3
Humans Animals
I
I
I
I
I
I
I
I
S
S.
S
S
S
S
L
S
L/S
S
Grouping
based on
IARC
criteria
2B
2B
28
2B
2B
3
2B
3/2B
1
SI opeb
(mg/kg/day)'1
302.6
4.92x10-3
4.34
1.99x10-2
1.56x10+5
5.1xlO-2
1.13
1.1x10-2
1.75x10-2(1)
Molecular
weight
103.1
198
324
197.4
322
165.8
414
131.4
62.5
Potency
index0
3x1 0+4
1x10°
lxlO+3
4x10°
5x1 0+7
8x10°
5xlO+2
1x10°
1x10°
Order of
magnitude
(Io9in
index)
+4
0
+3
+1
+8
+1
+3
0
0
    aS = Sufficient  evidence;  L  = Limited evidence; I = Inadequate evidence.
    ^Animal  slopes are 95% upper-bound slopes based on the linearized multistage model.  They are calculated based on
      animal  oral  studies, except for those indicated by I (animal inhalation), w (human occupational  exposure), and H
      (human  drinking water exposure).  Human slopes are point estimates based on the linear nonthreshold model.  Not all
      of the  carcinogenic potencies presented in this table represent the same degree of certainty.  All are subject to
      change  as new evidence becomes available.  The slope value is an upper bound in the sense that the true value (which
      is unknown)  is  not likely to exceed the upper bound and may be much lower, with a lower bound approaching zero.
      Thus, the use of the slope  estimate in risk evaluations requires an appreciation for the implication of the upper
      bound concept as well as  the "weight of evidence" for the likelihood that the substance is a human carcinogen.
    cThe potency  index is a rounded-off slope in (mmol/kg/day)-1 and is calculated by multiplying the slopes in
      (mg/kg/day)"1 by the molecular weight of the compound.

-------
8.5.2  Potency Index
     The potency index for TCI based on mouse hepatocellular carcinomas in
the NTP and NCI gavage studies is 1.4 x 10°.  This is derived as follows:   the
mean slope estimate from both studies, 1.1 x 10~2 (mg/kg/day)~l, is multiplied
by the molecular weight of 131.4 to give a potency index of 1.4 x 10°.   Round-
ing off to the nearest order of magnitude gives a value of 1()0, which is the
scale presented on the horizontal axis of Figure 8-15.  The index of 1.4 x 10°
is among the least potent of the 54 suspect carcinogens, ranking in the lowest
quartile.  Ranking of the relative potency indices is subject to the uncertain-
ties involved in comparing estimates of potency for a number of chemicals  based
on different routes of exposure in different species, using studies whose
quality varies widely.  All  of the indices presented here are based on  estimates
of low-dose risk, using linear extrapolation from the observable range.  These
indices may not be appropriate for the comparison of potencies if linearity
does not exist at the low-dose range, or if comparison is to be made at the
high-dose range.  If the latter is the case, then an index other than the  one
calculated above may be more appropriate.

8.6  SUMMARY
8.6.1  Qualitative Asssessment
8.6.1.1  Animal Studies —In  a 90-week carcinogenicity bioassay with male and
female B6C3F1 mice given estimated maximally (2,339 mg/kg in males  and  1,739
mg/kg in females) and one-half maximally tolerated doses of technical grade,
epoxide-stabilized TCI in corn oil daily, 5 days/week, for 78 weeks, statisti-
cally significant increases  in the incidence of hepatocellular carcinomas  in
male and female mice compared to corn oil controls were found (NCI, 1976).
Statistically significant increases in hepatocellular carcinoma incidence  in
                                    8-137

-------
treated male and female mice compared to vehicle-control  mice  were  also  ob-
served in a repeat of the above carcinogenicity  bioassay  in  which male and
female B6C3F1 mice were treated with  purified  TCI,  containing  no detectable
epoxides, in corn oil by gavage at  a  dose of 1,000  mg/kg/day,  5 days/week, for
103 weeks (NTP, 1982).  In effect,  this  repeat study  indicated that epoxide
stabilizers were not necessary factors in the  TCI-induced increases in hepa-
tocellular carcinoma incidence in male and female B6C3F1  mice  under the  con-
ditions of these bioassays.  It is  noted that  spontaneous hepatocellular tumor
formation occurs in the B6C3F1 mouse  strain, particularly in the male.
     A 110-week carcinogenicity study with male  and female Osborne-Mendel  rats
given technical grade, epoxide-stabilized TCI  in corn oil  by gavage at esti-
mated maximally (1,097 mg/kg) and one-half maximally  tolerated doses daily, 5
days/week, for 78 weeks, was negative for carcinogenicity; however, this study
may be inconclusive because of high mortality  in the  treatment groups (NCI,
1976).  In a carcinogenicity study  on male and female Fischer  344  rats treated
with purified TCI, containing no detectable epoxides, in  corn  oil  by gavage at
estimated maximally (1,000 mg/kg) and one-half maximally  tolerated  doses for
103 weeks, a small incidence of renal adenocarcinomas in  high-dose  males at
terminal sacrifice (3/16) was calculated to be a statistically signifi-
cant increase over that in matched  vehicle treated  controls  (0/33)  by life
table and incidental tumor tests, which adjust for  mortality,  but  not by the
Fisher Exact Test (NTP, 1982).  However, the NTP,  sponsor of the  study,  has
concluded that this study is inadequate to evaluate the presence  or absence of
a carcinogenic response to TCI.  Under the conditions of  this  study in Fischer
344 rats, there was a statistically significant  increase  in  mortality in
treated males  and 2%, 6%, and 20% of the matched vehicle-control,  low-dose,  and
                                    8-138

-------
high-dose males, respectively, were accidentally killed during the study.
Toxic nephrosis characterized as cytomegaly was commonly found in the treated
male and female Fischer 344 rats, with a stronger effect evident in males.   A
carcinogenicity study of purified TCI administered by gavage in olive oil  to
male and female Sprague-Dawley rats daily, 4 or 5 days/week for 52 weeks,  at
doses of 250 or 50 mg/kg followed by lifetime observation was negative, but
the 52-week exposure period was below potential lifetime exposures for these
animals (Maltoni, 1979).
     Inhalation exposure of male Han:NMRI mice, male and female Han:Wist rats,
and male and female Syrian hamsters to purified TCI vapor at levels of 100  ppm
and 500 ppm 6 hours/day, 5 days/week, for 78 weeks followed by lifetime obser-
vation did not induce a carcinogenic effect (Henschler et al., 1980).  Overt
toxicity of TCI was not evident in rats and hamsters, which suggests that
testing at higher levels of TCI was possible in this study, but mortality
was greater in male mice compared to controls.   A statistically significant
increase in malignant lymphoma incidence was found in low-dose and high-dose
female Han:NMRI mice compared to controls in this study;  however, the sponta-
neous incidence of malignant lymphomas in controls was high (30%), and it was
postulated in the study report that the increased incidence of malignant lym-
phomas in female mice could have been a result  of an immunosuppressive response
capable of enhancing susceptibility to tumor induction by specific, mostly
inborn, viruses (Henschler et al., 1980).
     A chronic inhalation study of technical  grade, epoxide-stabilized TCI  as
a vapor .was done in which male and female Charles River rats and  male and
female B6C3F1 mice were exposed to nominal  levels of TCI  at 100,  300, and 600
ppm 6 hours/day, 5 days/week, for 24 months (Bell  et al.,  1978).   An  audit  of
liver pathology revealed no carcinogenic effect in rats and a  dose-related
                                    8-139

-------
increase in the incidence of hepatocellular tumors in male and female B6C3F1
mice; however, this study is weakened by wide variability in the actual  TCI
exposure levels and lack of a concurrent matched control  group of mice.
     A series of studies done to evaluate the carcinogenic potential  of  puri-
fied TCI in lifetime treatment studies with female ICR/Ha Swiss mice  included
an initiation-promotion study in which a single topical  application of 1 mg
TCI in acetone was followed by lifetime repeated topical  applications with the
promoter phorbol myristate acetate;  a study of complete  carcinogenicity  with
repeated topical applications of 1 mg TCI in acetone; a  study with once  weekly
subcutaneous injections of 0.5 mg TCI in trioctanoin; and a study in  males as
well as females with once-weekly gavage administrations  of 0.5 mg TCI in
trioctanoin (Van Duuren et al., 1979).  Complete carcinogenic and initiating
activities for TCI were not found in these studies.   The dose used for topical
application was below that estimated as maximally tolerated, but the  authors
concluded that no activity or low activity of chloroolefins tested for carci-
nogenicity on mouse skin would be expected because of a  relatively low level
of epoxidizing enzymes' in mouse skin.  More than once-weekly dosing by gavage
and subcutaneous administration could have permitted a broader evaluation of
TCI carcinogenicity.
     Carcinogenic activity for TCI was not observed in several animal studies
designed for an evaluation of general toxicity; however, these studies involved
small groups and durations of treatment and observation  that were less than
the lifetime of the animals or, in one study, a duration that was not speci-
fied.  Exposure of newborn Wistar rats for 10 weeks to 2,000 ppm TCI  or  vinyl
chloride, 8 hours/day, 5 days/week,  induced ATPase-deficient foci, reported  as
premalignant lesions, in liver only  in rats exposed to the latter agent  (Laib
et al., 1978, 1979); however, while only vinyl chloride was effective with this

                                     8-140

-------
exposure duration, it is not certain whether longer exposure to TCI could
ultimately have induced similar foci.
     The carcinogenic potential of TCI oxide, a putative direct-acting meta-
bolite of TCI, has also been evaluated in lifetime treatment studies with
female ICR/Ha Swiss mice (Van Duuren et al., 1983).  These studies included an
initiation-promotion study in which a single topical  application of 1 mg TCI
oxide in acetone was followed by lifetime repeated topical  applications with
phorbol  myristate acetate; a study of complete carcinogenicity with repeated
topical  applications of 2.5 mg TCI oxide in  acetone;  and a study with once-
weekly subcutaneous injections of 0.5 mg TCI oxide in-trioctanoin.  These
studies  were negative for initiating activity and complete carcinogenic activ-
ity at estimated maximum tolerated doses.

8.6.1.2   Cell Transformation Studies—Cell transformation activity by TCI,  at
minimally toxic doses, and 3-methylcholanthrene, a positive control  agent,  was
found in an in vitro study with Fischer rat  embryo cells .(F1706, subculture
108) containing genetic information of the Rauscher leukemia virus (Price et
al., 1978).  Cell transformation has been developed as an end point for predic-
ting carcinogenic activity; however, it is not certain whether these agents
elicited transformation in this Fischer rat  cell  line through an action on  the
host cells, their integrated oncornavirus, or both.
     Exposure of baby Syrian hamster kidney  (BHK-211C1 13)  cells in the
presence of rat liver microsomes in vitro to TCI  solution in DMSO at doses
including those producing toxicity did not induce cell  transformation,  whereas
vinyl  chloride induced cell transformation when administered as  a gas at levels
including those producing toxicity, but not  as a solution in DMSO tested at
levels where toxicity was not clearly evident (Styles,  1979).   A broader
                                    8-141

-------
evaluation could have been made with an additional  test where cells were
exposed to TCI as a vapor.
     Cell transformation was induced by high concentrations (>1 mM) of TCI
oxide capable of producing dose-related cytotoxicity in cultures of Syrian
hamster embryo cells.  The transformed colonies in  this system were similar
to those induced by benzo[a]pyrene and other chloroalkene epoxides  (DiPaolo
and Doniger, 1982).

8.6.1.3  Human Studies—Three cohort studies of workers exposed to  TCI did  not
find that these workers experienced an excess risk  of cancer (Axelson et al.,
1978; Tola et al., 1980; Malek et  al., 1979).  Each of these studies, however,
suffered from one or more of the following deficiencies:   small  sample size,
lack of analysis by tumor site, problems with exposure definition,  and problems
with length of exposure.
     A case-control study of malignant lymphoma cases provided some suggestion
of an association of TCI exposure  with malignant lymphoma (Hardell  et al.,
1981).  However, failure by the authors to adjust for age and the possibility
of inaccurate and biased exposure  classification limit the usefulness of the
results.  A study of 56 liver cancer cases found no evidence of TCI exposure
(Novotna et al., 1979); however, no controls were used in the study.
     Paddle (1983) found that none of the primary liver cancer cases residing
in an area near an industrial plant that produced TCI had been employed by  the
plant.  This study had several deficiencies, and there are also limitations
in the author's description of the study which make it difficult to evaluate.
No controls were used in the study.  The author provided no indication of the
probability of finding a liver cancer case who had  been an employee of the
plant and had been exposed to TCI.  Also, the author provided no information on
                                     8-142

-------
 the  distribution  of workers at the plant exposed to TCI by length and time of
 exposure.   Furthermore, persons with liver cancer who may have worked at the
•plant  but who  did not  live in the area of the plant or who had moved from the
 area were excluded from the study by study definition.
      In  summary,  the epidemiologic data are inadequate for evaluating the car-
 cinogenicity of TCI.
 8.6.2   Quantitative Assessment
      Four sets of gavage  bioassay data on hepatocellular carcinomas in male
 and  female  mice  (NTP,  1982 and NCI, 1976) are used to calculate the upper-bound
 slope  estimate for TCI from the linearized multistage model.  Since the upper-
 bound  slope estimates  calculated on the basis of these data sets are comparable,
 ranging  from 5.8  x 10~3 to 1.9 x 10~2 mg/kg/day, the geometric mean, 1.1 x 10-2
 mg/kg/day is used to calculate the incremental lifetime cancer risk (i.e., unit
 risk)  due to a unit exposure of trichloroethylene in drinking water and in air.
 The  upper-bound estimate  of the cancer risk due to 1 yg/L of TCI in drinking
 water is 3.2 x 10~7.   The upper-bound estimate of the cancer risk due to 1 yg/
 m^ of TCI in air  is 1.3 x 10'^.  The relevant information on metabolism and
 kinetics for TCI  by oral  and inhalation exposures has been used to calculate
 the  unit risks for drinking water and air.
      None of the  epidemiologic studies reviewed in this report provide posi-
 tive evidence  from which  to estimate a unit risk for exposure to TCI.  As an
 alternative, an upper-bound estimate has been calculated from one negative
 study.  The calculation on the basis of human data results in a greater risk
 estimate than  the estimate from animal data and thus does not appear to con-
 tradict  the risk  estimate calculated from the animal data.
      The potency  index of TCI is 1 x 10°, ranking it in the lowest quartile of
 54 chemicals that the  CA6 has evaluated as suspect carcinogens.

                                     8-143

-------
8.7  CONCLUSIONS
     Evidence reviewed in this document  for the  carcinogenicity  of TCI  in
experimental  animals includes:  increases  in the incidences  of hepatocellular
carcinoma in male and female B6C3F1  mice (three  studies), malignant  lymphoma in
female Han: NMRI mice, and renal  adenocarcinoma  in  male  Fischer  344  rats.   The
gavage treatment study of purified TCI  in  male and  female B6C3F1 mice was  done
with basically the same experimental  design as the  gavage treatment  study  of
technical grade TCI in male and female  B6C3F1 mice  to evaluate the role of
epoxide stabilizers in the induction of  hepatocellular carcinomas.   Similar
carcinogenic responses were observed in  both studies. However,  as discussed
earlier in this chapter, a third study  in  B6C3F1 mice, the  inhalation study,
was weakened by deficiencies in its  conduct.  In the inhalation  study in
Han:NMRI mice, there was a 30% incidence of spontaneous  lymphoma in  control
mice and the possibility of an indirect  effect in treated mice;  the  gavage
treatment study in Fischer 344 rats  was  considered  by the study's sponsor, NTP,
to be inadequate for a judgment of TCI  carcinogenicity because of experimental
deficiences (See Section 8.1).
     While six epidemiologic studies or related  surveys  have focused on TCI
exposure, the total body of evidence is  considered  to be inadequate  to  evaluate
the carcinogenic potential.  Three cohort studies and two surveys showed no
excess risk of cancer resulting from exposure to TCI; however, in each  case,
study deficiencies limited their usefulness and/or diminished their  sensitivity
to detect a response.  A case control study with a suggestive association
between exposure and malignant lymphoma has flaws which  limit the usefulness
of the results.
     Cell transformation activity by TCI in Fischer rat  embryo  cells containing
rat oncornavirus was found, but it is not certain whether the effect could have

                                     8-144

-------
occurred in the host genetic material,  the oncornavirus,  or both.   As  concluded
in this document, currently available data provide  suggestive  evidence that
commercial  grade TCI is weakly active as an indirect  mutagen,  i.e.,  it requires
metabolic activation.  Data on pure TCI do not  allow  a  conclusion  to be drawn
about its mutagenic potential; although if it is  mutagenic, the  available  data
suggest that pure TCI would be a very weak indirect mutagen.   As discussed
in the metabolism chapter herein,  there is evidence that  TCI metaolite(s)  can
react with  cellular protein macromolecules in vivo  and  in vitro  and  with
exogenous DNA and RNA in vitro.  There are no known differences  among  species
with regard to metabolic pathways  or metabolic  profiles for TCI.  A  study  of
purified radio-labeled TCI binding to DNA in liver  in vivo in  male B6C3F1  mice
given a gavage dose of 1,200 mg/kg of TCI - 14C(7,500 yCi/animal)  suggested
a low order of binding.  Cell transformation with high  doses of  TCI  oxide, a
putative direct-acting carcinogenic metabolite  of TCI,  has been  observed j_n_
vitro; however, animal studies of  initiating and  complete carcinogenic poten-
tial of TCI oxide were negative.
     TCI is positive for the induction of malignant tumors of  the  liver in both
male and female B6C3F1 mice in multiple studies.  This  constitutes a signal that
TCI might be carcinogenic in humans.  The positive  lymphoma response in females
of the Han:NMRI mouse strain does  qualify as a  separate strain,  but  that
result is merely suggestive of a carcinogenic response  because it  occurs only
in females, has a high spontaneous rate, and has  been found to involve an
indirect mechanism of action.  The kidney tumor results in the NTP rat study
are not strong indications of a response in a second  species because of the
small number of animals responding (3 of 49 animals)  and  because of  the high
mortality,  although the statistical  significance  after  mortality corrections
suggests that a carcinogenic effect may be taking place.

                                     8-145

-------
     The TCI carcinogenicity results  could be  classified  under  the  criteria  of
the International  Agency for Research  on  Cancer (IARC)  as either  "sufficient"
or "limited" depending on which  of the differing current  scientific views
about chlorinated  organic compound induction of liver tumors  in mice is  chosen.
Since there are no adequate epidenriologic data in humans, the overall  ranking
of TCI under the criteria of the IARC  depends  primarily upon  the  position  taken
regarding the mouse liver tumor.  Thus, the overall  ranking could be either
Group 2B or Group  3.   The more conservative public health view  would regard
TCI as a probable  human carcinogen (Group 2B), but there  is also  scientific
sentiment for regarding TCI as an agent that cannot  be  classified as to  its
carcinogenicity for humans (Group 3).
     The U.S. Environmental Protection Agency's Proposed  Guidelines for  Carci-
nogen Assessment (U.S. EPA, 1984) regard  the mouse-liver-tumor-only response
as sufficient evidence for carcinogenicity in  animals.  A variety of factors
such as lack of mutagenicity, low malignancy,  and other such  measures,  if
present, would downgrade this evidence; however, downgrading  would  not  occur
with the TCI results.  Thus, based on  EPA's proposed cancer guidelines,  the
overall evidence for TCI results in a  classification of B2, i.e., a probable
human carcinogen.
     Additional carcinogenicity studies on purified TCI (no detectable  epox-
ides) including those sponsored by the NTP with lifetime  gavage treatment  in
four strains of rats (Osborne-Mendel,  Marshall 540,  August 28807, and ACI)
and those performed in the laboratories of Dr. Caesar Maltoni with  lifetime
inhalation exposure in B6C3F1 mice, Swiss mice, and Sprague-Dawley  rats  are
not finalized or published.  The Fukuda et al. (1983) and Henschler et  al.
(1984) studies were not available at the  time  this document was submitted  for
public comment and Science Advisory Board review; thus, these studies have not

                                     8-146

-------
been included in the final  document  but will  be  incorporated into any updates
of this Health Assessment Document.
     Gavage studies on hepatocellular carcinomas  in mice provide a basis to
culate incremental  lifetime cancer risks  due  to  exposure to TCI.  The develop-
ment of the risk estimates is  for the purpose of  evaluating the "what-if"
question:   If TCI is carcinogenic in humans,  what  is the possible magnitude of
the public health impact?  The upper-bound  estimate of the incremental cancer
risk due to 1 yg/L of TCI in drinking water is 3.2 x 10~7.  The upper-bound
estimate of the incremental cancer risk due to 1  yg/m^ of TCI in air is
1.3 x 10~6.  The upper-bound nature  of these  estimates is such that the true
risk is not likely to exceed this value and may  be lower.
     The potency index of TCI  is  1 x 10^, ranking  it in the lowest quartile of
54 chemicals that the CAG has  evaluated as  suspect carcinogens.
                                    8-147

-------
8.5  REFERENCES

Adams, E.M., H.C. Spencer,  V.K.  Rowe,  D.D.  McCollister,  and  D.D.  Irish.
     1951.  Vapor toxicity  of trichloroethylene determined by  experiments  on
     laboratory animals.   A.M.A. Arch. Indus.  Hyg.  Occup. Med.  4:469-481.

Andersen, M.E., M.L. Gorgas,  R.A. Jones,  and L.J.  Jenkins.   1980.   Determina-
     tion of the kinetic  constants for metabolism  of inhaled toxicants  in  vivo
     using gas uptake measurement.  Toxicol. Appl.  Pharmacol.  54:100-116.

Axelson, 0.  1980.  Letter  to Dr. Larry Anderson,  Carcinogen Assessment  Group,
     U.S. Environmental  Protection Agency.   November 14,  1980.

Axelson, 0.  1983.  Letter  to Mr. Herman Gibb,  Carcinogen Assessment  Group,
     U.S. Environmental  Protection Agency.   August  17,  1983.

Axelson, 0., K. Anderson, C.  Hogstedt, B. Holmberg,  G.  Molina,  and  A. de Verdier.
     1978.  A cohort study  on trichloroethylene exposure  and cancer mortality.
     J. Occup. Med. 20:194-196.

Beliles, R.P.  1975.  The metals.  In:  Casarett and Doull  (eds.).  Toxicology:
     the basic science of poisons.  New York:  Macmillan  Co., Chapter  18.

Bell, Z.G., K.J. Olson,  and T.J. Benya.  1978.   Final  report of audit findings
     of the Manufacturing Chemists Associaton  (MCA):   Administered  trichloro-
     ethylene (TCE) chronic inhalation study at Industrial Bio-Test Laboratories,
     Inc., Decatur, Illinois.  Unpublished.

Buben, J.A., and E.J. O'Flaherty.  1985.   Delineation  of  the role of  metabolism
     in the hepatotoxicity  of trichloroethylene and  perchloroethylene:   A  dose-
     effect study.  Toxicol.  Appl. Pharmacol.  78:105-122.

Cox, C.R.  1972.  Regression  model and life tables.   J.  Roy. Stat.  Soc.  B
     34:187-220.

Daffer, P.; Crump, K.; Masterman, M.  (1980)  Asymptotic  theory for analyzing
     dose-response survival data with  application  to low-dose  extrapolation
     problems.  Mathematical  Biosciences 50:207-230.

Davidson, I.W.F.  1984.   Biological basis for extrapolation  across  mammalian
     species.  EPA Issue  Paper.   U.S.  Environmental  Protection Agency,
     Washington, DC.

Dekant, W., M. Metzler,  and D. Henschler.  1984.  Novel  metabolites of  tri-
     chloroethylene through dechlorination  reactions in  rats,  mice, and  humans.
     Biochem. Pharmacol.  33:2021-2027.

Diem, K., and C. Lentner, eds.  1970.   Scientific  tables.  7th ed.  Basel,
     Switzerland: Ciba-Geigy  Ltd.

DiPaolo, J.A., and J. Doniger.  1982.   Neoplastic  transformation  of Syrian
     hamster cells by putative epoxide metabolites  of commercially  utilized
     chloroalkenes.  J.  Natl. Cancer Inst.  69:531-534.
                                     8-148

-------
DiRenzo, A.B., A.J. Gandolfi, and I.G.  Sipes.   1982.   Microsomal  bioactiva-
     tion and covalent binding of aliphatic halides to DNA.   Toxicol.  Lett.
     11:243-252.

Doll, R.  1971.  Wei bull  distribution of cancer.   Implications for models  of
     carcinogenesis.  J.  Roy. Stat. Soc. A 13:133-166.

Fernandez,  J.G., P.O. Droz, B.E. Humbert, and  J.R.  Caperos.   1977.  Tri-
     chloroethylene exposure.  Simulation of uptake,  excretion,  and metabo-
     lism using a mathematical model.  Brit. J.  Ind.  Med.  34:43-55.

Filser, J.G., and H.M. Bolt.  1979.  Pharmacokinetics of halogenated ethylenes
     in rats.  Arch. Toxicol. 42:123-136.

Fukuda, K., K. Takemoto,  and H. Tsuruta.  1983.   Inhalation  carcinogenicity of
     trichloroethylene in mice and rats.  Ind. Health 21:243-254.

Green, T.,  and M.S. Prout.  1984.  Species differences in  response to  tri-
     chloroethylene.  II.  Biotrans format ion in  rats  -and mice.  Toxicol. Appl.
     Pharmacol.  In press.

Hardell, L., M. Eriksson, P. Lenner,  and E. Lundgren.  1981.   Malignant  lym-
     phomas and exposure  to chemicals,  especially organic  solvents, chloro-
     phenols and phenoxy  acids: a case-control study.  Br. J.  Cancer 43:169-
     176.

Hathaway, D.E.  1977.  Comparative mammalian metabolism of vinyl  chloride  and
     vinylidene chloride  in relation  to oncogenic potential.   Environ. Health
     Perspect. 21:55-59.

Hathaway, D.E.  1980.  Consideration  of the evidence  for mechanisms of 1,1,2-
     trichloroethylene metabolism, including new  identification  of its dichlo-
     roacetic acid and tricloroacetic acid metabolites in  mice.   Cancer  Lett.
     8:263-269.

Henschler,  D., H. Elsasser, W. Romen,  and E.  Eder.   1984.  Carcinogenicity
     study  of trichloroethylene, with and without epoxide  stabilizers, in
     mice.   J. Cancer Res. Clin. Oncol. 104:149-156.

Henschler,  D., W. Romen,  H.M. Elasser,  D. Reichert, E. Eder,  and Z. Radwan.
     1980.   Carcinogenicity study of  trichloroethylene by  long-term inhalation
     in the animal  species.  Arch. Toxicol. 43:237-248.

Howe, R.  1983.  GLOBAL83: an experimental program developed  for the U.S.
     Environmental  Protection Agency  as an update to  GLOBAL82: a  computer
     program to extrapolate animal toxicity data  to low  doses  (May 1982).
     K.S. Crump and Co.,  Inc., Ruston,  LA.  Unpublished.

Kaplan, E.L., and P. Meier.  1958.  Nonparametric estimation  of  incomplete
     observations.   J. Amer. Stat. Assoc. 53:457-481.

Katz, R.M., and D.  Jowett.  1981.  Female laundry and dry-cleaning workers in
     Wisconsin:  A mortality analysis.   Am J.  Public  Health 71:305-307.


                                     8-149

-------
Klassen, C.D., and G.L.  Plaa.   1966.   The relative  effect  of  various  chlori-
     nated hydrocarbons  on liver and  kidney  function  in  mice.   Toxicol.  Appl.
     Pharmacol. 9:129-131.

Klassen, C.D., and G.L.  Plaa.   1967.   Relative  effects of  various  chlorinated
     hydrocarbons on liver and kidney function  in dogs.  Toxicol.  Appl.  Phar-
     macol. 10:119-131.

Kline, S.A., and B.L. Van Duuren.   1977.   Reaction  of epoxy-l,l,2-trichloro-
     ethane with nucleophiles.  J.  Heterocycl.  Chem.  14:455-458.

Kline, S.A., J.J. Solomon, and B.L. Van Duuren.   1978.   Synthesis  and reactions
     of chloroalkene epoxides.  J.  Org. Chem. 43:3596-3600.

Krueger, G.R.F.  1972.  Chronic immunosuppression and lymphomagenesis in man
     and mice.  Natl. Cancer Inst.  Monogr. 35:183-190.

Laib, R.J., G. Stockle,  and H.M. Bolt.   1978.   Induction of pre-malignant
     hepatic lesions:  Comparative  effects of vinyl chloride  and trichloro-
     ethylene.  20th Congress  of the  European Society of Toxicology.   West
     Berlin.  Abstract 67.

Laib, R.J., G. Stockle,  H.M. Bolt., and W. Kung.  1979.  Vinyl  chloride  and
     trichloroethylene:   Comparison of  alkylating effects  of  metabolites and
     induction of preneoplastic enzyme  deficiencies in rat liver.   J. Can.
     Res. Clin. Oncol.  94:139-147.

Lin, R.S., and I.I. Kessler.  1981.   A  multifactorial model for pancreatic
     cancer in man.  J.  Am. Med. Assoc. 245:147-152.

Luz, A.  1977.  The range of incidence  of spontaneous neoplastic and  nonneoplas-
     tic lesions of the  laboratory  mouse.  Z. Versuchstierkd  19:342-343.

Malek, B., B. Krcmarova, and 0. Rodova.  1979.   An  epidemiological  study of
     hepatic tumor incidence in subjects  working with trichloroethylene.  II.
     Negative result of  retrospective investigations  in  dry-cleaners.  Prakov.
     Lek. 31:124-126.

Maltoni, C.  1979.  Results of long-term carcinogenicity bioassays of trichloro-
     ethylene experiments by oral administration  on Sprague-Dawley rats.  In
     press.

Mantel, N., and M.A. Schneiderman.   1975.  Estimating "safe"  levels,  a hazardous
     undertaking.  Cancer Res. 35:1379-1386.

Miller, R.E., and P.P. Guengerich.   1982.  Oxidation  of  trichloroethylene by
     liver microsomal cytochrome P-450:  evidence for chlorine  migration in
     a transition state  not involving trichloroethylene  oxide.  Biochemistry
     21:1090-1097.

Monsinger, M., and H. Fiorentini.   1955.   Reaction  hepatique,  renales,
     ganglionnaires et spleniques dans  1'intoxication experimentale par  le
     trichloroethylene chez le chat.   C.  R.  Soc.  Biol.  (Paris)  149:150-172.


                                     8-150

-------
Monster, A.C., G. Boersman, and W.C. Duba.  1976.  Pharmacokinetics of tri-
     chloroethylene in volunteers:  Influence of workload and exposure con-
     centration.  Int. Arch. Occup. Environ. Health 38:87-102.

Monster, A.C., G. Boersman, and W.C. Duba.  1979.  Kinetics of trichloroethy-
     lene in repeated exposure of volunteers.  Int. Arch. Occup. Environ.
     Health 42:283-292.

National Cancer Institute (NCI).  1976.  Carcinogenesis bioassay of trichloro-
     ethylene.  CAS No. 79-01-6.  NCI-CG-TR-2.

National Toxicology Program (NTP).  1982.  Carcinogenesis bioassay of trichloro-
     ethylene.  CAS No. 79-01-6.  NTP 81-84.  NIH Publication No. 82-1799.   Draft,

Nomiyama, K., and H. Nomiyama.  1977.  Trichloroethylene metabolism in man  and
     animals, with special reference to host and agent factors modifying the
     trichloroethylene metabolism.  Ind. Environ. Xenobiot. 2:173-176.

Novotna, E., A. David, and B. Malek.  1979.  An epidemiological  study on
     hepatic tumor incidence in subjects working with trichloroethylene: I.
     Negative results of retrospective investigations in subjects with primary
     liver carcinoma.  Pracovni Lekarstvi 31(4):121-123.

Paddle, G.M.  1983.  Incidence of liver cancer and trichloroethylene manufacture:
     joint study by industry and a cancer registry.  British Medical Journal
     286:846.

Parchman, C.G., and P.N. Magee.  1982.  Metabolism of 14C-trichloroethylene to
     1 CO2 and interaction of a metabolite with liver DNR in rats and mice.
     J. Toxicol. Environ. Health 9:797-813.

Parker, J.C., and I.W.F. Davidson.  1984.  Extrapolation of the  carcinogenic
     response.  Abstract of paper presented at the annual meeting of the Society
     for Risk Analysis:  Uncertainty in Risk Assessment, Risk Management, and
     Decision Making, held in Knoxville, TN, Sept. 30-Oct. 4, 1984.

Prendergast, J.A., R.A. Jones, L.J. Jenkins Jr.,  and J. Siegel.   1967.  Effects
     on experimental animals of long-term inhalation of trichloroethylene,
     carbon tetrachloride, 1,1,1-trichloroethane, dichlorodifluoromethane,  and
     1,1-dichloroethylene.  Toxicol. Appl. Pharmacol. 10:270-289.

Price, P.J., C.M. Hassett, and J.I. Mansfield.  1978.  Transforming activities
     of trichloroethylene and proposed industrial alternatives.   In Vitro
     14:290-293.

Prout, M.S., W.M. Provan, and T. Green.   1984.  Species differences in response
     to trichloroethylene.  I.  Pharmacokinetics  in rats and mice.   Toxicol.
     Appl.  Pharmacol.  In press.

Purchase, I.F.H., E. Longstaff, J. Ashby, J.A. Styles,  D. and P.A.  Lefevre,
     and F.R. Westwood.  1978.  An evaluation  of  6 short-term tests for
     detecting organic chemical carcinogens and recommendations  for their
     use.  Br. J. Cancer 37:873-959.


                                     8-151

-------
Rudali, G.  1967.  Potential  carcinogenic hazards from drugs.   Evaluation of
     risks.  U.I.C.C. Monograph Series 7:138-143.

Sanders, V.M, A.N. Tucker, K.L. White, Jr., B.M. Kauffman,  P.  Hallett, R.A.
     Carchman, J.F. Borzelleca, and A.E. Munson.  1982.  Humoral  and cell-
     mediated immune status in mice exposed to trichloroethylene  in the drinking
     water.  Toxicol. Appl. Pharmacol. 62:358-368.

Seifter, J.  1944.  Liver injury in dogs exposed to trichloroethylene.  J.
     Indus. Hyg. Toxicol. 26:250-253.

Smith, G.F.  1966.  Trichloroethylene.  A review.  Br. J. Ind. Med. 23:249-
     262.

Stott, W.T., J.F. Quast, and P.G. Watanabe.  1982.  Pharmacokinetics and macro-
     molecular interactions of trichloroethylene in mice and rats.  Toxicol.
     Appl. Pharmacol. 62:137-151.                     '  ,,.,.,.

Styles, J.A.  1977.  A method for detecting carcinogenic organic  chemicals
     using mammalian cells in culture.  Br. J. Cancer 36:558-563.

Styles, J.A.  1979.  Cell transformation assays.  Pages 147-164 in G.E. Pagett,
     ed.  Topics in Toxicology:  Mutagenesis in Sub-mammalian  Systems:  Status
     and Significance.  University Park Press, Baltimore, MD.

Tola, S., R. Vilhuner, E. Jaruinen, andpM.L. Korkale.  1980.  A cohort study
     on workers exposed to trichloroethylene.  J. Occup. Med.  22:737-740.

Tucker, A.N., V.M. Sanders, D.W. Barnes, T.J. Bradshaw, K.L. White, L.E. Sain,
     J.F. Borzelleca, and A.E. Munson.  1982.  Toxicology of trichloroethylene
     in the mouse.  Toxicol. Appl. Pharmacol. 62:351-357.

U.S. Environmental Protection Agency.   1984 (Nov. 23).  Proposed  guidelines
     for carcinogen risk assessment.  Federal Register 49:46294-46301.

Van Duuren, B.L., B.M. Goldschmidt, G. Lowengart, A.C. Smith,  S.  Melchionne,
     I. Seldman, and D. Roth.  1979.  Carcinogenicity of halogenated olefinic
     and aliphatic hydrocarbons in mice.  J. Natl. Cancer Inst. 63:1433-1439.

Van Duuren, B.L., S.A. Kline, S. Melchionne, and I. Seldman.  1983.  Chemical
     structure and carcinogenicity relationships of some chloroalkene oxides
     and their parent olefins.  Cancer Res. 43:159-162.

Waters, E.M., H.B. Geistner, and J.E.. Huff.  1977.  Trichloroethylene.  I.
     An overview.  J. Toxicol. Environ. Health 2:671-707.
                                     8—152               * USOOVERNMENTPHINHNQOFFICE-1W5- 559-111/20607

-------