United States
            Environmental Protection
            Agency
            Office of Health and
            Environmental Assessment
            Washington DC 20460
EPA/600/8-82/005F
July 1985
Final Report
            Research and Development
&EPA
Health Assessment
Document for
Tetrachloroethylene
(Perchloroethylene)
 Final
 Report

-------

-------

-------
                                    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 are 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.
     In view of  the  pending  release  of National Toxicology Program  reports on
long-term animal studies with tetrachloroethylene, the carcinogenicity conclu-
sions of this  document  are  considered interim  and  will  be  updated when the
reports can be evaluated.
                                      111

-------
                               CONTENTS
LIST OF TABLES	   vi i
LIST OF FIGURES	   ix
AUTHORS AND REVIEWERS.	   xi

1.   EXECUTIVE SUMMARY	   1-1

2.   INTRODUCTION.	   2-1

3.   GENERAL BACKGROUND INFORMATION	   3-1
     3.1  PHYSICAL AND CHEMICAL PROPERTIES	   3-1
     3.2  PRODUCTION	   3-1
     3. 3  USE	   3-3
     3.4  EMISSIONS	   3-3
     3.5  ENVIRONMENTAL FATE AND TRANSPORT	   3-4
          3.5.1  Ambient Air	   3-4
          3.5.2  Water		„.   3-7
     3. 6  LEVELS OF EXPOSURE	   3-9
          3.6.1  Mixing Ratios	   3-9
     3.7  ANALYTICAL METHOD.	   3-18
          3.7.1  Ambient Air..	   3-18
          3.7.2  Water	   3-22
          3.7.3  Biological Media	   3-24
          3.7.4  Cal ibration	   3-24
          3.7.5  Storage and Stability of PCE	   3-25
     3.8  REFERENCES	   3-26

4.   ECOSYSTEM CONSIDERATIONS	   4-1
     4.1  EFFECTS ON AQUATIC ORGANISMS AND PLANTS	   4-1
          4.1.1  Effects on Freshwater Species	   4-1
          4.1.2  Effects on Aquatic Plants	   4-2
          4.1.3  Effects on Saltwater Species	   4-2
     4.2  BIOCONCENTRATION AND BIOACCUMULATION	   4-3
          4.2.1  Levels of PCE in Tissues of Aquatic Species	   4-4
     4.3  BEHAVIOR IN WATER AND SOIL	   4-13
     4.4  SUMMARY.	   4-15
     4.5  REFERENCES	   4-16

5.   MAMMALIAN METABOLISM AND PHARMACOKINETICS.	   5-1
     5.1  ABSORPTION AND DISTRIBUTION	   5-1
          5.1.1  Dermal Absorption	   5-1
          5.1.2  Oral Absorption	   5-2
          5.1.3  Pulmonary Absorption	   5-3
          5.1.4  Tissue Distribution and Concentrations	   5-13
     5.2  EXCRETION	   5-15
          5.2.1  Pulmonary Elimination in Man	   5-15
          5.2.2  Urinary Metabolite Excretion in Man	   5-18
          5.2.3  Chronic Exposure	   5-20
          5.2.4  Excretion Kinetics in the Rodent	   5-21
     5.3  MEASURES OF EXPOSURE AND BODY BURDEN	   5-23


                                      1v

-------
                             CONTENTS (continued)
     5.4  METABOLISM	    5-27
          5.4.1  Known Metabolites	    5-27
          5.4.2  Enzymic Pathways of Metabolism	    5-28
          5.4.3  Magnitude and Dose-Dependency on Metabolism	    5-32
          5.4.4  Covalent Binding	    5-37
          5.4.5  Interactions with Metabolism	    5-40
     5.5  SUMMARY	    5-43
     5.6  REFERENCES	    5-45

6.    TOXIC EFFECTS	    6-1
     6.1  HUMANS	    6-1
          6.1.1  Effects on the Liver	    6-1
          6.1.2  Effects on Kidneys.,	    6-3
          6.1.3  Effects on Other Organs/Tissues	    6-3
          6.1.4  Effects on CNS and Behavior	    6-4
     6. 2  LABORATORY ANIMAL STUDIES	    6-7
          6.2.1  Lethality and Anesthesia	    6-7
          6. 2.2  Behavioral Effects	    6-18
          6.2.3  Effects on the Liver and Kidney	    6-19
          6.2.4  Effects on the Heart	    6-23
          6.2.5  Effects on the Skin and Eye	    6-24
     6.3  ADVERSE EFFECTS OF SECONDARY POLLUTANTS	    6-24
     6.4  SUMMARY OF ADVERSE HEALTH EFFECTS AND ASSOCIATED LOWEST
            OBSERVABLE EFFECT CONCENTRATIONS	    6-25
           6.4.1  Inhalation Exposure	    6-25
           6.4.2  Oral Exposure	    6-26
           6.4. 3  Dermal Exposure	    6-26
     6.5   REFERENCES	    6-27

7.    TERATOGENICITY, EMBRYOTOXICITY, AND REPRODUCTIVE EFFECTS	    7-1
     7.1  ANIMAL STUDIES	    7-2
          7.1.1  Mice	    7-2
          7.1.2  Rats	    7-2
          7.1.3  Rabbits	    7-6
     7.2  SUMMARY	    7-9
     7.3  REFERENCES	    7-10

8.    MUTAGENICITY 	    8-1
     8.1  GENE MUTATION TESTS	    8-1
          8.1.1  Bacteria	    8-1
          8.1.2  Drosophila	    8-12
     8.2  CHROMOSOME ABERRATION TESTS	    8-13
          8.2.1  Whole-Mammal Bone Marrow Cells	    8-13
          8.2.2  Human Peripheral Lymphocytes	    8-14
          8.2.3  Drosophila	    8-15
     8. 3  OTHER TESTS INDICATIVE OF DNA DAMAGE	    8-15
          8.3.1  DNA Repair	    8-15
          8.3.2  Mitotic Recombination	    8-18
     8.4  DNA BINDING STUDIES	    8-21
     8.5  STUDIES INDICATIVE OF MUTAGENICITY IN GERM CELLS	    8-21

-------
                             CONTENTS (continued)
     8. 6  MUTAGENICITY OF METABOLITES	           8-22
     8. 7  SUMMARY AND CONCLUSIONS	        8-23
     8. 8  REFERENCES	   8-25

9.   CARCINOGENICITY	   9-1
     9.1  ANIMAL STUDIES	   9-1
          9.1.1  National Cancer Institute Bioassay (1977a)	   9-2
          9.1.2  Dow Chemical  Company Inhalation Study (Rampy et al.,
                  1978)	   9-12
          9.1.3  Intraperitoneal Administration Study (Theiss et al.
                  1977)	   9-14
          9.1.4  Skin Painting Study (Van Duuren et al. ,  1979)	   9-16
     9.2  EPIDEMIOLOGIC STUDIES	   9-17
          9.2.1  Kaplan (1980)	   9-18
          9.2.2  Blair et al.  (1979)	   9-22
          9.2.3  Katz and Jowett (1981)	   9-23
          9.2.4  Lin and Kessler (1981)	   9-24
          9.2.5  Dun and Asal  (1984)	   9-24
          9.2.6  Asal (personal communication,  1985)	   9-26
     9.3  RISK ESTIMATES FROM ANIMAL DATA	   9-27
          9.3.1  Selection of Animal Data	   9-28
          9. 3.2  Interspecies Dose Conversion	   9-30
          9.3.3  Choice of Risk Model	   9-46
          9.3.4  Unit Risk for a Compound	   9-51
          9.3.5  Alternative Low-Dose Extrapolation Models	   9-53
     9.4  CALCULATION OF UNIT RISK	   9-53
          9.4.1  Potency (Slope) Estimate on the Basis of Animal Data.   9-53
          9.4.2  Risk Associated with 1 ng/L of PCE in Drinking  Water.   9-56
          9.4.3  Risk Associated with 1 ug/m3 of PCE in Air	   9-56
     9.5  RISK PREDICTION AGAINST HUMAN EXPERIENCE	   9-63
     9.6  COMPARISON OF POTENCY WITH OTHER COMPOUNDS	   9-65
     9. 7  SUMMARY	   9-71
          9. 7.1  Qualitative	   9-71
          9.7.2  Quantitative	   9-72
     9.8  CONCLUSIONS	   9-73
     9.9  REFERENCES	   9-75
                                       VI

-------
                                LIST OF TABLES


Table                                                                 Page

3-1  Major U.S.  producers of PCE	     3-3
3-2  Ambient air mixing ratios of PCE measured at sites around
      the world	     3-10

4-1  Levels of PCE in tissues of marine organisms, birds,  and
      mammals	'.	     4-5
4-2  Accumulation of PCE by dabs	     4-9
4-3  Concentration of PCE and trichloroethylene in mollusks and
      fish near the Isle of Man	     4-12

5-1  Physical properties of PCE and other chloroethylenes	     5-2
5-2  Estimated uptake of six individuals exposed to PCE
      at rest and after exercise	     5-10
5-3  Disposition in rats of 14C-PCE radioactivity 72 hr following
      oral or inhalation administration	     5-12
5-4  Disposition in mice of 14C-PCE radioactivity 72 hr following
      oral or inhalation administration	     5-13
5-5  Rat organ content of PCE after daily inhalation exposure
      of 200 ppm for 6 hours per day	     5-14
5-6  Disposition in rats of 14C-PCE radioactivity 72 hr following
      drinking water ingestion	      5-23
5-7  Reported metabolites of PCE (other than TCA)	      5-24
5-8  Metabolism of PCE by rat hepatic microsomes and the
      effect of various inducers	      5-31
5-9  Disposition of 36C1-PCE after oral administration to
      Wistar rats	      5-35
5-10 Irreversible hepatic binding of 14C-PCE in rats 72 hr
      after exposure	      5=39
5-11 Comparison of irreversible hepatic binding of 14C-PCE in
      Sprague-Dawley rats and B6C3F1 mice after inhalation
      (6 hr) and oral exposures	      5-40
5-12 Effects of chronic oral administration of PCE on hepatic
      DNA content and DNA synthesis in mice and rats	      5-41

6-1  Summary of the effects of PCE on animals	      6-8
6-2  Toxic dose data	      6-14

7-1  Summary of reproductive/teratogenic effects of PCE in
      laboratory animals	     7-7

8-1  Summary of mutagenicity testing of PCE	     8-2
8-2  Results of bacterial tests of different purities and
      sources of PCE	     8-3

9-1  Incidence of hepatocellular carcinomas in B6C3F1 mice fed
      PCE	     9-4
9-2  Cumulative survival of Sprague-Dawley rats exposed to PCE
      for 12 months	     9-13
9-3  Pulmonary tumor response to PCE	     9-15
                                      vn

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


Table

9-4  Incidence of hepatocellular carcinomas in male and female
      B6C3F1 mice fed PCE by gavage		      9-29
9-5  Disposition of 14C-PCE radioactivity for 72 hours after sin-
      gle oral doses to Sprague-Dawley rats and B6C3F1 mice......      9-33
9-6  Disposition of 14C-PCE radioactivity for 72 hours after
      inhalation exposure for 6 hours to Sprague-Dawley rats and
      B6C3F1 mice	      9-39
9-7  Total hepatic macromolecular binding following oral exposure
      of rats and mice to 500 mg/kg 14C-PCE	      9-43
9-8  Lifetime average daily exposure (LAE) for B6C3F1 mice in the
      NCI bioassay, assuming 17% of dose is metabolized	      9-45
9-9  Lifetime average daily exposure (LAE) for B6C3F1 mice in the
      NCI bioassay, using a dose-metabolism (in urinary excretion)
      relationship.	      9-45
9-10 Dose-response data and potency (slope) estimates	      9-54
9-11 Relative carcinogenic potencies among 54 chemicals evaluated
      by the Carcinogen Assessment Group as suspect human
      carcinogens	    9-67

-------
                            LIST OF FIGURES


Figure                                                                  Page

4-1  Accumulation and loss of PCE by dabs	   4-10
4-2  Relation between flesh and liver concentrations of PCE
       in dabs	   4-11

5-1  First-order excretion curves for PCE in blood of rats after
       exposure to 600 ppm for 6 hr or to 500 mg/kg gavage doses	   5-4
5-2  PCE concentrations in blood and exhaled air following
       i nhal ati on exposure for 4 hr.	   5-5
5-3  Predicted uptake and distribution of PCE to tissue groups
       during and after an 8-hr exposure to 100 ppm..................   5-6
5-4  PCE alveolar air concentration during exposure of 5 subjects
       for 8 hr at 100 ppm	   5-9
5-5  Mean exhaled breath concentrations of PCE for five volunteers
       exposed to 100 ppm PCE for 7 hr per day for 5 days	   5-11
5-6  Daily (8-hr) occupational inhalation exposure to PCE............   5-17
5-7  Trichloroacetic acid blood concentrations following inhalation
       exposure to PCE for 4 hr	   5-19
5-8  Urinary excretion of trichloroacetic acid during and following
       inhalation exposure to PCE for 4 hr.			   5-19
5-9  Relationship between PCE occupational inhalation exposure and
       urinary concentration of total trichloro-compounds at end of
       work shift for 36 male and 25 female workers.	   5-20
5-10 Daily (8-hr) occupational inhalation exposure to PCE	   5-21
5-11 Relationship between PCE and dose and the amount of total
       urinary metabolite excreted per day by mice.	   5-25
5-12 Direct linear relationship between the time-weighted average
       occupational exposure to PCE over the last 4 hr of a work
       day and the concentration of PCE in exhaled air 15-30 min
       after the end of exposure.	   5-26
5-13 Postulated scheme for the metabol i sm of PCE.	   5-29
5-14 Production of TCA from PCE by hepatic microsomas from differ-
       ently pretreated rats as a function of time.	   5-31
5-15 Relationship between hepatotoxicity parameters from PCE oral
       administration and total urinary metabolite excreted per
       day by mice at increasing dose levels of PCE	   5-38

6-1  Concentration-time curves for various effects of PCE	   6-16
6-2  Concentration-time curves for PCE-induced anesthesia	   6-17

8-1  Dose response curves of Perchlor 200 and Perchlor 230 using
       Salmonella typhimurium tester strains TA100 and TA1535....	   8-7
8-2  Induction of mitotic recombination by PCE in Saccharomyces
       cerevisiae D7	   8-19
                                      IX

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


Figure                                        '                          Page

9-1  Growth curves for male and female mice in the PCE chronic
       study 	    9-5
9-2  Survival comparisons of male and female mice in the PCE
     chronic study 	    9-7
9-3  Growth curves for male and female rats in the PCE chronic
       study 	    9-8
9-4  Survival comparisons of male and female rats in the PCE
       chronic study 	    9-9
9-5  Relationship between the PCE dose and the amount of total
       urinary metabolite excreted per day by mice in each group	    9-36
9-6  Histogram representing the frequency distribution of the
       potency indices of 54 suspect carcinogens evaluated
       by the Carcinogen Assessment Group	    9-66

-------
                             AUTHORS AND REVIEWERS


The principal authors of this document are:


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 and Pharmacology, The Bowman Gray
     School of Medicine, Wake Forest University, Winston-Salem, N.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, N.C.

Charalingayya B. Hiremath, Carcinogen Assessment Group, U.S. Environmental
     Protection Agency, Washington, D.C.

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

Vicki Vaughan-Dellarco, Reproductive Effects Assessment Group, U.S. Environ-
     mental Protection Agency, Washington, D.C.


The following individuals reviewed earlier drafts of this document and submitted
valuable comments.
All Members of the
Interagency Regulatory Liaison Group
Subcommittee on Organic Solvents

Dr. Joseph Borzelleca
Dept. of Pharmacology
The Medical College of Virginia
Virginia Commonwealth University
Richmond, VA  23298

Dr. Mildred Christian
Argus Laboratories
Perkasie, PA  18944

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

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

Dr.  Lawrence Fishbein
National Center for lexicological Research
Jefferson, AR  72079

Dr.  Thomas Haley
National Center for Toxicology Research
Jefferson, AK  72079

Dr.  Rudolf Jaeger
Institute of Environmental Medicine
New York, NY  10016

Dr.  John G. Keller
P. 0. Box 10763
Research Triangle Park, NC  27709

Dr.  John  L. Laseter
Director, Environmental Affairs, Inc.
New Orleans, LA  70122

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

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

Dr. James Withey
Food  Directorate
Bureau  of Food Chemistry
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.
Chao W.  Chen, Ph.D.
Margaret M.L. Chu, Ph.D.
I.W.F.  Davidson, Ph.D
Herman J. Gibb, B.S. ,
Bernard H. Haberman,  D.V.M., M.S
Charalingayya B. Hiremath, Ph.D.
Robert E. McGaughy, Ph.D.
Charles H. Ris, M.S., P.E.
Dharm W. Singh, D.V.M., Ph.D.
Todd W.  Thorslund, Sc.D.
 (consultant)
M.P.H.
Participating Members of the Reproductive Effects Assessment Group

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

Members of the Agency Work Group on Solvents
Elizabeth L. Anderson
Charles H. Ris
Jean Parker
Mark Greenberg
Cynthia Sonich
Steve Lutkenhoff
Arnold Edelman
James A. Stewart
Paul Price
William Lappenbusch
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 Tcxic 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 Water Regulations and Standards
             Office of Solid Waste
             Office of Pesticide Programs
             Office of Emergency and Remedial Response
                                      XI11

-------
                            1.  EXECUTIVE SUMMARY

     Tetrachloroethylene (PCE) is a moderately volatile chlorinated hydro-
carbon which has important commercial applications in the dry cleaning of
fabrics and in the degreasing of fabricated metal parts.  It  is  estimated that
approximately 265,000 metric tons were produced in the United States in 1982.
Approximately 90 percent of production is estimated to be released  eventually
to the atmosphere.  Because PCE is relatively insoluble in water (150 mg/L)
and has a vapor pressure of 19 torr at 25°C, PCE in natural waters  would  be
conveyed to the atmosphere rapidly, through evaporation.  There  are no known
or expected natural sources of emissions.
     PCE has been detected in the ambient (natural environment)  air of a
variety of urban and nonurban areas of the United States and  other  regions of
the world.  Levels can range from trace amounts in rural areas to as much as
10 parts per billion (ppb) or 0.068 mg/m^ in some large urban centers.  The
global average background level is estimated at about 25 parts per  trillion
(ppt) or 0.175 x 10~3 mg/m^.  PCE has been detected less frequently in  water;
it has been monitored in surface and drinking waters, generally  at  levels
between 1 and 2 ppb.  In certain instances involving contamination  of ground-
water, much higher levels have been reported.  Although there is very limited
information on the behavior of PCE in soil, PCE can be expected  to  leach
through soils of low (< 0.1 percent) organic carbon content.   The amount  of
PCE adsorbed to soils is dependent on the partition coefficient, the organic
carbon content, and the concentration of PCE in the liquid phase.
     In the troposphere, a region of the atmosphere extending to between  8
and 16 kilometers above the earth's surface, PCE undergoes photochemical
degradation to the extent that its estimated lifetime is appreciably less
than 1 year.  Little PCE is expected to be conveyed to the stratosphere.
Recent studies have shown that, in real atmospheres, neither  atomic chlorine-
nor hydroxy radical-induced photooxidation of PCE generates substantial
concentrations of ozone or other oxidants; thus, PCE is not believed to be a
significant factor in production of photochemically induced pollution often
experienced near large urban centers.  Because of the reduced solar flux  in
winter and seasonal variations in hydroxy radical concentration, PCE levels
in ambient air are expected to be higher in winter than in summer.   On a  daily

                                     1-1

-------
 basis,  PCE  levels fluctuate considerably.
      Inhalation is the principal route of concern by which PCE enters the body.
 Ingestion of drinking water contaminated by PCE is another important concern.
 PCE  is  rapidly and virtually completely absorbed following oral administration,
 while pulmonary uptake of PCE during inhalation exposure is linearly propor-
 tional  to exposure duration and the air concentration; pulmonary uptake is
 also influenced by physical activity and body mass.  The metabolism and
 pharmacokinetics of PCE are highly contingent on its physicochemical proper-
 ties.  Controlled inhalation studies with human volunteers (at 100 ppm)
 suggest that whole-body, steady-state conditions are not established within
 short exposure periods (e.g., 8 hours).  Because partitioning into adipose
 tissue  is slow, steady-state (the rate at which whole body uptake is balanced
 by clearance) may require considerably longer periods of exposure.  PCE
 distributes widely into body tissues.  PCE is eliminated from the body mainly
 by the  pulmonary excretion of unchanged parent compound.  Limited metabolism
 of PCE  occurs; metabolism is dose-dependent and saturable in mice, rats, and
 humans;  in  humans saturation would not be expected until air exposure levels
 approximate 100 ppm (678 mg/m3) or greater.  The principal site of metabolism
 is in the liver where PCE is oxidized to PCE oxide, which rearranges to
 trichloroacetic acid.  Controlled studies with humans have demonstrated that
 PCE metabolism (urinary trichloroacetic acid) represents 1 to 3 percent of
 the amount of PCE absorbed during 8-hour exposures to between 100 to 400 ppm
 (678 to  2,112 mg/m3)-.  While the metabolic profiles of PCE are not as yet
 fully established in mice, rats, and humans, there is no convincing evidence
 of differences in the pathways in these species.  PCE metabolites are known
to covalently bind, in vitro and in vivo, to cellular macromolecules such as
protein and lipid.  Since tissue-bound metabolites have a slow rate of turn-
over, cumulative cellular changes may occur in humans with chronic exposure.
 Indices of hepatotoxicity of PCE in the rat and mouse have been shown to be
highly correlative with the dose-dependent nature of PCE metabolism.
     Excluding carcinogenicity as an end point, toxicity testing in experi-
mental animals,  coupled with limited human data derived principally from
overexposure situations,  suggest that long-term exposure of  humans  to
                                     1-2

-------
environmental levels found or expected (10 ppb or less)  of PCE is not  likely
to present a health concern.
     Decrements in task performance and coordination are the first gross  signs
of central nervous system (CNS) and behavioral alterations observed in  con-
trolled human studies in which individuals were exposed  to about  100 ppm  (678
mg/m^) for up to 7 hours.  More sensitive tests, however, would have to be  per-
formed to determine if PCE affects the nervous system at even lower concentra-
ti ons.
     Transient liver damage in humans is generally associated with short-term
exposures greatly in excess of 100 ppm (678 mg/m^).   In  rodent species  tested,
intermittent or prolonged exposure to PCE has been observed to result in  liver
and kidney damage at levels exceeding 200 ppm (1,356 mg/m^).  Since ambient
air levels are generally orders of magnitude lower than  that associated with
liver damage, long-term exposure of humans to ambient air or water levels
would not be expected to cause adverse liver or kidney effects.
     Similarly, ambient air and water levels of PCE  are  unlikely  to cause
adverse effects upon the heart.
     The mammalian animal tests performed to date do not indicate any signifi-
cant teratogenic potential of PCE in the species tested.  On this basis, there
is no evidence to suggest that the conceptus is uniquely susceptible to the
effects of PCE.  The anatomical effects observed primarily reflect delayed
development  and generally can be considered reversible.   It is important  to
note, however, that the reversible nature of an embryonic/fetal  effect  in  one
species might, in another species, be manifested in  a more serious and  irre-
versible manner.  The teratogenic potential of PCE for humans is  unknown.
     PCE has been evaluated for its ability to cause gene mutation, chromoso-
mal aberrations, unscheduled DNA synthesis, and mitotic  recombination.  These
tests were conducted using bacteria, Drosophi1 a, yeast,  cultured  mammalian
cells, whole mammal systems, and cytogenetic analyses of exposed  humans.   Cer-
tain technical and commercial samples of PCE elicited increased responses  in
the Ames bacterial test, a yeast mitotic recombination assay, a host-mediated
assay using  Salmonella, and DNA repair tests.  Exogenous metabolic activation
was not required for detection of these increased effects.  In general, the
responses were weak and were observed at high concentrations that were  cyto-
toxic; dose-dependent relationships were not established.  The positive

                                     1-3

-------
 findings  may  be  the  result of mutagenic contaminants and/or added stabilizers.
 Several  other tests  of  commercial and technical samples of PCE have been
"reported  as  negative.   The epoxide of PCE, which is thought to be the active
 biological  intermediate, was found to be positive in bacterial studies.
      In  a gavage bioassay, PCE has induced a statistically significant increase
 of malignant  liver tumors in both male and female B6C3F1 mice.  No carcinogenic
 effects  were  observed in  lifetime studies of rats exposed to PCE in gavage and
 inhalation  bioassays; however, these latter studies had diminished sensitivity
 to detect a  response due to excessive dose-related mortality in the gavage
 study and a  low  dose level in the inhalation study.
      Intraperitoneal injection of PCE in Strain A mice did not produce an
 increased incidence  of  pulmonary adenomas,; mouse skin initiation-promotion
 experiments  also did not show a tumor response.  However, because of inherent
 limitations  in these assays, the negative results do not detract from the
 positive  findings of liver tumors in mice.
      A cohort study  of  dry-cleaning workers exposed to PCE showed that workers
 were at  an  elevated  risk of colon cancer mortality; however, the elevated risk
 cannot be totally assigned to PCE since as many as one-half of the workers may
 have been exposed to petroleum distillates in their working history.  Other
 studies that  found an association between cancer and employment in the dry-
 cleaning  industry did not identify the dry-cleaning solvents to which the
 employees were exposed.
      The  positive response in both male and female mice constitutes a signal
that  PCE  and/or  its  reactive metabolites might be a carcinogen for humans.  In
terms  of  strength of evidence in animal test systems, the mouse bioassay
constitutes limited  evidence.
     According to the Agency's Proposed Guidelines for Carcinogen Risk
Assessment, the  cancer  evidence of PCE in animal test systems is limited, and
the cancer evidence  in  epidemiologic studies is inconclusive.  The overall
weight-of-evidence classification for PCE would be Group C, i.e., a possible
human carcinogen.
     When the  criteria of the International  Agency for Research on Cancer
(IARC) are applied,  the animal  data supporting the carcinogenicity of PCE is
classified as  limited.   Because existing epidemiologic data for PCE is
inconclusive,  its overall  IARC  ranking  should  be  Group 3,  meaning  that the

                                     1-4

-------
evidence is inadequate to evaluate the carcinogenic potential  of PCE.
     The NTP is completing PCE inhalation cancer bioassays for rats  and mice.
These findings should be available for scientific review in the fall  of 1985.
When the results of these bioassays are made available,  consideration  will  be
given to updating the carcinogenicity evaluation in this assessment  document.
     Although PCE has only limited carcinogenicity evidence, a carcinogenic
potency and related upper-bound estimate of incremental  lifetime cancer risk
can be estimated from the NCI male and female mouse bioassay.
     The development of these risk estimates is for the  purpose of evaluating
the "what if" question:  If PCE is carcinogenic in humans, what is the possible
magnitude of the public health impact?  Any use of the risk estimates  should
include a recognition of the weight-of-evidence likelihood for the carcinogenic
potential of PCE in humans.  The upper-bound incremental cancer risk  is
calculated to be 5.0 x 10~2 (mg/kg/day )"•'- for a continuous lifetime  exposure
to PCE, under the presumption that PCE is carcinogenic in humans.   The upper-
bound nature of this estimate is such that the true risk is not likely to
exceed this value, and may be lower.  Expressed in terms of relative  potency,
PCE ranks in the lowest quartile among the 54 suspect or known human  carcino-
gens evaluated by EPA's Carcinogen Assessment Group.
                                     1-5

-------
                               2.  INTRODUCTION

     Tetrachloroethylene  (PCE)  is one  member of a  family of  unsaturated
chlorinated aliphatic compounds.   Other common names/acronyms are perchloro-
ethylene, Perk, PER, and PERC.  Its synonyms include carbon dichloride, tetra-
chloroethene, and 1,1,2,2-tetrachloroethylene.
     PCE, though a water and solid waste contaminant, is primarily of interest
in ambient air  exposure situations.   It is  released  into  ambient air as a
result of  evaporative  losses  during production, storage,  and/or  use.  It is
not known to be generated from natural sources.  It has negligible photochemical
reactivity in the troposphere and is removed by scavenging mechanisms, princi-
pally via hydroxyl radicals.
     The scientific  data  base is  limited with  reference to the  effects of PCE
on humans.   Effects  on  humans have generally  been  ascertained  from studies
involving  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 PCE exposure studies have been directed toward eluci-
dating the effects  on  the central nervous  system, effects on clinical chem-
istries, and pharraacokinetic parameters of PCE exposure.
     Since epidemiologic  studies  have not been able to assess adequately the
overall  impact  of PCE  on human health, it has 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 among animal species increases the confidence that similar
effects may occur in humans.
     This document is intended to provide an evaluation of the scientific data
base concerning PCE.  The publications cited  in  this document represent a
majority of  the known scientific  references  to PCE.   Reports which  had little
or no bearing upon the issues discussed were not cited.
     The basic  literature  search that supports this assessment is current up
to 1984.   On-going  literature  searches have  been  conducted  through 1985,
resulting  in  the inclusion of selected references  in all  chapters  of this
document.
                                    2-1

-------
                      3.    GENERAL BACKGROUND INFORMATION

3.1  PHYSICAL AND CHEMICAL PROPERTIES
     Tetrachloroethylene, also called PCE (1,1,2,2-tetrachloroethylene or per-
chloroethylene), is a colorless,  heavy  liquid with a chloroform-1ike  odor.  It
is used as a solvent for organic substances and is commercially important as a
solvent in the  dry  cleaning of fabrics  and  in the degreasing  of  metals.  It
has a molecular weight of 165.85 and is relatively insoluble in water (150 mg/L)
(Handbook of  Chemistry  and Physics,  1976;  Chemical  and Process  Technology
Encyclopedia, 1974).  Its  specific gravity at 20°C is 1.624.  Its  CAS  registry
number is 127-18-4.   In air, at 25°C and standard pressure, 1 part per million
(ppm) is equivalent to 6.78 mg/m3.
     PCE  has  negligible  photochemical  reactivity (Dimitriades  et a!., 1983)
and, in  the  troposphere,  is decomposed via free radical mechanisms.   When in
contact with  water  for prolonged  periods, PCE slowly decomposes to yield tri-
chloroacetic  and hydrochloric  acids.  Upon prolonged storage in  light, it was
reported to decompose slowly to trichloroacetyl  chloride and phosgene by auto-
oxidation (Hardie,  1966).  At 700°C, it decomposes,  when in contact with acti-
vated charcoal,  to  hexachloroethane and hexachlorobenzene (Gonikberg, 1956).
PCE has a boiling point of 121.1°C at 760 mm Hg and a vapor pressure of 14 torr
at  20°C.  MacKay et al.  (1982)  have  calculated a vapor  pressure of 19 torr at
25°C.
     The  chemical reactivity of PCE  has  been discussed  by  Bonse and Henschler
(1976) in terms  of  the electron-inductive effect of the chlorine  atoms, which
reduce electron  density  about  the ethylene bond.  This  effect, in combination
with a steric protective effect afforded by the chlorine atoms,  provides in-
creased stability against electrophilic attack.   This  is  exemplified in the
reaction of PCE with ozone.  Compared to ethylene and less-substituted chlorina-
tion hydrocarbons,  PCE has  a  low  rate of reaction (Williamson  and Cvetanovik,
1968).

3.2  PRODUCTION
     PCE may be produced by several  processes:
     1.    Chlorination of trichloroethylene:
                                 onop
               CHC1  = CC12 + C12 frFf* CHC12CC13
                                    3-1

-------
          2CHC12CC13 + Ca(OH)2 ^-H-C12C = CC12 + CaCl2 + 2H20

     2.   Dehydrochlorination of S-tetrachloroethane:
          CHC12-CHC12 + C12 	» CC12 - CC12 + 2HC1

     3.   Oxygenation of S-tetrachloroethane:
          2CHC12CHC12 + 02 	> 2CC12 = CC12 + 2H20

     4.   Chlorination of acetylene:
          CC12 -  CC12 + C12 200^cc-,3-CC-|3
          CH  - CH +  3 CC1 CC1   2QO-400°C.4CC12 =  CC12  + 2HC1
          CH  - CH +  3 CU3LU3  Cata1yst

     5.   Chlorination of hydrocarbons:
          C3H8 +  8 C12 	>•   CC12  = CC12  + CC14 + 8HC1
          (propane)
               2CC14 	>   CC12  = CC12  + 2C12

     6.   Oxychlorination of  1,2-dichloroethane:
          2C2H4C12 + 5C12 	> C2H2C14 + C2HC15 + 5HC1
          C2H2C14 +  C2HC15  	> C2HC13 + 2HC1 + CC12 = CC12
          7HC1 +  1.75  02  	> 3.5 H20 + 3.5 C12
          2C2H4C12 + 1.5  C12  + 1.75 02 	»•  C2HC13 + CC12 = CC12 + 3.5 H20

     The  majority of PCE produced in the  United  States  is derived from the
oxychlorination   of  1,2-dichloroethane (reaction  6)  or via Chlorination of
hydrocarbons  (reaction  5)  (Lowenheim and Moran,  1975).
     In  1980, 329,000 metric  tons  of PCE were consumed in the United States
(SRI,  1982).   This  figure  represents production plus imports, minus exports.
According to  the  U.S.  International  Trade  Commission  (1983), 265,770 metric
tons were produced in the United States  in  1982.   Production  in 1983 was esti
mated at 263,000  metric tons (Chemical and  Engineering News,  1984).
                                    3-2

-------
                        TABLE 3-1.   MAJOR U.S.  PRODUCERS OF PCEC
     Organization
Yearly 1981 capacity, MTC
     Dow Chemical
     PPG
     Vulcan
     Diamond Shamrock
     Ethyl Corporation
     E.  I. du Pont de Nemours
     Stauffer Chemical
          145
           91
           91
           75


           73
           32
 Adapted from Chemical Economics Handbook (SRI, 1983).   MT = Metric tons.
 Terminated production (Halogenated Solvents Industry Alliance, 1983).
C0utput for captive use only.

3.3  USE
     PCE has  the following  uses  (Gosselin et al.,  1976;  Fishbein,  1977):
(1) dry cleaning  solvent;  (2)  textile  scouring solvent; (3) dried vegetable
fumigant;  (4) rug and  upholstery cleaner;  (5)  stain, spot,  lipstick, and rust
remover; (6) paint  remover;  (7)  heat  transfer media ingredient; (8)  chemical
intermediate in  the  production of  other organic compounds; and (9) metal  de-
greaser.
     It is  estimated  that  the  use  of PCE in the dry cleaning industry repre-
sents about 42  percent of  1980 consumption in the United States (SRI,  1982).

3.4  EMISSIONS
     Emissions of PCE  arise  during  its  production,  from its  use as a chemical
intermediate in industrial  processes,  from storage containers, during disposal,
and from its  use as a solvent.  Because emissions  are  almost  exclusively to
the atmosphere,  the information presented in this section focuses on air.   Data
available  concerning  discharges  to water are  discussed in Section 3.6.1.2.
Emissions  estimates  reflect  a  diversity of sources  throughout the country.
Dry cleaning operations  are  located primarily in urban areas.  Approximately
                                    3-3

-------
 26,000 establishments are estimated  to  exist,  according to Bureau of Census
 data (U.S. EPA, 1979).   There  are approximately 18,000 retail  establishments
 plus much smaller  numbers of  industrial  and coin-operated facilities (Inter-
 national  Fabricare Institute,  1984).
      In 1977, global  emissions  were estimated at 570,000 ± 285,000 metric tons
 (Singh et al., 1979).  It was also estimated that emissions accounted for ap-
 proximately 90 percent of the  amount of PCE produced  in the United States.

 3.5  ENVIRONMENTAL FATE AND  TRANSPORT
      The potential for ambient  air and  water mixing ratios of PCE to pose a
 hazard to human health is influenced by many processes.  Such factors include
 transformation into  secondary pollutants  of concern and degradation rates in
 air and water.

 3.5.1 Ambient Air
 3.5.1.1  Tropospheric Reactivity-- Reaction with the hydroxyl radical (-OH) is
 the principal process by which  many organic compounds,  including PCE, are scav-
 enged from the troposphere.   Hydroxyl radicals  are produced when 03 is  irradi-
 ated,  resulting in excited atomic oxygen, which then reacts with water vapor.
 The tropospheric  lifetime of a  compound  is related to the -OH mixing ratio ac-
 cording to the expression:
                                           _
where  k  is  the  rate  constant of reaction.
     Singh  et al.  (1979,  1981)  calculated a tropospheric residence of PCE of
about  68  days.   This calculation was based on an average 24-hour -OH abundance
of  106 molecules cm 3 in the boundary layer of a polluted atmosphere.  Justi-
fication  for this  -OH mixing  ratio stems from the  field studies of Calvert
(1976) and  from  Singh  and  coworkers (1979a).   Because this -OH mixing ratio is
more typical  of  summer months,  Singh et  al.  (1981)  suggested that a seasonally
adjusted  mixing  ratio would result in a longer chemical residence time.  If a
seasonally  averaged  -OH  mixing  ratio  of 4 x  105  molecules cm"3  (a  level  sup-
ported by the field  measurements of Campbell  et al.,  1979) and a  weighted global
                                    3-4

-------
average temperature  (265°K), an average  residence  time  of  PCE would  be  calcu-
lated to be 292 days or 0.8 year (Singh et al. ,  1979).
     Estimations of  a  residence time  of  PCE  of  one year or less  have been  re-
ported by a number of investigators (Dilling, 1982; Altshuller,  1980; Singh et
al., 1978a; Singh,  1977;  Crutzen  et al.  , 1978;  Lillian et al.,  1975; Yung et
al., 1975; Pearson and McConnell ,  1975).
     Dimitriades et  al.  (1983)  calculated a very  low tropospheric reactivity
for PCE based on observations that ambient levels  of PCE are constant.  Atmos-
pheric consumption of PCE is 0.02 percent per daylight  hour.
     Higher levels  of  -OH  have  been reported for the  southern hemisphere com-
pared to those in the northern hemisphere (Singh,  1978).  This gradient probably
is due to the fact that carbon monoxide  levels are much higher in the northern
hemisphere, thus  reducing  -OH  levels  (Singh, 1978).   Measurements of PCE and
other  reactive  halocarbons  indicate that mixing  ratios  are higher  in the nor-
thern  hemisphere  where the -OH mixing ratio is low and where most of the PCE
is released (Singh et al., 1978b).
     Chamber studies indicate that PCE, on irradiation  in the presence of other
atmospheric constituents,  can  be  transformed into secondary products.  This
area of  study  has been recently reviewed and further explored by Dimitriades
et al. (1983).   These investigators have confirmed that PCE, under smog chamber
conditions having  high reactant concentrations, reacts to  form  03 and  ozone
precursors by means of a Cl-initiated photooxidation mechanism.   However, such
photooxidation  is  not  expected  to  occur  in  the  real atmosphere at  a  rate high
enough for substantial  03  production.   It is the  authors'  contention that Cl
atoms  are  effectively  scavenged by the  hydrocarbons  normally present in the
atmosphere; thus, PCE was judged to contribute less to  03 production than equal
concentrations  of  ethane.   Ethane  is  regarded by the  authors  to  be a boundary
species  separating  the reactive volatile organics from the unreactive  ones.
     Studies on PCE reactions with 03, 0, and -OH  have  indicated that rate con-
stants are lower than with Cl (Dimitriades et al. , 1983).
     Gay et al. (1976) had determined that trichloroacetyl  chloride  was a photo-
oxidation product of PCE in smog chamber studies.  While the studies of Gay et
al. (1976) indicate that trichloroacetyl chloride  may be formed through chlorine
atom migration  in  an epoxide intermediate,  evaluation  of  -OH and oxygen atom
rate constants  indicate  that  less  than  1 percent of PCE in ambient air  reacts
with atomic oxygen and, of the activated epoxides  formed,  only a small percent-
age undergo rearrangement  (Graedel,  1978).   Dimitriades et  al.  (1983)  found
                                    3-5

-------
that,  on  irradiation,  the  only  product  observed was  CO.   Phosgene was  not de-
tected.  When  2  ppm  (13.6  mg/m3)  PCE  and  20  ppb trichloroacetyl  chloride were
.irradiated together, the phosgene level reached  0.1 ppm.
     Phosgene production from the photochemical  oxidation of PCE  in the  presence
of  other  substances  has  been  reported by  others (Lillian et al. ,  1975; Gay et
al., 1976).  The extent  to which phosgene may be formed  in real  atmospheres,
based  on smog chamber results, would  also be expected to  be minimal.
3.5.1.2   Tropospheric  Removal Mechanisms  for PCE--The  reaction  sequence  by
which  PCE may be scavenged from the troposphere  is as follows (Graedel,  1978):

        C2C14 +  HO- 	>   HOC(C1)2C(C1)2-
        HOC(C1)2C(C1)2- + 02 	»•   HOC(C1)2C(C1)202-
        HOC(C1)2C(C1)202- 	oxygen	»•  HOC(C1)2C(C1)20-              (3-2)
                           abstraction
        HOC(C1)2C(C1)20-  	>   HOC(C1)2- + COC12
        HOC(C1)2- + 02 	»•  COC12 + H02-

     Howard (1976) suggested that the reaction path for the atmospheric  oxida-
tion of PCE may  follow the scheme below,  leading to the production of oxalyl
chloride:

          C2C14 + -OH 	» C2C12OH-


                      	> CC12CC10H- + Cl-                           (3-3)


          CC12CC12OH-  + 02  	> 02CC12CC12OH


         02CC12CC12OH  +  NO 	> COC1CC12OH- + N02 + Cl•


          •OH + COC1CC12OH  	> COC1COC1  +  H20 + Cl•
                                   3-6

-------
Compared to  other  ethylene  compounds  studies,  Howard  (1976)  reported  that  PCE
exhibits unusually low reactivity toward hydroxyl radicals.
     Snelson et al.  (1978)  suggested that trichloroacetyl chloride and phos-
gene would  hydrolyze to  the  corresponding  chloroacetic acids and  hydrogen
chloride via homogeneous gas phase  hydrolysis.   The  acids then would pre-
sumably be washed out of the atmosphere.
     The environmental significance of the production of phosgene from PCE has
been discussed by  Singh  and coworkers (Singh et al.,  1975; Singh,  1976).   As
PCE emissions are  likely to be higher in urban areas, the reactivity of this
halocarbon may result in concentrations of phosgene in the low ppb  range during
adverse meteorological conditions  in and around urban centers (Singh, 1976).
It should be noted that overt adverse health damage would not be expected at
these phosgene levels.  Considering the smog chamber results of Dimitriades et
al.  (1983),  in which PCE was found to have negligible reactivity,  it appears
unlikely that phosgene  would  be produced at other  than trace  levels.  Singh
(1976) concluded that phosgene is removed slowly from the atmosphere.   Rainfall
appears to lower atmospheric levels of phosgene (Singh et al., 1977a).  On the
other hand,  phosgene has  been  reported  to hydrolyze to  C02 and HC1  rapidly in
liquid water.  Manogue (1958) reported a half-life of 0.1 second at 25°C.   Thus,
it is not expected to persist in the troposphere because of rainout and hydrol-
ysis in aqueous aerosols.
     The observed diurnal variations in PCE levels suggest that PCE has higher
mixing ratios in the morning and evening hours than at other times  (Lillian et
al., 1975; Singh et al., 1977a).  Ohta et al.  (1977) reported that  mixing ratios
tended to be highest on cloudy days and lowest on rainy days.
     Solar flux is a major factor in the rate at which PCE is removed from the
atmosphere.   Singh et al. (1977a) suggested that the reduced solar  flux in winter
months would permit  a much  longer transport of  PCE because  of reduced reac-
tivity.   The effect of solar flux was calculated by Altshuller (1980) who esti-
mated that a 1  percent consumption of PCE by reaction with -OH would take 14
days during  the month of January as opposed to one day in July.

3.5.2  Water
     Jensen  and Rosenberg (1975)  investigated  the degradability of  PCE (~  0.1
to 1.0 ppm)  in  both  open and closed systems,  with sea water and fresh water.
In open aquaria,  with 20 L of sea water, PCE levels decreased 50 percent within
                                    3-7

-------
 200 hours (daylight).   In a closed system, levels decreased approximately 25 per-
 cent over the  same  interval  (daylight).   It was reported that PCE  levels  in
 boiled,  deionized water in a closed system did not exhibit any significant de-
 crease after 8 days.   Analysis  was by headspace collection,  followed  by gas
 chromatography-electron capture detection  (GC-ECD)  quantification.   Accuracy
 and limits of  detection were not reported.  Variation  in detector sensitivity
 was checked daily by an injection of PCE  in hexane.
      Billing et al.  (1975) reported that  PCE in water slowly decomposes to form
 trichloroacetic and hydrochloric  acids.   The  evaporation rate was determined
 by  dissolving  1  ppm  (w/w) PCE in  200 ml of water.  Solution  temperature was
 approximately 25°C.   The solution was stirred  magnetically in a sealed system.
 Quantification was made  by mass  spectroscopy.   The evaporation rate of PCE,
 determined from  measurements  over  a 2-week period,  was characterized by a 50
 percent  decrease  in the  initial  mixing ratio  in 24 to 28 minutes.   The stir-
 ring speed had a  marked effect on  the evaporation rate.  With no stirring ex-
 cept for 15 seconds every 5 minutes, the  time  required for 50 percent depletion
 ranged from 72 to 90  minutes.   The evaporative half-life was 27 ± 3 minutes.
 Addition of dry,  granular bentonite clay (500 ppm) appeared  to increase the
 rate of disappearance  by 33 percent at 20 minutes.   However, when the clay was
 allowed to remain in contact with purified water for several days and then added
 to the solution,  there  was no change in the rate compared with control.  In
 closed-system  investigations,  Dilling et  al.  (1975)  used dry, powdered dolomi-
 tic limestone,  bentonite,  and peat moss  to determine the adsorption rate  for
 PCE.   With 500  ppm bentonite,  there was a 22 percent absorption after 30 minutes.
 There  was no further absorption.   Addition of  limestone resulted in a 50 percent
 depletion in 20 ± 2  minutes.   Addition of silica sand had no effect on the  dis-
 appearance  rate.   When 500 ppm peat moss was added, up to 0.4 ppm PCE was ab-
 sorbed after 10 minutes.   At longer times, no  further absorption was observed.
 It  was  concluded  that evaporation  probably is the major pathway by which  PCE
 is  lost from water.
     Parsons et al.  (1984) obtained evidence that PCE can be converted to other
chlorinated  compounds  following  incubation  of PCE  in  microcosms  containing
muck from an  aquifer recharge basin.   Results  suggested  that PCE can  be con-
verted to trichloroethylene, chloroethene, cis- and trans-l,2-dichloroethene.
and dichloromethane (methylene chloride).
                                    3-8

-------
     In reactivity  studies, Dilling  et al.  (1975)  found  that  sunlight  had  the
greatest effect on  the  rate of  PCE disappearance.  The PCE  level  in water  was
6 uM.  Over a 12-month  period,  in which  PCE  solutions were  stored  in the dark
as well as in the light, PCE levels decreased from 1 ppm (0 time), to 0.63 ppm
(6 months), to  0.35 ppm (12 months)  in  samples  stored  in the dark.   In the
light-exposed solution,  the  level  decreased to 0.52 ppm (6 months),  and 0.24
ppm (12 months).
     Schwarzenbach et al. (1979), in measurements of PCE levels in Lake Zurich,
reported findings compatible with those of Dilling et al. (1975) in that evapo-
ration is the dominant elimination process from surface waters. The annual  re-
lease  to  the atmosphere was estimated by  applying  a  steady-state mass  balance
model.  Based on  vertical  concentration  profiles from the lake, about 240 kg
PCE was released from the central basin annually.
     Wood et al. (1981) demonstrated that PCE in sediment-water samples can be
degraded  under  anaerobic conditions.  PCE was  not  degraded  in  autoclaved con-
trols.  Under experimental  conditions, the  resulting half-life  for PCE was 34
days.

3.6  LEVELS OF  EXPOSURE
3.6.1.  Mixing  Ratios
3.6.1.1   Ambient Air--A  wide  variety of  halogenated aliphatic  hydrocarbons,
including PCE,  have been detected in ambient air.  Ambient measurements of PCE
have been  conducted in both the United  States and other areas  of  the  world.
These  determinations  provide  a basis for assessing  the  levels to which human
populations may be exposed.
     Measured ambient air concentrations differ widely and undoubtedly reflect
the influences  of a variety of  factors, e.g., meteorological conditions, tropos-
pheric reactivity,  diurnal  variations,  sampling times,  and source emissions.
     Table 3-2  provides  summary information  regarding background and urban con-
centrations of  PCE.  It  should  be noted that, in general, these values reflect
short  sampling  times.
     Measurements of ground-level samples by Singh et al. (1978b), in both the
northern  and southern  hemispheres,  gave  average background levels of 0.040 ±
0.012  ppb (2.7  x 10"4 ± 0.08 x  10~4 mg/m3) and 0.012 ± 0.003 ppb (0.081 x 10~3
±0.02  x 10 3 mg/m3),  respectively.   Globally,  the  average background  level  of
                                    3-9

-------
TABLE 3-2.   AMBIENT AIR MIXING RATIOS OF PCE MEASURED AT SITES AROUND THE WORLD
Location
Alabama
Birmingham
Arizona
Grand Canyon3
Phoenix
California
San Bernardino Mtns.
oo Badger Pass
h"—*
O Point Arena
Stanford Hills
Point Reyes
Dominguez
El Cajon
La Jolla
Los Angeles
Menlo Park
Mill Valley
Mt. Cuyamaca
Date of Reported Concentration, ppb (mg/m3)
Measurement Maximum Minimum Average
April 12-22, 1977 0.008 0 0.001 ± 0.003
Nov. 28-Dec. 5, 1977 0 00
Apr. 23-May 6, 1979b 3.696 (0.025) 0.129 0.9938± 0.7155
(0.0008) (0.0067 ± 0.0048)
Fall, 1972 0.09 (6.1 x 10~4)
May 12-16, 1976 0.03 (2 x 10"4)
1976 0.03 (2 x 10~4)
Nov. 24-30, 1975 0.04 (2.6 x 10~4)
Dec. 2-12, 1975 0.043 (2.9 x 10~4)
May 14, 1976 2.9
April 9, 1975 0.31
Apr 9, 1974- Jan 6, 1976 2.3 0 0.53 ± 0.63
Sept. 22, 1972- 2.2 0.067 1.1 ± 0.45
April 19, 1979
Nov. 24-30, 1975 0.20 ± 0.21
Jan. 1-27, 1977 0.065 ± 0.075
Mar. 15, 1975 0.22
Reference
Pellizzari, 1979
Pellizzari, 1979
Singh et al. , 1981
Simmonds et al . , 1974
Singh et al . , 1977a
Singh et al. , 1978a
Singh et al . , 1977b
Singh et al. , 1977b
Pellizzari, 1977
Su and Goldberg, 1976
Su and Goldberg, 1976
Simmonds et al . , 1974;
Singh, 1976;
Singh et al. , 1977a;
Su and Goldberg, 1976
Singh et al . , 1977a
Singh et al. , 1979
Su and Goldberg, 1976

-------
TABLE 3-2.   (continued)





CO
1
1— >
1 — •







Location
California
Oakland
Palm Springs
Riverside
San Jose
Santa Monica
Upland
Colorado3
Denver
Delaware
Delaware City
Louisiana
Baton Rouge, Geismar,
and Plaquemine
Maryland
Baltimore
Michigan
Detroit
New Jersey
Bayonne
Date of
Measurement
June 30-July 8, 1979
May 5-11, 1976
April 25, 1977-
July 12, 1980
Aug. 21-27, 1978
April 6, 1974
Aug. 13-Sept. 23, 1977
Jun 16-26, 1980
July 8-10, 1974
Fall, 1978

July 11-12, 1974

Oct 27-Nov 5, 1978b
March, 1973-Dec. , 1973
Reported Concentration, ppb (mg/m3)
Maximum Minimum Average
0.64 0.12 0.31 ± 0.17
1.1 0.12 0.28 ± 0.084
0.98 0.37 0.49 + 0.13
1.1 ± 0.036
2.3
1.1 0.01 0.19 ± 0.35
0.47 0.24 0.39 ± 0.077
0.51 (0.0034) <0.02 (<0.0001) 0.24 (0.0016)
0.18 (0.001) .001 (0.007 x 10~3) 0.017 (0.118 X 10~~3)

0.29 (0.0019) <0.02 (<0.0001) 0.18 (0.0012)

2.16 (0.004) <0.1 (<0.001) 0.35 (0.002)
8.2 (0.0055) 0.30 (0.0020) 1.63 (0.0110)
Reference
Singh et al . , 1979;
Singh et al . , 1981
Singh et al. , 1978a
Singh et al . , 1979; 1980
Singh et al. , 1979
Su and Goldberg, 1976
Pellizzari , 1979
Singh et al. , 1980
Lillian et al . , 1975
Pellizzari et al . , 1979b

Lillian et al . , 1975

Evans et al. , 1979
Lillian et al . , 1975

-------
TABLE 3-2.   (continued)
Location
New Jersey
New Brunswick
New Brunswick
Seagirt
Sandy Hook
Boundbrook, Rahway,
Edison and Passaic
Batsto3
Bridgeport3
Burlington3
Camden3
Carlstadt3
Edison3

Elizabeth3

Fords3

Middlesex3
Newark3

Date of
Measurement

-
-
June 18-19, 1974
July 2, 1974

Sept 18-22, 1978
Feb. 26-Dec. 29, 1979
Sept. 22, 1977
Sept. 19, 1977
April 3-Oct. 24, 1979
Sept. 28-30, 1978
March 24, 1976-
Sept. 24, 1978
Sept. 15, 1978-
Dec. 29, 1979
Mar. 26, 1976-
Sept. 27, 1978
July 23-28, 1978
Mar. 23, 1976-
Dec. 29, 1979
Reported Concentration,
Maximum

-
-
0.88 (0.059)
1.4 (95 x 10~4)

58 (0.394)c
0.53
0.041

30
5.5
8.7

14

5.6

0.21
32

ppb (mg/m3)
Minimum

0.5 (0.003)
0.12 (0.0081)
0.10 (0.067)
0.15 (10 x 10"«)

trace
0
0

0
1.1
0.11

0

0

0
0

Average

-
-
0.32 (0.0022)


30.9 (0.210)
0.034
0.020
0.027
1.8 ± 6.2
3.5 ± 2.3
2.8 ± 3.1

2.0 ± 3.1

2.8 ± 2.7

0.068 + 0.090
1.3 ± 3.1

Reference

Lillian et al. , 1976
Lillian and Singh, 1974
Lillian et al . , 1975
Lillian et al . , 1975

Pellizari et al . , 1979
Bozzelli et al . , 1980
Pellizzari and Bunch, 1979
Pellizzari and Bunch, 1979
Bozzelli et al. , 1980
Pellizzari et al . , 1979
Pellizzari et al . , 1979; Pelliz-
zari, 1978; Bunn et al., 1975
Bozzelli et al. , 1980;
Pellizzari, 1979
Pellizzari et al . , 1979;
Pellizzari, 1977
Bozzelli and Kebbekus, 1979
Bozzelli and Kebbekus, 1979;
Bozzelli et al . , 1980;
                                                         Pellizzari,  1977

-------
TABLE 3-2.   (continued)





CO
1
I—1
CO









Location
New Jersey
Rahway8
Rutherford9
Somerset3
South Amboy
New York
New York City
Niagara Falls and
Buffalo
Whiteface Mtn.
Ohio9
Wilmington
Texas
Houston
Aldinea
Deer Park8
El Pasoa
Freeport8
Houston3
Date r'
Measurement

SepL. 20-22, 1978
May 1, 1978-Dec 29, 1979
July 18-26, 1978
Jan. 27-Dec. 29, 1979
June 27-28, 1974
Aug 18-27, 1978
Fall 1978
Sept. 17, 1974

July 16-26, 1974
Sept 16-25, 1978b
June 22-Oct. 20, 1977
July 29-30, 1976
April 5-May 1, 1978
Aug. 9, 1976
July 27, 1976-
May 24, 1980
Reported Concentration,
Maximum

5.0
9.2
0.068
2.2
9.75 (0.0661)
10.61 (0.0721)
2.0 (0.014)
0.19 (12.8 x 10~")


4.52 (0.030)
0.037
0.15
0.39

1.3
ppb (mg/ma)
Minimum

2.7
0
0
0
1.0 (0.006)
0.16 (0.001)
0,02 (0.122 x 1Q~3)
0.02 (0.13 x 10~3)


<0.1 (<0.001)
0
0.01
0.11

0
Average

3.8 + 1.2
0.89 ± 0.14
0.036 ± 0.26
0.21 ± 0.53
4.5 (0.030)
1.00 (0.006)
1.0 (0.0068)


0.15 ± 0.015
0.11 (0.001)
0.012
0.07
0.15
0.12
0.33
Reference

Bozzelli and Kebbekus, 1979;
Bozzelli et al. , 1980
Bozzelli et al. , 1980
Bozzelli and Kebbekus, 1979
Bozzelli et al . , 1980
Lillian et al . , 1975
Evans et al. , 1979
Pellizzari et al. , 1979b
Lillian et al. , 1975

Lillian et al . , 1975
Evans et al. , 1979
Pellizzari et al. , 1979
Pellizzari et al. , 1979
Pellizzari, 1979
Pellizzari et al . , 1979
Pellizzari et al. , 1979;
Singh et al. 1980

-------
TABLE 3-2.   (continued)




CO
-pi


Location
Texas
LaPorte3
Pasadena3
Utah3
Magna
Washington3
Auburn
West Virginia3
Charleston
St. Albans
aOata obtained
24-hour value.
Sampling time:
Sampling time:
Date °f Reported Concentration, ppb (mg/m3)
Measurement Maximum Minimum Average Reference
Aug. 12-13, 1976 0.49 Pellizzari et al., 1979
July 28, 1976 0.003 Pellizzari et al . , 1979

Oct. 24-Nov. 3, 1977 0.012 0 0.004 ± 0.006 Pellizzari, 1979

Jan. 10-11, 1977 0.76 0.18 0.59 ± 0.27 Battelle, 1977
Sept. 27-Nov. 20, 1977 0.16 0 0.004 ±0.008 Pellizzari, 1978
Sept. 27-Oct. 25, 1977 0.064 0 0.016 ±0.032 Pellizzari, 1978
from summary report of Brodzinsky and Singh, 1982. Values reported are 24-hour sampling concentrations.
14 minutes.
100 minutes.

-------
PCE was 0.026 ±  0.007.7 ppb  (1.7  x  10  4  ±  0.47  x  10"4  mg/m3);  the  coefficient
of variation was  27  percent.  The average urban  level of PCE was found to be
about 0.8 ppb.
     Evidence for  considerable  variability in ambient air  levels  of  PCE was
shown by Lillian et al.  (1975).   The authors attributed the variability of PCE
to its tropospheric reactivity (reaction with hydroxyl radicals).
     Some of the  highest  air levels  of PCE reported  have  been  associated with
waste disposal  sites.  Pellizzari (1978) reported levels to ranges at sites in
New Jersey that ranged from trace amounts  to a maximum of 58 ppb (0.394 mg/m3)
in a  14-minute sampling  period.  PCE  was  adsorbed using Tenax  cartridges.
However, Tenax is reported to generate artifacts  (Singh, 1982).
     Coefficients of variation for most of the recent studies reported by Singh
et al. have been less than 30 percent.
     Howie (1980)  reported ambient  air levels of  PCE  in the vicinity  of laun-
dries to  be  as  high  as 32 ppb (0.22  mg/m3).   In this  study  of  indoor  PCE con-
centrations,  outdoor samples provided background  data.  Measurements were made
by adsorbing PCE onto charcoal filters, followed  by desorption with carbon di-
sulfide and  quantification by GC-ECD and GC-MS.   Outdoor  samples were collec-
ted for 24 hours.  Of 124 measured samples, 56 had 24-hour levels of less than
1 ppb.  Replicate  sample  analyses were reported  to give an overall precision
of better than 20 percent for both indoor  and outdoor samples.
3.6.1.2  Water—Various studies  have shown that  PCE is found in both natural
and municipal waters.  A  review by  Deinzer et al. (1978) has summarized many
of the findings.  Love and Eilers (1982),  in their review, reported that halo-
genated solvents  such  as  PCE are seldom detected in concentrations greater
than  a  few micrograms per liter  in  surface waters.   In  the  U.S.  Environmental
Protection Agency's STORET System for the  period  from August 1977 to September
1984, the mean concentration of PCE  in all  measured water supplies was 2 ug/L.
Maximum observed  concentrations were about 20 ug/L.   The  information  was com-
piled from 66 measurement stations  in  10 states.  With  respect to  streams  and
other surface waters,  analysis  of data from 1,102 measurement stations in 45
states indicates that, from August 1975 to September 1984, the mean PCE concen-
tration was 1 ug/L.
3.6.1.2.1  Natural waters.   Surface  waters,  such  as  rivers  and  lakes, are  the
most  important  sources of drinking water in  the United  States.   Attempts have
been made to show  an  epidemiological link  between the presence of halogenated
                                    3-15

-------
 organic  compounds in drinking water and cancer (Harris and Epstein, 1976) but
 a cause-effect  relationship  has  not been established.
      Dowty et  al.  (1975a,b)  detected PCE by  GC-MS  techniques in untreated
 Mississippi  River water as well  as in treated water.  An approximate six-fold
 reduction in  concentration occurred after sedimentation and chlorination.  PCE
 in water  from a commercial deionizing charcoal filtering unit showed a marked
 increase over the amount found in finished water from treatment facilities or
 commercial sources of bottled water.   The value of charcoal filtering to remove
 organics from water requires  further study.
      Suffet et al. (1977) reported detection of PCE in river waters supplying
 drinking water to Philadelphia,  Pennsylvania.   The Belmont Water Treatment Plant,
 with an  average capacity of 78 million gallons per day, obtains influent from
 the Schuykill River.
      In  a study  designed  to  detect pollutants in surface water at different
 U.S. sites,  Ewing et al.  (1977)  identified PCE among the  pollutants.   Detection
 limits were  not  reported.  The  highest  level  reported was 45 (jg/L.   In all
 samples taken in California,  Oregon,  and Washington,  PCE  was  either not detec-
 ted or was found at a concentration of 1 pg/L  or less.   Sampling sites  included
 those in the vicinity of Los  Angeles Harbor, Santa Monica Bay, and San  Francisco
 Bay, at three sites along the Willamette River in  Oregon,, and two in the Puget
 Sound area.
      PCE was  among a number  of halogenated organics  found in  community  drinking
 water supply wells in Nassau  County,  New York.  Because  of contamination,  16
 of these wells were  closed  by the New York State  Health  Department (Ewing et
 al.,  1977).   The maximum  detected level  of PCE in the  contaminated wells was
 375  (jg/L.  Since  PCE is generally  not  used  as a cesspool cleaning  agent,
 previous  industrial  dumping may  be the source  of contamination.
      Pearson  and  McConnell (1975)  found  an  average  PCE concentration of 0.12
 ppb  in  Liverpool  Bay sea  water;  the  maximum concentration found was  2.6 ppb.
 Sediments  from  Liverpool  Bay were  found  to  contain  4.8 ppb (w/w).   No  direct
 correlation was found between  PCE  concentration in sediments  and in  the waters
 above.  Rainwater collected near an  organochlorine manufacturing site was found
 to contain 0.15 ppb (w/w)  PCE  (Pearson and McConnell, 1975);  it  was  not detec-
ted in well waters.  Upland waters  of  two  rivers in Wales  were found to contain
approximately 0.15  ppb PCE;  similar  levels  of trichloroethylene were  found
(Pearson and McConnell, 1975).
                                    3-16

-------
     Lochner (1976)  found  that levels of PCE  in  Bavarian  lake  waters  ranged
from 0.015 to 3 ppb (0.015 x 10~3 to 2.7 x 10~3 mg/L).  European surface waters
were reported  to  have  uniform PCE concentrations  ranging  from  0.2  x 10"3  to
0.002 mg/L.  Analyses  of river,  canal,  and  sea water,  all  containing effluent
from production and user sites in four countries, revealed  PCE concentrations
ranging from 0.01  to 46 ppb (0.01 to  46 ug/liter) (Correia et a!., 1977).
3.6.1.2.2  Municipal waters.   Bellar  et al.  (1974)  measured the concentration
of PCE in water obtained from sewage treatment plants in several cities.  Before
treatment, the average PCE concentration was 6.2 |jg/L.  The treated water before
chlorination contained 3.9 |jg/L PCE.   After chlorination, the effluent contained
4.2 [jg/L PCE.
     PCE has been  detected in the drinking water of a number of U.S.  cities.
These  include  Evansville,  Indiana (Keith et al.,  1977);  Kirkwood,  Missouri
(Keith et  al. ,  1977);  New Orleans, Louisiana (Dowty et al.,  1975);  Jefferson
Parish, Louisiana (Dowty et al., 1975b); Cincinnati, Ohio (Keith et al., 1977);
Miami, Florida  (Keith  et al.,  1977);  Grand  Forks,  North  Dakota  (Keith et al.,
1977); Lawrence, Kansas (Keith et al., 1977); New York City (Keith et al.,  1977);
and Tucson, Arizona (Keith et al. , 1977).
     Concentrations  recorded  for  the  above cities were less  than 1 pg/L.   An
exception was Jefferson Parish, which had a measured concentration of 5 ppb (5
ug/L).  Keith et al. (1977) did not detect PCE in the drinking water of Phila-
delphia.   PCE was found in Evansville tap water from July 1971 to December 1972.
The Ohio River Basin, a heavily industralized area, is upstream from Evansville
and serves as a major source of drinking water for that community.
     Dowty et  al.  (1975b)  determined levels of PCE in the drinking water for
New Orleans.    Considerable variation in the relative  concentrations  of the
various halogenated compounds was observed from day to day.
     Contamination of drinking water by PCE was recently investigated by Wake-
ham et al.  (1980).   It was reported that elevated concentrations of PCE were
found  in  drinking  water transported in vinyl-coated asbestos-cement pipes in
areas of the town of Falmouth, Massachusetts.  PCE is used as a solvent during
the application of the vinyl coating to the pipe during manufacturing.   It was
suggested that residual solvent leaches into the water carried in these pipes.
                                    3-17

-------
      Using  a charcoal trap with  flame  ionization detection, Wakeham and co-
 workers  (1980)  detected  PCE  levels  ranging  from 140  to 18,000 ppb in unflushed
 pipes.   In  other  parts of  the  distribution  system,  levels were less than 2 ppb.
 The authors reported that vinyl-coated  asbestos-cement pipe has been used in
 parts of the northeastern United States over the past decade in response to
 concerns that water carried in  uncoated pipes  could contain asbestos fibers.
      In  municipal waters  supplying the  cities of  Liverpool,  Chester,  and
 Manchester, England,  0.38  ppm  (w/w)  PCE  was found (Pearson and McConnell,  1975).
      Munich (Germany) drinking water was analyzed by Lochner  (1976).  Samples
 taken at various  sampling  points  and times  gave a range of 1.1 x 10 3 to 2.4 x
 10"3 mg/L.   Raw sewage at  Munich  contained  0.088 mg/L PCE.   On mechanical  clari-
 fication,  the 24-hour average  concentration of  PCE was 0.0068 mg/L.
 3.6.1.3   Sediments--Sediment  levels  for  PCE documented in the U.S.  Environmental
 Protection Agency's  STORE! system indicate  that mean levels from 1,102 measure-
 ment stations  in 45 states are about 33 M9/kg (dry weight).  Maximum reported
 levels have been  as  high as  3,700 ug/kg.  Measurement period covers June,  1977
 to November, 1984.

 3.7  ANALYTICAL METHODOLOGY
      PCE has been analyzed in  air and  in water,  as well  as in biological fluids,
 by a variety of  methods.   Separation  of PCE from other compounds is usually
 carried  out by  gas  chromatography (GC).   Quantification is usually made by either
 electron capture  detection (ECD) or mass spectroscopy (MS).  These analytical
 methods, GC/ECD or GC/MS,  have a  lower  limit of detection of a few ppt.

 3.7.1 Ambient  Air
      Because PCE  levels  in air are  typically in the  sub-ppb range,  the sampling
 and  analysis  techniques  have been designed  to detect trace gas levels.
 3.7.1.1  Sampling and Sources  of  Error—Because of the low levels occurring  in
 ambient  air,  sampling techniques have focused  on adsorption onto solids such
 as charcoal  (Evans et al., 1979)  or  on concentration methods that increase  the
 amount of PCE to above detection  limits  (Rasmussen et al. ,  1977).   In the  upper
 troposphere,  PCE has been  sampled by pumping air into stainless  steel  or glass
containers  until there is a positive pressure relative  to the  surrounding  atmos-
phere (Singh et al., 1979).
                                    3-18

-------
     Evans et  al.  (1979)  sampled PCE using a method based on adsorption onto
activated charcoal,  followed  by  desorption  by  carbon  disulfide/methanol.   The
precision of the  analytical  method, expressed as  a coefficient  of  variation
for the  total  measurement system (including sample collection, handling, and
preparation) was reported as 16 percent.  When 49 quality control samples were
analyzed, the  overall  percent recovery  from the  charcoal  tubes was  70.2  ±1.7
percent.  When the total measurement system was  independently checked by using
Tenax  GC,  another solid adsorbent, the  paired  data  were correlated with  a
coefficient of 0.82, with the average Tenax result exceeding the average char-
coal result  by 21  percent.  Samples were in the  sub-  to  low-ppb  range.   Evans
et  al.  (1979)  reported that PCE is stable  on charcoal tubes for at least one
month  at  0°C.   The lower limit of  detection for the total measurement method
(to include GC/ECD) was estimated at 0.68 p.g/m3  (0.1 ppb).
     Pellizzari and  Bunch  (1979)  reported the  use  of  Tenax GC, a porous  poly-
mer based  on  2,6-diphenyl-p-phenylene  oxide,  to adsorb PCE 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
0.38 ppt (2.5 x 10 6 mg/m3).  Accuracy of analysis was reported at ± 30 percent
Included among the inherent analytical  errors were (1) the ability to accurate-
ly  determine the  breakthrough volume,  (2)  the percent recovery from the sam-
pling  cartridge after a period of storage,  and (3) the reproducibi1ity of ther-
mal 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, protec-
ted from  light,  and  stored at 0°C. The  advantages reported for Tenax include
(1) low water  retention,  (2)  high  thermal  stability,  and (3)  low background
levels  (Pellizzari,  1974,  1975,  1977; Pellizzari  et al.,  1976).  Singh et  al.
(1982)  have  cautioned  that Tenax  suffers  from  serious  artifact problems.
     Krost et  al.  (1982)  reported  an estimated detection  limit of 0.3 ppt  for
PCE using  high resolution GC/MS.    The detection  limit was calculated on the
basis  of  the  breakthrough volume for a  known amount of Tenax GC at 10°, 21°,
and 32°C.  Field sampling and analysis  precision of the Tenax method was found
to  range  from  ± 10 to ± 40 percent relative standard  deviation for different
substances when replicate field sampling cartridges were  examined.
     Knoll et al.  (1979) reported resolution of  PCE from  other chlorinated hy-
drocarbons with Carbopak C-HT, a graphitized thermal  carbon black treated with
                                    3-19

-------
 hydrogen  at  1000°C.   Carbowax 20M,  reacted  with nitroterephthal ic acid, was
 reported  not to give good  separation.   Porapak T, a porous polymer based on
 ethylene  glycol  dimethacrylate, was  reported to give  good  separation.
      A  freezeout concentration has  been  developed by Rasmussen and coworkers
 (1977)  to determine trace  levels  of PCE in the presence of other compounds.
 The detection limit was reported  at 0.2 ppt (1.36 x 10 6  mg/m3)  for 500-mL
 aliquots  of  ambient  air  samples measured  by  GC  coupled  with EDC.   When freeze-
 out is  complete,  PCE remains  behind,  and  such gases as  oxygen  and  nitrogen are
 passed  through as the freezeout  loop is heated.  The carrier gas sweeps the
 contents  onto the column.
      Singh et al. (1979, 1982) have employed the  cryogenic trapping  of air
 containing trace levels of PCE and  other compounds of  interest.   During sam-
 pling,  traps are maintained at liquid oxygen temperature.   Traps were made of
 stainless steel  packed with  a 4-inch bed of glass beads or glass  wool.   Ali-
 quots are thermally desorbed  and  injected directly into the gas chromatograph.
 Both electric heating and  hot water desorption techniques were found to be
 satisfactory.
      Makide  et al.  (1980)  employed  stainless steel  canisters  for  sampling of
 air containing PCE at levels  of about 20  ppt.   Canisters were  polished electro-
 chemically.   Canisters were evaluated to  10 4  Pa at 200°C  before  sampling.
 The composition of the  samples was reported  to  have remained unchanged for over
 a year.
      Budde and  Eichelberger  (1979)  reported that  carbon  adsorption  methods
 generally have more disadvantages than  those methods using porous polymers.
 The advantage of porous polymers  coupled with thermal desorption, as  contras-
 ted with  solvent desorption,  is  higher  sensitivity,  because the total sample
 is  measured  and  there  is no background from  the solvent.   However,  because the
 total sample  is measured, multiple samples must be collected to insure against
 accident  and  loss of  sample,  and  to  obtain  information  on  the  precision  of the
 method.    Tenax  GC was reported to be superior  to other polymers for  organics
 analysis.   Samples  are taken  by  pulling  air through glass tubes packed with
 Tenax GC,  60/80 mesh and supported  by plugs of glass wool.  After  a  suitable
 sampling  period  (about  2 to 4 hours in  urban  areas),  tubes  are capped and
 stored.    Samples  are thermally desorbed  (250 to 270°C)  for 3  minutes  under  a
10-mL helium flow.
                                    3-20

-------
     Criteria for  evaluating  methods  using solid sorbents to collect organic
compounds from air have  been discussed by  Melcher  et  al.  (1978).   Among  the
factors to be considered are effects of (1) size of collection tube (2) break-
through volume,  (3)  humidity,  (4) temperature, (5) migration, (6) desorption
efficiency,  and (7) concentration.
3.7.1.2  Analysis—The sampling methods which use solid adsorbents or cryogenic
techniques have  the  trap connected to the gas chromatograph by multiple-port
gas sampling  valves.   With solid traps, the  collected organics  are quickly
heated  and  the desorbed organics  are  passed  through  capillary columns.   A
number  of coating  materials in the capillary  columns  have been successfully
used for  separating  PCE.   These materials  include  (1) SF-96  on 100/120 mesh
Chromosorb W  (Cronn et al.,1977); (2) SP-2100 on 80/100 mesh Supelcoport (Singh
et al., 1979); (3) 80/100 mesh Carbopak C-HT,  Porapak T, and SP-2100/0.1 percent
Carbowax 1500 on 100/ 120 mesh Supelcoport (Knoll et al.,  1979).   In the method
used by  Evans et al.  (1979)  in  field  studies,  a 1.8-m glass column with  a
2-mm i.d., packed with 0.1 percent SP-1000 on Carbopack C, 80/100 mesh was used
to separate  PCE  from  other organics  in  ambient air  samples.  Twenty-four-hour
samples were  adsorbed onto charcoal.  After desorption with carbon  disulfide/
methanol, a 1.0-uL aliquot was injected into the gas chromatograph.   The separa-
tion conditions  included an oven temperature of 125°C  (all transfer lines at
least 170°C), and the carrier gas was 5 percent methane in argon.   Quantifica-
tion was  made by ECD (Nickel 63;  ECD  temperature  of  218°C) and  a  standing
current of 0.5  ampere.   To insure that the cell  was  not contaminated, the
sensitivity  of the detector was evaluated by  comparing the standing current
with the pulse frequency curve.
     Electron capture detection is a method of choice  used by a number of inves-
tigators (Singh et al., 1977, 1979, 1982;  Rasmussen et  al., 1977).  Singh et al.
(1979) maintained the ECD at a higher temperature (325°C) than did Evans et al.
(1979), because  it was found  that  the ECD  response  increased with an increase
in temperature.  The  identity of PCE was confirmed by  determining its ionization
efficiency as well as the EC thermal response.
     More recently, Singh  et al.  (1982) maintained  the ECD  at 275°C with a
carrier flow  rate  of  40  mL/min to analyze  for  PCE and 11 other  halogenated
organics  in  ambient air.   Identity  of PCE  and other compounds was established
from retention times  on  multiple columns, the ECD  thermal response, and  the
ECD ionization efficiency.   Lillian and Singh (1974) reported that the accuracy
associated with  GC-ECD measurements  is  75 percent  or  greater with  compounds
                                    3-21

-------
having ionization efficiencies exceeding 50 percent.   Using two  ECDs  in series,
PCE was found to have an ionization efficiency of 70  percent.   In a comparison
of GC-ECD with GC-MS, Cronn et al.  (1976) judged GC-ECD to be superior in re-
producibility for quantitating halocarbons.   Of four  halocarbon standards (PCE
not among them)  measured by GC-ECD, the coefficients  of variation  ranged from
1.4 to 4.3 percent,  compared  to a  range of 4  to  19 percent when  11 halocarbon
standards were measured by GC-MS.   A close agreement  between the levels of PCE
and other halocarbons  determined  by GC-ECD and GC-MS on the same ambient air
samples was obtained by Russell and Shadoff (1977).
     GC-ECD was used by Pellizzari  et al.  (1979) to measure PCE  in ambient air
samples.   Samples were  adsorbed  onto charcoal and desorbed with a mixture of
methanol  and  carbon  disulfide, and aliquots were separated on  a 2.5-mm (i.d.)
Pyrex  column  containing 0.2 percent Carbowax 1500 on Carbopack  C.  The esti-
mated detection  limit was 2.5 x 10~6 mg/m3 (0.38 ppt).
     Makide et al. (1980) separated trace  levels of  PCE from other halogenated
organics on a silicone  OV-101 column (10 percent by  weight  coated  on Chromosorb
W-HP, 80-100 mesh) of 5-mm  i.d. and 3 m long.   Samples were transferred to the
column cooled at -40°C  during preconcentration.   Separation was  carried out by
raising  column  temperature  5°C per minute up to 70°C.   Methane was  added to
the carrier gas  (nitrogen)  to improve the  signal-to-noise  ratio  and to stabilize
the baseline.  Quantification was  made  by  a  constant-current ECD.  The detection
limit  for PCE was reported  as <0.05 ppt,  a level unattainable under nearly all
conditions.   Precision  was  reported to  be  within 2 percent.
     To  measure  PCE  at levels expected to occur in  occupational  air, flame
ionization detection has  been used.   The analytical  method S335, suggested by
the National  Institute for Occupational  Safety and  Health  (1977)  for  organic
solvents in air, utilizes adsorption onto  charcoal,  followed by  desorption with
carbon disulfide (CS2).  PCE is separated by GC.   The method  is recommended
for the  range 96 to  405 ppm (655 to 2749 mg/m3).  The coefficient  of variation
for the  analytical sampling method is 5.2  percent.   With the method, interfer-
ences are minimal, and  those  that  do occur can  be eliminated by  altering chroma-
tographic conditions.

3.7.2  Water
3.7.2.1  Sampling--A variety  of  techniques and  methods are commonly used to
sample trace  levels  of  PCE  and other halogenated organics  in water samples.
Coleman  et al. (1981) have  reported that the  Grob closed-loop-stripping tech-
nique  is an excellent tool  to monitor organics  in water at the ppt level   It
                                    3-22

-------
was reported  that  a million-fold concentration of  most  low and  intermediate
molecular weight organics can be achieved.  Quantisation is performed by spik-
ing the initial water sample with a series of  internal standards, stripping at
30°C for  two  hours, and by chromatographing the CS2 extract on a wall-coated
open-tubular capillary.
3.7.2.1.1  Gas purging and trapping.  In this method, finely divided gas bubbles
are passed  through  the  sample, transferring the organic compounds to the gas
phase.   The gas is  then passed through a solid adsorbent in a trap.   Compounds
are desorbed  at  elevated temperature by backflushing with a carrier gas into
the gas chromatograph (Budde and Eichelberger, 1979).  Since the boiling point
of PCE is 121°C, Tenax GC would be an effective absorbent.
     The  purge  and  trap procedure is widely  used,  as  is the purging device
developed by Bellar and Lichtenberg (1974).  For most organic compounds, detec-
tion limits as  low  as 1 ug/L  can  be  obtained when GC/MS  is  used for  analysis.
Mieure (1980) reported that adding salt to the sample or increasing its tempera-
ture dramatically improves the removal of most organic compounds.
     Otson  and  Williams  (1982)  have  described  a modified purge and trap tech-
nique  for evaluation  of volatile organic pollutants in water.   The detection
limit  reported for  PCE was 0.1 ug/L with ECD, and 1 ug/L with flame ionization
detection (FID).  Tenax GC was used as packing for the combined trap/chromatog-
raphic column.
3.7.2.1.2   Headspace analysis.  This  method  describes static sampling of the
vapor  phase that is in equilibrium with the aqueous sample.   The concentration
in the headspace  is proportional  to  the concentration  in the water (Kepner  et
al., 1964).   In  this procedure, trace organics  in the  range  of 10 to 100 ug/L
can be sampled  (Mieure,  1980).   Mieure (1980) reported a detection limit for
PCE (analyzed by ECD) of 0.01 ug/L.   With flame ionization detection, the limit
was 32 ug/L.   Typically,  a 1- to 2-mL sample of the headspace is removed and
injected  into  the gas chromatograph.  Headspace extraction  coupled with mixed
column separation and  ECD analysis  was reported by Caste!lo et al.  (1982) to
be suitable for rapid screening of drinking water supplies.
3.7.2.1.3   Liquid/liquid extraction.  Mieure  (1980)  reported that recovery  of
PCE from water spiked with 2.3 to 90 ug/L ranges from 100 to 113 percent.   The
precision ranges  from 10 to 12 (RSD).  These  results  were  obtained from a
round-robin study,  by the American  Society for Testing  and  Materials  (ASTM)
Committee D19  on  Water,  using liquid/liquid extraction.  The extractant was
not identified.
                                    3-23

-------
     Budde and Eichelberger (1979) cautioned that a disadvantage to this method
 is  that  very volatile compounds may  be  lost  during  extract concentration  or
 during  solvent elution from the gas  chromatograph.   Methylene chloride was
 recommended  as the only general-purpose solvent.
     Sheldon  and  Hites (1978)  used methylene  chloride in a sampling procedure
 applied  to  the identification of  PCE and 98 other organic  compounds  in river
 water.   Grab samples were collected  in amber  glass  bottles and samples for
 solvent  extraction were  immediately  preserved by acidifying  to pH  2 with
 hydrochloric  acid and by  adding 250 ml  of methylene  chloride.   The analytical
 techniques used were  those reported by Jungclaus  et  al.  (1978).  Solvent extrac-
 tion efficiencies were not determined.   While  PCE was  previously  detected in
 vapor  stripping  analysis  of prior  samples,  it was not detected in the water
 samples  cited  in  their report.
 3.7.2.2   Analysis—Schwarzenbach et al.  (1979) used ECD to measure PCE levels
 in  water samples.  Volatile organics were  purged and adsorbed onto charcoal.
 Desorption was by CS2 Quantification  was  made  by  FID and ECD.
     Dowty  et al. (1975a) used  Tenax GC in trapping purgeable organics  from
 water  samples.   The polymer containing the trapped organics was placed in  the
 GC  injection port maintained  at 200°C.   Final  separation was  made on a glass
 capillary  column  coated with  Pluronics  121.   The effluent of the column was
 split to allow for FID and  ECD.

 3.7.3  Biological  Media
     Ramsey  and  Flanagan (1982) have described a gas chromatographic method
 reported to  be suitable for  analysis  of  PCE and other organics present in blood.
 Detection was both by flame  ionization  and  ECD.   Approximately 200 uL of blood
 or  200 mg tissue  is  required  for analysis.

 3.7.4  Calibration
     Singh et  al.  (1982)  found  that  primary  standards of PCE in the low-ppb
 range  could  be satisfactorily calibrated  using permeation tubes  maintained
 either at 30° or 70°C.  Permeation tubes were standard  FEP  or  TFE  Teflon.   All
 permeation tubes were  conditioned  for two weeks or longer.  Errors  in the per-
meation  rate were ± 10 percent.
                                    3-24

-------
3.7.5  Storage and Stability of PCE
     Sampling of  exhaled breath commonly  is  accomplished  by  use of  Saran bags
or glass pipettes.  Temperature and storage time of the samples before analysis
are factors to be considered in obtaining accurate data.
3.7.5.1  Glass Sampling Tubes--Eva1uation  of glass  sampling  tubes was  made  by
Pasquini (1978).  Serial  alveolar breath samples were collected in the tubes
and the  concentrations  of  PCE were analyzed  by  a  gas  chromatograph equipped
with a  flame  ionization detector.  Analysis  of vapor  retention over 169 hours
indicated  that  glass  tubes  can be acceptable containers  for  breath  samples  if
precautions are  taken.  Moisture,  temperature, and  tube  surface and condition
can greatly alter vapor retention.
     In  tubes  filled  with  breath samples  taken  at  room temperature and also
stored at  room temperature, the mean percent  loss of PCE was 64.8 ± 9.4.   Par-
titioning  of PCE  between the vapor and liquid states appears to be  a reasonable
explanation for  vapor retention loss.  It was shown for trichloroethylene that
if storage tubes  were maintained  at 37°C, vapor  retention was greater.  It was
also greater if  si 1 iconized tubes were used.
3.7.5.2  Saran,  Teflon and Tedlar Containers--Saran bags as  storage containers
for  PCE  vapors have been evaluated by Desbaumes and Imhoff  (1971).   Although
it was concluded  that Saran can be an acceptable container,  the diffusion rate
was  appreciable  over  a 24-hour storage  period.   Storage temperature  was  not
reported.
     Teflon  containers  were judged by Drasche et al.  (1972) to be  more suit-
able  than  Saran even  though losses of PCE due to adherence to  Teflon  surfaces
were appreciable.  Within the first 30 minutes after  introduction of a mixture
(relative  humidity =  45 percent)  of benzene,  trichloroethylene, and PCE into a
Teflon  bag,  vapor concentrations of  each  dropped 40  to  60 percent.  However,
when  the bag was heated to  100°C for 30 minutes  after  the  mixture had been
stored  for 44  hours at 25°C,  concentrations  rose to the  initial values.
      Knoll  et  al. (1979) reported that  PCE, when stored at ambient tempera-
tures  for  10 days or  less  in Tedlar bags, was stable.   When the vapor mixture
is heated  to 70°C,  PCE  is stable  for  no  more than  5 hours.
                                     3-25

-------
3.8  REFERENCES

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

Bellar, T. A., and J. J.  Lichtenberg.  1974.  Determining volatile organics at
     the microgram per  litre  levels by  gas  chromatography.   J. Am. Water Works
     Assoc. 66:739-744.

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.

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

Bozzelli,  J.  W., and  B.  Kebbekus.  1979.  Analysis of selected volatile organic
     substances  in ambient  air.  New  Jersey Institute  of Technology.
     PB80-144694, Newark, New Jersey.

Bozzelli,  J.  W., B.  Kebbekus, and  A.  Greenberg.  1980.  Analysis of selected
     toxic and carcinogenic substances  in ambient air  in New Jersey.  New
     Jersey  Institute of Technology,  Newark,  New Jersey.

Brodzinsky,  R.,  and  H.  B. Singh. 1982.  Volatile organic chemicals in the
     atmosphere:  an  assessment of available  data.   Final Report.  SRI Inter-
     national, Menlo  Park,  CA.   Prepared for  the U.S.  EPA.

Budde,  W.  L., and J.  W.  Eichelberger. 1979.  An EPA manual for organics analysis
     using gas chromatography--mass spectrometry.

Bunn,  W.  W.,  E.  R. Deane, D.  W.  Klein,  and  R.  D.  Kleopfer.  1975. Sampling and
     characterization of air for organic compounds.  Water Air Soil Pollut.
     4(3-4):367-380.

Calvert,  J.  G. 1976.  Hydrocarbon involvement  in photochemical smog.  Environ.
     Sci.  Technol. 10:256-262.

Campbell,  M.  J., J. C.  Sheppard, and  B.  F.  Aiu.  1979.  Measurement of hydroxyl
     radical  concentrations in  boundary layer air by monitoring CO oxidations.
     Geophys. Res. Lett.  6:185-188.

Castello,  G. , T. C. Gerbino,  and S. Kanitz.  1982.  Gas  chromatographic analysis
     of halocarbons in  drinking water by headspace extraction and mixed column
     separation.   J.  Chromat.   247:263-272.

Chemical and Engineering News.  1984.  June 11.

Coleman, W. E. , R.  G. Melton,  R. W. Slater,  F.  C.  Kopfler,  S. J. Voto, W.  K.
     Allen, and T.  A; Aurand.  1981. Determination of organic contaminants by
     the Grob closed-loop-stripping technique.   J.  Am. Water Works Assoc.
     73(2):119-125.
                                     3-26

-------
Considine, D. M., ed. 1974. Chemical
     New York. NY:  McGraw Hill.
and Process Technology Encyclopedia.
Correia, Y., G. J. Martens, F. H. Van Mensch,  and  B.  P.  Whim.  1977.  The
     occurrence of trichloroethylene, tetrachloroethylene,  and 1,1,1-trichloro-
     ethane in Western Europe  in air and water.  Atmos.  Environ.   11:1113-1116.

Cronn, D. R., R. A. Rasmussen, and  E. Robinson.  1976.  Phase I  report:  measure-
     ment of tropospheric halocarbons by gas  chromatography-mass  spectrometry.
     Washington State University, August.

Cronn, D. R, R. A. Rasmussen,  and E. Robinson.  1977.  Report for Phase  II:
     Measurement of tropospheric halocarbons  by gas  chromatography-mass  spec-
     trometry.  Washington State University,  October.

Cronn, D. R., R. A. Rasmussen, E. Robinson, and D. E.  Harsch.  1977b. Halo-
     genated compound identification and measurement in  the troposphere  and
     lower stratosphere.  J. Geophys. Res. 82(37):5935-5944.

Crutzen, P. J., I. S. A. Isaksen, and J. R. McAfee.  1978.  The  impact of  the
     chlorocarbon  industry on  the ozone  layer.   J. Geophys.  Rev.  83:345-363.

Deinzer, M., F. Schaumburg, and E.  Klein.  1978.  Environmental  Health Sciences
     Center Task Force Review  on halogenated  organics  in drinking water.
     Environ. Health Persp.  24:209-239.

Desbaumes, E., and C. Imhoff.  1971. Use of Saran bags  for the  determination
     of solvent concentration  in the air of workshops.   Staub-Reinhalt Luft
     31(6);36-41.
Dilling, W. L. 1982. In:  Environmental  Risk Analysis  for  Chemicals.
     Nostrand Reinhold Co., N.Y.
                                 Van
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(a):833-838.

Dilling, W. 1977. Interphase transfer processes.   II.   Evaporation  rates  of
     chloromethanes, ethanes, ethylenes, propanes,  propylenes  from  dilute
     aqueous solutions.  Comparisons with theoretical predictions.   Environ.
     Sci. Technol. 11(4):405-409.

Dimitriades, B., B. W. Gay, Jr.,  R. R. Arnts,  and R.  L.  Seila.  1983.  Photo-
     chemical reactivity of perchloroethylene:  a new appraisal.  J.  Air Pollut.
     Control Assoc. 33(6):575-587.

Dowty, B. J., D. R. Carlisle, and J. L.  Laseter.  1975a.  New Orleans  drinking
     water sources tested by gas  chromatography - mass  spectrometry.  Environ.
     Sci. Technol.  9:762-765.

Dowty, B. J., D. R. Carlisle, J.  L. Laseter,  and  J.  Storer.  1975b.  Halogenated
     hydrocarbons in New Orleans  drinking water and blood plasma.   Science
     187:75-77.
                                    3-27

-------
Drasche, H.,  L. Funk, and R. Herbolsheimer.  1972.  Storing of air samples for
     the analysis of contaminants especially,  of chlorinated hydrocarbons.
     Staub-Reinhalt. Luft 32(9):20-25.

Evans, G.  F.,  R. E. Baumgardner, J.  E.  Bumgarner,  P.  L.  Finkelstein, J. E.
     Knoll,  and B.  E. Martin. 1979.  Measurement of perchloroethylene in
     ambient air.   EPA-600/4-69-047.   U.S.  Environmental  Protection Agency,
     August.

Evans, G.  F.  1980.  Measurement of perchloroethylene in ambient air.  Proceed-
     ings  of the International Technical  Conference on Toxic Air Contaminants,
     October 9-10.

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 pollutants
     in  surface waters.  EPA-560/6-77-015.   U.S.  Environmental Protection
     Agency  July.

Fishbein,  L. 1977.  Potential industrial carcinogens and mutagens.
     EPA-560/5-77-005, Office of Toxic Substances,  U.S.  Environmental Protection
     Agency.   May.

Fuller,  B. B.  1976. Air pollution assessment of tetrachloroethylene, Mitre
     Corp.,  February.

Gay, B.  W.,  P.  L.  Hanst, J.  J. Bufalini,  and R.  C.  Noonan.  1976. Atmospheric
     oxidation of  chlorinated ethylenes.   Environ.  Sci.  Techno!. 10(1):
     58-67.

Gonikberg, M.  G. ,  V. M. Zhulin, and  V.  P.  Butuzor.  Bull.  Acad. Sci. USSR  Dir.
     Chem. Sci. 739:1956 (English translation).

Gosselin et  al. 1976. Clinical toxicology of commercial  products.   4th Edition.

Graedel, T.  E.  1978: Bell  Laboratories to H.  B.  Singh,  Personal Communication,
     SRI International, October.

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.

Halogenated  Solvents Industry Alliance. 1983.  Comments submitted to the U.S.
     Environmental  Protection Agency,  August.

Handbook of  Chemistry and Physics. 1976.  57th  Edition, CRC Publishing Co.,
     Cleveland, Ohio.

Hardie, D.  W. F- 1966.  Chlorocarbons and  Chlorohydrocarbons.  In:   The encyclo-
     pedia of chemical  technology, 2nd Edition.

Harris, R.H., and S. S.  Epstein.  1976. Drinking water  and cancer mortality in
     Louisiana.  Science  193:55.
                                    3-28

-------
Howard, C. J. 1976. Rate constants for the  gas-phase  reactions  of OH  radicals
     with ethylene and halogenated ethylene  compounds.   J.  Chem.  Phys.  65(11):
     148-154.

Howie, S. J.  1980. Ambient perchloroethylene  levels  inside  coin-operated
     laundries with dry-cleaning machines on  the  premises.   Environmental
     Monitoring Systems Laboratories, U.S.  Environmental  Protection Agency,
     December.

Huybrechts, G., J. Olgregts, and K. Thomas.  1967.  Trans.  Faraday  Soc.  63:
     1647.

International Fabricare Institute. 1984. Testimony provided to  U.S. Environ-
     mental Protection Agency.

Jensen, S., and R. Rosenberg. 1975. Degradability of  some chlorinated aliphatic
     hydrocarbons  in sea water and sterilized water.   Water Res.  9:659-661.

Jungclaus, L. A.,  V. Lopez-Avila, and R. A.  Hites. 1978.  Environ.  Sci.  Technol.
     12:88.

Keith, L. H., A. W. Garrison, F. R. Allen,  H.  H.  Carter,  T.  L.  Floyd,  J.  D.
     Pope, and A.  D. Thruston, Jr. 1977. Identification  of  organic compounds
     in drinking water from  thirteen U.S. cities:  chap.  22.   In:   Identifica-
     tion and analysis of organic pollutants  in water.   L.  H.  Keith,  ed.   Ann
     Arbor Science.

Kepner, R. E., H.  Maarse, and J. Strating.  1964.  Gas  chromatographic  head
     space techniques for the quantitative  determination of volatile  components
     in multicomponent aqueous solutions.   Anal.  Chem. 36:  77-82.

Knoll, J. E., M. A. Smith, and M. Rodney Midgett.  1979.  Evaluation of emission
     test methods  for halogenated hydrocarbons.   Volume  1.   EPA-600/4-79-025.
     U.S. Environmental Protection Agency,  March.

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

Lillian,  D., and H. B. Singh. 1974. Absolute  determination  of  atmospheric
     halocarbons by gas phase coulometry.   Anal.  Chem. 46(8):1060-1063.

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.
     Atmospheric fates of halogenated compounds.   Environ.  Sci.  Technol.
     9(12):1042-1048.

Lillian,  D., H. B. Singh, and A. Appleby. 1976. Gas  chromatographic analysis
     of ambient halogenated  compounds.  J.  Air Pollut. Control  Assoc.   26(2):
     141-143.

Lochner,  F- 1976.  Perchloroethylene:  taking  stock.   Umwelt 6:434-436.
     (English translation).
                                     3-29

-------
Love, 0. T., and R. G. Eilers. 1982. Treatment  of drinking water containing
     'trichloroethylene and related  industrial solvents.   J.  Am.  Water Works
     Assoc.  74:413-425.

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

Lowenheim,  F. A.,  and M. K. Moran.  1975.  Perchloroethylene.   In:  Faith,
     Keyes,  and Clark's  Industrial  Chemicals, Fourth edition,   pp.  604-611.

MacKay,  D. ,  A. Bobra, D. W. Chan, and W.  Y.  Shiu.  1982.  Vapor pressure corre-
     lations for low-volatility environmental chemicals.   Environ.  Sci.  Technol
     16:645-649.

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

Mathias,  E. , E. Sanhueza, I.  C. Hisatsune, and  J.  Heicklen.  1974.  The chlorine
     atom  sensitized oxidation and  ozonolysis of  C^Cl,.   Can.  J. Chem. 52:
     3852-3862.

McConnell,  G., D.  M. Ferguson, and  C. R.  Pearson.  1975.  Chlorinated hydrocar-
     bons  and the  environment.  Endeavor   34(121):13-18.

Melcher,  R.  G. , R.  R. Langner, and  R. 0.  Kagel. 1978.  Criteria for the evalua-
     tion  of methods for the  collection of organic pollutants in air using
     solid  sorbents.  Am. Ind. Hyg. Assoc. J. 39:349-361.

Mieure,  J.  P.  1980. Determining volatile organics in water.   Environ. Sci.
     Technol. 14(8):930-935.

Monogue, W.  H.  1958. Diss. Abstr.  18:984.

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

National Institute  for Occupational Safety and  Health.  1977.  Manual of analyt-
     ical methods,   2nd Edition, Part II.   NIOSH Monitoring Methods, Vol.  3,
     April.

Ohta, T., M. Morita, and I.  Mizoguchi.  1976.  Local  distribution of chlori-
     nated hydrocarbons  in the ambient  air in Tokyo.   Atmos.   Environ.  10:
     557-560.

Ohta, T., M. Morita, I.  Mizoguchi,  and  T.  Tada. 1977.  Washout effect and
     diurnal variation for chlorinated  hydrocarbons in ambient air.  Atmos.
     Environ. 11:985-987.

Otson,  R., and D.  T. Williams. 1982. Headspace  chromatographic determination
     of water pollutants.  Anal.  Chem.  54:942-946.
                                    3-30

-------
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(2): 56-59.

Pasquini, D.  A. 1978. Evaluation of glass sampling tubes  for  industrial
     breath analysis.  Am. Ind. Hyg. Assoc. J.  39(1):55-62.

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

Pellizzari, E.  D.  1974.  Development of method for carcinogenic  vapor  analysis
     in ambient atmospheres.    EPA-650/2-74-121, July.

Pellizzari, E.  D.  1975.  Development of analytical techniques  for measuring
     ambient atmospheric carcinogenic vapors.  EPA-600/2-75-075, November.

Pellizzari, E.  D.  1977.  The measurement of carcinogenic vapors  in  ambient
     atmospheres.   EPA-600/7-77-055, June.

Pellizzari, E.  D.  1978.  Measurement of carcinogenic vapors  in ambient atmos-
     pheres.    EPA-600/7-78-062, April.

Pellizzari, E.  D.  1979.  Information on the characteristics  of ambient organic
     vapors in areas of high chemical production.  Research Triangle  Institute,
     Research Triangle Park,  NC.  Available from U.S. Environmental Protection
     Agency.

Pellizzari, E.  D.  , and J.  E.  Bunch. 1979a.  Ambient air carcinogenic vapors:
     improved  sampling and analytical techniques and  field  studies.
     EPA-600/2-79-081, May.   U.S.  Environmental Protection  Agency.

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 a Tenax GC cartridge sampler  for  hazardous vapors.
     Anal. Lett.  9:45-63.

Pellizzari, E.  D.  , M. D. Erickson, and R.  A. Zweidinger.  1979b. 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.

Ramsey, J. D.  and R. J.  Flanagan.  1982. Detection and identification  of
     volatile organic compounds in blood by headspace gas chromatography as an
     aid to the diagnosis of solvent abuse.  J. Chromat.  240:423-444.

Rasmussen, R.  A.,  D. E.  Harsch, R. 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(6):579-581.

Russell, J. W., and L. A.  Shadoff. 1977. The sampling and determination of
     halocarbons in ambient air using concentration on porous polymer.  J.
     Chromat.  134:375-384.
                                    3-31

-------
 Schwarzenbach,  R.  P.,  E.  Molnar-Kubica, W. Giger, and S. G. Wakeham.  1979.
      Distribution,  residence  time, and fluxes of tetrachloroethylene  and
      1,4-dichlorobenzene in Lake Zurich, Switzerland.   Environ.  Sci.  Techno!.
      13(11):1367-1373.

 Sheldon, L.  S. ,  and R.  A.  Hites. 1978.  Organic compounds in the  Delaware
      River.   Environ.  Sci.  Technol.  12(10):1188-1194.

 Simmonds, P.  G.,  S.  L.  Kerrin, J.  E.  Lovelock, and F. H. Shair.  1974.  Distri-
      .bution of atmospheric halocarbons in the air over  the Los Angeles  basin.
      Atmos.  Environ.   8:209-216.

 Singh, H. B.  1976.  Phosgene in the ambient air.   Nature 264: (5585):428-429.

 Singh, H. B.  1977.  Atmospheric halocarbons:  evidence in favor of  reduced
      average hydroxyl  radical  concentration in the troposphere.  Geophys.  Res.
      Lett.  4(3):101.

 Singh, H. B.  1978.  Personal communication.

 Singh, H. B.,  D.  Lillian,  and A. Appleby.  1975a.  Anal.  Chem. 47:860-864.

 Singh, H. B.,  D.  Lillian,  A.  Appleby,  and L.  Lobban.  1975b. Atmospheric
      formation of carbon tetrachloride from tetrachloroethylene.   Environ.
      Lett.  10(3):253-256.

 Singh, H. B.,  L.  J.  Salas,  and L.  A.  Cavanagh. 1977b. Distribution, sources,
      and sinks  of atmospheric  halogenated compounds.   J. Air Pollut.  Control
      Assoc.  27(4):332-336.

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

 Singh,  H. B.,  L.  Salas,  H.  Shigeishi,  and A.  Crawford. 1977a.  Urban-nonurban
      relationships  of halocarbons,  SF6, N20,  and  other atmospheric trace
      constituents.  Atmos.  Environ.  11:819-828.

 Singh,  H. B. , L.  J. Salas,  H.  Shigeishi,  and  E.  Scribner.  1978b. Global
      distribution of selected  halocarbons, SF6,  and N20.  Phase  II interim
      report.  SRI International, Menlo  Park,  California, May.

 Singh,  H. B., L. J. Salas,  H.  Shigeishi,  and  A.  H.  Smith.  1978.  Fate  of
      halogenated compounds  in  the  atmosphere.   Interim Report 1977,
      EPA-600/3-78-017.

 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.  Environ-
     mental 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,  CA.


                                    3-32

-------
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.

Singh, H.  B.,  L. J. Salas, R. Stiles, and H.  Shigeishi.  1982.  Measurements  of
     hazardous organic chemicals in the ambient atmosphere,  final  report.
     Prepared for the U.S. EPA by SRI International,  Menlo Park,  CA,  August.

Snelson, A., R.  Butler, and F. Jarke. 1978. Study  of  removal processes  for
     halogenated air pollutants.  EPA-600/3-78-058.   Environmental  Sciences
     Research Laboratory, U.S. Environmental  Protection  Agency,  Research
     Triangle Park, N.  C.

SRI International Chemical Economics Handbook,  C-2 Chlorinated Solvents.
     February, 1982.  p. 632, 3001 H.

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.

Suffet, I.  H., L. Brenner, and J. V. Radziul. 1977. GC/MS  identification  of
     trace organic  compounds  in Philadelphia  waters,  chap.  23.   In:  Identifica-
     tion and analysis of organic pollutants  in water, L.  H. Keith,  ed.   Ann
     Arbor Science.

U.S. Environmental  Protection Agency. 1979. Draft  environmental  impact  state-
     ment:   perchloroethylene - background information for proposed standards.
     EPA-450/3-79-029a.  August.

U.S. International  Trade Commission 1983. Synthetic Organic  Chemicals.  United
     States Production and Sales.

Wakeham, S.  G. , A.  C. Davis,  R. T. Witt, B. W.  Tripp, and  N. M.  Frew. 1980.
     Tetrachloroethylene contamination of drinking water by  vinyl-coated
     asbestos-cement pipe.  Bull. Environ. Contain. Toxicol.  25:639-645.

Williamson,  D. G.,  and R. J.  Cvetanovic. 1968.  Rates  of  reaction  of ozone
     with chlorinated and conjugated olefins.   JACS 90:4248.

Wood, P. R., F.  Z.  Parsons, J. DeMarco, H. J. Harween, R.  F. Lang,  I. L.
     Payan,  and M.  C. Ruiz. 1981. Introductory  study  of  the  biodegradation  of
     the chlorinated methane, ethane, and ethane compounds.  Presentation
     before the American Water Works Association,  St. Louis, MO,  June.

Yung, Y. L., M.  B.  McElroy, and S. C. Wofsy.  1975. Atmospheric halocarbons:
     a discussion with emphasis on chloroform.  Geophys. Res.  Lett.   2(9):
     397-399.
                                    3-33

-------
                       4.  ECOSYSTEM CONSIDERATIONS

4.1  EFFECTS ON AQUATIC ORGANISMS AND PLANTS
     Tetrachloroethylene (PCE)  has  been tested for acute aquatic toxicity in
12  species.   The information  presented in  this  chapter presents observed
levels reported  to  result  in adverse  effects under  laboratory  conditions.   It
is  recognized  that  such  parameters  of toxicity  are  not  easily  extrapolated  to
environmental  situations.  Test populations  themselves  may  not be representa-
tive of  the  entire  species,  in which susceptibility of various lifestages to
the test  substance  may vary  considerably.   Guidelines for the utilization of
these data in the development of criteria levels for PCE in water are discussed
elsewhere (U.S. EPA, 1979).
     The  toxicity of  PCE to  fish and other aquatic organisms  has been gauged
principally  by  flow-through  and static  testing methods (Committee on Methor^
for Toxicity Tests  with Aquatic Organisms,  1975).   The flow-through method
exposes  the  organism(s)  continuously  to a constant  concentration  of  PCE  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  concen-
tration  only.   Both types  of tests  are  commonly used  as  initial  indicators  of
the potential of substances to  cause adverse effects.

4.1.1  Effects on Freshwater Species
     Alexander  et  al.  (1978)  used  both flow-through (measured)  and  static
(unmeasured)  methods  to  investigate the acute  toxicity  of  four  chlorinated
solvents,  including PCE,  to  adult  fathead  minnows (Pimephales  promelas).
Studies  were conducted  in accordance with  test  methods described by the
Committee on Methods for Toxicity Tests with Aquatic Organisms (1975).
     The  static and flow-through results for the 96-hour experiments indicated
that PCE  was the most  toxic  of the  solvents  tested.   The  lethal  concentration
(96-hour  LC50)  necessary to  kill 50  percent  of the fathead minnows  in  the
flow-through test was  18.4 mg/L (18.4 ppm); the 95 percent confidence limits
were 14.8  to 21.3 mg/L (14.8 to 21.3  ppm).   In  comparison,  the static  experi-
ments gave  a 96-hour  LC50 of 21.4 mg/L (21.4 ppm); the 95 percent confidence
limits were  16.5 to 26.4 mg/L  (16.5 to  26.4  ppm).   Fish  affected  during  expo-
sure were  transferred  to static freshwater  aquaria at  the  end of exposure.
Only those  fish severely  affected  by high  concentrations did  not  recover.
                                    4-1

-------
     When the  minnows  were exposed to  sublethal  levels for short  exposure
intervals, only reversible effects  were observed.   Endpoints  evaluated were
loss of  equilibrium, melanization,  narcosis, and swollen,  hemorrhaging  gills.
The effective  flow-through concentration  (EC50)  of PCE  that produced one  or
more of these reversible effects was 14.4 mg/L (14.4 ppm).
     The 96-hour LC50,  in a static test with the bluegill (Lepomis macrochirus).
was reported as 12.9 mg/L (12.9 ppm) (U.S.  EPA, 1978,  1980).  The most  sensitive
species tested is the rainbow trout (Salmo gairdneri).   The LC50 determined  by
a  flow-through measured  procedure was  5.28 mg/L (5.28 ppm) (U.S. EPA,  1980).
With embryo-larval  test  procedures, a  chronic  value of  0.840 mg/L  (0.840 ppm)
was obtained by the U.S.  EPA (1980) for the fathead minnow.
     With the  freshwater invertebrate  Daphnia magna,  a 48-hour EC50 value of
17.7 mg/L (17.7  ppm) was obtained  (U.S.  EPA,  1980).   The  midge Tanytarsus
dissimilis was more resistant,  with a  48-hour  LC50 value of 30.84  mg/L  (30-84
ppm) determined under static, measured conditions.

4.1.2  Effects on Aquatic Plants
     As  cited  in  the U.S.  EPA  Ambient  Water Quality Criteria  Document  (U.S.
EPA, 1980),  no adverse effects  on chlorophyll  a or cell  numbers  of  the  fresh-
water alga Selenastrum capricornutum were observed at exposure concentrations
as high as 816 mg/L (816 ppm).
     For  the saltwater  species  Skeletonema costatum a  96-hour EC50  of  about
500 mg/L  (500 ppm) was  determined for effects on chlorophyll a and cell   number.
This alga species is more resistant than the alga Phaeodactylum tricornutum  for
which the EC50 value was determined to be  10.5  mg/L  (10 ppm) (Pearson and
McConnell, 1975).

4.1.3  Effects  on Saltwater Species
     Pearson  and McConnell (1975)  investigated the acute  toxicity  of PCE  on
the dab (Limanda  limanda),  barnacle larvae (Barnacle nauplii), and on unicell-
ular algae (Phaeodactylum  tricornutum).   The LC50 was 5 mg/L  (5 ppm) for the
dab.   The 48-hour LC50  for barnacle larvae was 3.5 mg/L (3.5 ppm).
     Toxicity to  the unicellular alga was assessed by measuring alterations  in
the uptake of  carbon from  atmospheric  carbon  dioxide during photosynthesis.
                                                            14
Uptake  of carbon  dioxide was  measured by the use of sodium-  C-carbonate.  The
EC50  was  10.5 mg/L (10.5 ppm).
                                    4-2

-------
     Data collected  by  the  U.S.  Environmental Protection Agency (1980) indi-
cate that,  for  mysid shrimp (Mysidopsis bahia), the LC50 was 10.2 mg/L (10.2
ppm) in a 96-hour static, unmeasured procedure.  Chronic testing over the life
cycle of the mysid shrimp resulted in a chronic value of 0.450 mg/L (0.45 ppm)
(U.S. EPA, 1980).  The chronic value is 0.044 times the 96-hour LC50.

4.2  BIOCONCENTRATION AND BIOACCUMULATION
     An indicator of the potential for  a  substance  to  result  in cumulative  or
chronic toxic effects in aquatic species is the bioconcentration factor (BCF).
Bioconcentration refers  to  the  increased concentration of a substance within
an  organism  (e.g.,  fish)  relative to the  ambient  water concentration under
steady-state conditions.  As defined by Veith et al. (1979), the bioconcentra-
tion factor  (K,  f)  is a  constant  of  proportionality between the concentration
of  the  chemical  in  fish  (Cf)  and  in  the water  (C ).  This can be more  clearly
expressed as
                         Kl
                                     at steady state.  (4-1)
                    Cw
Bioaccumulation, a  term  often erroneously used in place of bioconcentration,
can be  defined  as that process which  includes  bioconcentration  and  any  uptake
of toxic  substances  through consumption of one organism by another.  The BCF
alone,  however,  may  .not  be the most  useful measure  of  the  overall  fate  of a
substance in water  or of its potential  for producing toxic effects,  for all
chemicals.
      In  the absence of  direct measurement, a  measure commonly  used to  assess
the degree  to  which a compound may be  bioconcentrated  is the octanol-water
partition coefficient.   Estimates  for the  octanol/water partition coefficient
range from  339  to 871 (Neely et  al., 1974;  U.S.  EPA,  1980;  Chiou  et al. ,
1977).  The partition coefficient has been shown to be directly related to
bioconcentration potential  in  fish (Neely et  al. , 1974).   According  to the
American Society for Testing and Materials (ASTM), a log partition coefficient
exceeding a value of three is considered an indication of a high probability
of measurable bioaccumulation in aquatic species (ASTM,  1978).  Compounds that
exhibit a large log  coefficient generally  are  those  with  low  water  solubility
and high  solubility  in organic solvents.   Although a compound may demonstrate

                                    4-3

-------
a high BCF or  log partition coefficient, other environmental  factors  that act
to reduce this potential often exist.  The compound may  be  rapidly  hydrolyzed
or degraded by other  mechanisms.   Measurable uptake  by  the organism may be
precluded if the  tissue depuration rate for the substance is great.
     With regard to PCE,  the  BCF was calculated to be 34 and 49  in two  fish
species  (U.S.   EPA, 1980; Neely et al., 1974).  Neely  et  al.  (1974)  found that
the BCF  for PCE  and  other chemicals was linearly  related to the respective
partition coefficients.  For PCE, the log partition coefficient was 2.88,  and
the BCF, determined in trout (rainbow) muscle, was 39.6 ± 5.5.  The trout were
exposed  to  two undefined  levels  of PCE for an undefined period of  time.   The
extent to which the levels approached the acute LC50  level for this species or
whether  a  steady-state  was  achieved was not  reported.   A direct  measurement
for BCF  of  49 for the bluegill  (whole  body)  is  cited in the water quality
criteria document  (U.S.  EPA,  1980).   The log partition coefficient was  2.53.
The depuration rate was rapid,  with a half-life of less than one day.
     Although these studies suggest that PCE does have bioconcentration  poten-
tial,  the  extent  to  which this potential can  be  manifested  in the form of
adverse  effects  can be  gauged  only  from  the results of toxicological  studies.

4.2.1  Levels of PCE  in Tissues of Aquatic  Species
     Pearson and McConnell (1975) suggested that chronic and  sublethal effects
of  PCE  may  result from exposure to  low  concentrations  of PCE, if  the halo-
carbon can  be  bioaccumulated.   As a first  step in addressing the question of
bioaccumulation,  these  investigators  determined levels of PCE in  a  variety of
invertebrate and vertebrate species (Tables 4-1 and 4-2).
     Among marine  invertebrates,  wet tissue concentrations of PCE  were  found
to range from 1 to 9 ppb.   The highest concentration  (8 to 9 ppb) found  was in
the crab (Cancer pagurus).   Higher levels were found  in marine algae  (13 to 20
ppb).   In tissues of fish, a range of 0.3 to 41 ppb was found.  Concentrations
in the livers  of  three species of  fish  were  found to greatly exceed those
found in the  flesh.   Tissue levels from all  species  are  shown in Table  4-1.
Concentrations reported for fish,  fish-eating birds,  and marine mammals were
for selected tissues  such as fish liver, sea  bird  eggs,  and seal  blubber.   If
the reported tissue concentrations  for birds and  mammals are converted  to  a
whole-body weight basis, concentrations are much lower and closer to  concentra-
tions  measured in seawater,  indicating little or no bioconcentration  and biomag-
nification (U.S.  EPA,  1981).
                                    4-4

-------
               TABLE 4-1.   LEVELS OF PCE  IN TISSUES  OF  MARINE  ORGANISMS,  BIRDS,  AND MAMMALS
Species
Invertebrates
Plankton
Plankton
Ragworm (Nereis diversicolor)
Mussel (Mytilus edulis)


Cockle (Cerastoderma edule)
Oyster (Ostrea edulis)
Whelk (Buccinum undatum)
Slipper limpet (Crej) \dula
fornicata)
Crab (Cancer pagurus)
Shorecrab (Carcinus maenus)
Hermit crab (Eupagurus
bernhardus)
Source
Liverpool Bay
Torbay
Mersey Estuary
Liverpool Bay
Firth of Forth
Thames Estuary
Liverpool Bay
Thames Estuary
Thames Estuary
Thames Estuary
Tees Bay
Liverpool Bay
Firth of Forth
Firth of Forth
Firth of Forth
Thames Estuary
Trichloro-
ethylene
Tissue (ppb by mass
0.05 - 0.4
0.9
Not detected
4 - 11.9
9
8
6 - 11
2
Not detected
9
2.6
10 - 12
15
12
15
5
PCE
on wet tissue)
0.05 - 0.5
2.3
2.9
1.3 - 6.4
9
1
2 - 3
0.5
1
2
2.3
8-9
7
6
15
2
Shrimp (Crangon crangon)
Firth of Forth
16

-------
TABLE 4-1.   (continued)
Species
Starfish (Asterias rubens)
Sunstar (Solaster sp.)
Sea Urchin (Echinus esculentus)
Marine Algae
Enteromorpha compressa
Ulva lactuca
Fucus vesiculosus
Fucus serratus
Fucus spiral is
Fish
Ray (Raja clavata)
Plaice (Pleuronectes platessa)
Flounder (Platyethys flesus)
Dab (Limanda limanda)
Source
Thames Estuary
Thames Estuary
Thames Estuary
Mersey Estuary
Mersey Estuary
Mersey Estuary
Mersey Estuary
Mersey Estuary
Liverpool Bay
Liverpool Bay
Liverpool Bay
Liverpool Bay
Trichloro-
ethylene
Tissue (ppb by mass
5
2
1
19 - 20
23
17 - 18
22
16
flesh 0.8 - 5
liver 5-56
flesh 0.8 - 8
liver 16 - 20
flesh 3
liver 2
flesh 3-5
!•;..»•/. no on
PCE
on wet tissue)
1
2
1
14 - 14.5
22
13 - 20
15
13
0.3 - 8
14 - 41
4 - 8
11 - 28
2
1
1.5-11
n c _ -an

-------
                                           TABLE  4-1.   (continued)
Species
Mackerel (Scomber scombrus)

Dab (Limanda limanda)

Plaice (Pleuronectes platessa)
Sole (Solea solea)

Red gurnard (Aspitrigla
cuculus)
Scad (Trachurus trachurus)
Pout (Trisopterus lus"us)
Spurdog (Squalus acanthi as)
Mackerel (Scomber scombrus)
Clupea sprattus
Cod (Gadus morrhua)


Sea and Freshwater Birds
Gannet (Sula bassana)

Source
Liverpool Bay

Redcar, Yorks
Thames Estuary
Thames Estuary
Thames Estuary

Thames Estuary

Thames Estuary
Thames Estuary
Thames Estuary
Torbay, Devon
Torbay, Devon
Torbay, Devon



Irish Sea

Tissue
flesh
1 iver
flesh
flesh
flesh
flesh
guts
flesh
guts
flesh
flesh
flesh
flesh
flesh
flesh
Air
bladder

1 iver
eggs
Trichloro-
ethylene
(ppb by mass
5
8
4.6
2
3
2
11
11
6
2
2
3
2.1
3.4
0.8
<0.1


4.5 - 6
9-17
PCE
on wet tissue)
1
not detected
5.1
3
3
4
1
1
2
4
2
1
Not detected
1.6
<0.1
3.6


1.5 - 3.2
4.5-26
Shag (Phalacrocerax aristotelis)    Irish Sea
eggs
2.4
                              1.4

-------
                                                   TABLE 4-1.  (continued)
oo
Species
Razorbill (Alca torda)
Kittiwake (Rissa tridactyla)
Swan (Cygnus olor)
Moorhen (Gallinula chloropus)


Mallard (Anas platyrynchos)
Mammals
Grey Seal (Halichoerus grypus

Common Shrew (Sorex araneus)

Source
Irish Sea
North Sea
Frodsham Marsh
Merseyside

Merseyside
Fame Island
Frodsham Marsh
Tissue
eggs
eggs
1 iver
kidney
1 iver
muscle
eggs
eggs
bl ubber
1 iver
-
Trichloro-
ethylene
(ppb by mass
28 - 29
33
2.1
14
6
2.5
6.2 - 7.8
9.8 - 16
2.5 - 7.2
3 - 6.2
2.6 - 7.8
*
PCE
on wet tissue)
32 - 39
25
1.9
6.4
3.1
0.7
1.3 - 2.5
1.9 - 4.5
0.6 - 19
0 - 3.2
1
         Levels for trichloroethylene included for comparative purposes.



        Source:  Pearson and McConnell, 1975.

-------
                    TABLE  4-2.   ACCUMULATION OF PCE BY DABS
Tissue
flesh
1 i ver
flesh
1 i ver
flesh
1 i ver
Period of
Exposure (days)
3-35
3-35
3-35
3 - 35
10
10
Mean Exposure
Concentration (ppm)
0.3
0.3
0.03
0.03
0.2
0.2
Mean
Concentration in
Tissue (ppm)
2.8a (13)
113 (14)
0.16 (9)
7.4b (9)
1.3 (7)
69 (7)
Accumu-
lation
Factor
x 9
x 400
x 5
x 200
x 6
x 350
 Numbers in parentheses are number of specimens analyzed.
-,                                             6
 One fish had a flesh concentration of 29.7/10  and was omitted from calcula-
 tions.
h                                           6
 One fish had flesh concentration of 50.3/10  and was omitted from calculations
Source:   Pearson and McConnell,  1975.

     The average concentration of PCE in seawater taken from Liverpool  Bay,  an
area where  many species  of  organisms were collected, was  0.00012  ppm.   A
comparison of this value with those presented in Table 4-1 indicates an uptake
of as much as 75-fold.   It was the authors' contention that, based on their  ob-
servations, there is little indication that bioaccumulation occurs in the food
chain.
     As shown  in  Table 4-2,  dabs  (Limanda  limanda)  exposed  to 0.3 ppm for 3
to  35  days  were found to  have  a BCF (liver) for  PCE  of  400.   It was not
reported whether  this  period of  exposure approximated  a steady-state for PCE.
After dabs were returned  to clean  seawater,  the  level  of  PCE dropped to 1/100
of  the  original  level  in 4 days  and  to 1/1000  of  the  initial  level after 11
days (Figure 4-1).  The ratio between liver and flesh concentrations is approx-
imately 100  to  1.   The relationship  between  flesh and liver concentrations
in the dab is shown in Figure 4-2.
     Dickson and  Riley (1976)  detected PCE  in  three species of mollusks and
in  five  species of  fish  collected near  Port Erin,  Isle of Man.  Levels of PCE
in  various tissues  are shown  in  Table 4-3.   Relative to the  PCE concentration

                                    4-9

-------
 100
        O LIVER ACCUMULATION
        D LIVER LOSS
        A FLESH
                   EXPOSURE TIME, rlays

Figure 4-1.  Accumulation and loss of  PCE  by dabs.
             Source:   Alexander et al. ,  1978.
                       4-10

-------
           100
            10
          E
          Q.
          0.


          I
          CO
          CJ
          CL
            0.1
           0.01
                              T
                      EXPOSURE LEVELS, ppm

                            O  0.3
                            D

                            A
0.2

0.02
                                                    O
                             10              100


                                PCE IN LIVER, ppm
                                                             1000
Figure 4-2.   Relation between  flesh and  liver  concentration  of PCE in dabs


              Source:  Alexander et al.,  1978.
                                   4-11

-------
    TABLE 4-3.   CONCENTRATION OF PCE AND TRICHLOROETHYLENE IN MOLLUSKS AND
                           FISH NEAR THE ISLE OF MAN
Species
Eel (Conger conger)
brain
gill
gut
1 i ver
muscle
Cod (Gadus morhua)
brain
gill
heart
1 iver
muscle
skeletal tissue
stomach
Coalfish (Pollachius birens)
alimentary canal
brain
PCE
(mg x

6
2
3
43
1

3
3
8
2
-
6
-
TrichloroethyTene
lOVg dry weight tissue)

62
29
29
43
70

56
21
11
66
8
-
7
306
71
     gill
     heart
     1 iver
     muscle

Dogfish (Scyl1iorhinus canicula)

     brain
     gin
     gut
     heart
     1 iver
     muscle
     spleen

Bib (Trisopterus  luscus)

     brain
     gut
     1 iver
     muscle
     skeletal tissue
 6
 2
12
13
 4
 0.3
 70
 40
176
 41
274
479
 41
307
143
187
185
                                   4-12

-------
                            TABLE 4-3.   (continued)
                                         PCE	Trichloroethylene
Species                                    (mg x 106/g dry weight tissue)
Baccinum undatum
     digestive gland                       33                      2
     muscle                                39
Modiolus modiolus
digestive tissue
mantle
muscle
Pecten maximus
gill
mantle
muscle
ovary
testis
63
16
88
40
24
176
56
250
33
detected
-
-
Source:  Dickson and Riley, 1976.

in  seawater,  there  was only a slight enrichment in the tissues (< 25 times).
PCE had one of the  lowest mean bioconcentration factors.

4.3  BEHAVIOR IN WATER AND SOIL
     The potential  of any substance for bioconcentration is influenced by many
factors, including  the rate at which it volatizes and its reactivity.
     In the  laboratory study  by  Oil ling  et  al.  (1975),  the  measured  half-life
of  PCE ranged from  24 to 28 minutes in water.  Factors affecting the evaporation
of  PCE were  surface wind speed, agitation  of  the  water, and water  and  air
temperatures.   Reactivity  of PCE  in water  was measured by exposing  sealed
quartz tubes  containing I  ppm PCE  to  sunlight  for  one year.   At 6  months,  the
level of PCE had declined to 0.52 ppm, and at 1 year, to 0.25 ppm.   Dilling et
al.  (1975)  reported that  the presence of  3 percent NaCl  (as in  seawater)
caused about  a  10 percent decrease in  the  evaporation  after 40 percent had
already evaporated.  The addition  of 500 ppm  clay appeared to increase the
rate of disappearance  to 85 percent soluble  depletion  at  20  minutes.  These
experiments were conducted to simulate the evaporation of PCE under conditions

                                    4-13

-------
more nearly like those found in the environment.   Evaporation  of  PCE  from the
hydrosphere is a rapid process.
     The field  studies of  Zoeteman et al.  (1980)  suggest that PCE is  more
persistent in natural water environments  than  is indicated  by  laboratory mea-
surements.   In a field study of the persistence of a variety of organic  chemi-
cals  in  different  aquatic  environments in  the  Netherlands, Zoeteman et al.
(1980)  estimated  the persistence of  PCE  in river  water  from  3  to 30 days
(half-life).   In lakes and groundwaters,  the half-life was  estimated at  10-fold
higher.   Estimates  were  derived from monitored values  of samples collected
between two sites along the Rhine River,  into which no discharges were expected.
PCE was analyzed by GC-MS.
      Estimates  of  the persistence  of PCE in rivers,  lakes,  and ponds,  by
calculation according  to  Smith  et al. (1980)  are  in  general   agreement  with
the  field  results  of Zoeteman and coworkers (1980).  The half-life of PCE  is
obtained from the expression:

                                  t1/2  =  0.693,                          (4-2)
                                             kc
                                             •v

                  where k  is the volatilization rate constant.

      Using the  data provided  by Smith et  al.  (1980),  the  t1/2 (days) is as
follows:  ponds, 9  to 20;  lakes, < 1  to 20; rivers, < 1 to  20.
      Bouwer et  al.  (1981)  found that PCE and other halogenated organics have
the potential to leach rapidly through soil.  When secondary treated municipal
wastewater containing  from 1  to 10 ng/L  PCE was applied  to soil  columns, at
rates  typical of  high-rate land application systems and  under conditions in
which  volatilization was prevented, PCE was detected in the effluent.  Leaching
of  PCE through  soil was  suggested by Zoeteman  et  al.  (1980)  as  a probable
factor in the contamination of groundwater supplies in the  Netherlands.
     The  potential  for halogenated organics,  including  PCE,   to  contaminate
groundwater supplies via leaching from surface waters was examined by Schwarzeir
bach  and Westall  (1981).   In  batch and column experiments  with various  types
of sorbents and organics designed to  simulate field conditions, these investi-
gators found  that  the partition coefficient for a particular  compound can  be
estimated from  its  octanol/water partition coefficient and from  the  fraction
                                    4-14

-------
of organic carbon  in the sorbent.  A  high  degree of correlation was  found
between the partition coefficient and organic carbon content when the fraction
of organic carbon  was  greater than 0.1 percent.   A  partition  coefficient of
0.56 ± 0.09 was  found  for  PCE,  using  natural  aquifer material  (organic carbon
= 0.15 percent)  from a  field  site in  Switzerland.   It was  concluded  that, for
concentrations typically encountered  in  natural  waters, sorption of PCE and
other organics of comparable 1ipophilicity by aquifer materials is reversible.
The expression

                                   S = kpC,                           (4-3)
                  where S = concentration in solid phase
                         kp = partition coefficient
                         C = concentration in liquid phase

was found satisfactory to describe sorption equilibrium.

4.4  SUMMARY
     The available  data  for PCE  indicate that acute and chronic  toxicity to
freshwater aquatic  life  can occur at  concentrations  around 5280 and  840 pg/L,
respectively.   For  saltwater aquatic  life,  the  acute  and  chronic toxicity
values are 10,200 and 450 M9/L, respectively.
     PCE does  not  appear to biomagnify or  concentrate as it moves up the food
chain.  The available  data suggest that the  bioconcentration potential of PCE
is low, and it appears to be eliminated rapidly from aquatic organisms.
     Contamination  of  groundwater supplies  by  PCE leaching through soil could
be a  concern,  particularly in  situations in  which soils of low organic carbon
content are involved.
                                    4-15

-------
4.5  REFERENCES


   Alexander,  H.  C.,  W.  M.  McCarty, and E. A. Bartlett. 1978. Toxicity  of
        perchloroethylene,  trichloroethylene, 1,1,1-trichloroethane,  and
        methylene chloride  to fathead minnows.  Bull.  Environ.  Contam.  Toxicol.
        20:344-352.

   American Society for Testing and Materials. 1978. Estimating the  hazard  of
        chemical  substances to aquatic life.  J.  Cavins, K.  L. Dickson,  and  A.
        W.  Maki,  eds.,  STP  657.   Committee D-19 on Water.

   Bouwer,  E.  J., P.  L.  McCarty, and J. C. Lance. 1981. Trace organic behavior
        in soil columns during rapid infiltration of secondary  wastewater.
        Water Res.  15(1):151-160.

   Chiou, C.  T.,  V.  H.  Freed, D. W. Schmedding,  and R. L. Kohnert. 1977.
        Partition coefficient and bioaccumulation of selected organic chemicals
        Environ.  Sci.  Techno!.  11:475-478.

   Committee on Methods for Toxicity Tests with Aquatic Organisms. 1975.
        methods for acute toxicity tests with fish, macroinvertebrates,  and
        amphibians.   Ecol.  Res.  Series, EPA 600/3-75-009.

   Dickson, A. G., and J.  P.  Riley. 1976.  The distribution  of short-chain
        halogenated aliphatic hydrocarbons in some marine organisms.  Marine
        Pollut. Bull.  7(9):167-169.

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

   Neely, W.  B.,  D.  R.  Branson, and G. E.  Blau.  1974.  Partition coefficient
        to measure bioconcentration potential of organic chemicals in fish.
        Environ.  Sci.  Technol. 8:1113.

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

   Schwarzenbach, R.  P.  and J. Westall. 1981. Transport of  nonpolar  organic
        compounds from surface water to groundwater.   Laboratory sorption
        studies.   Environ.  Sci.  Technol. 15:1360-1367.

   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. Technol. 14(11):1332-1337.

   U.S.  Environmental  Protection Agency.  1978.  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.  Tetrachloroethylene:  water
        quality criteria.   Federal Register 44(52):15966-15969.

                                    4-16

-------
U.S.  Environmental Protection Agency. 1980. Tetrachloroethylene:  ambient
     water quality criteria.  Office of Water Regulations and Standards,
     EPA  440/5-80-073, October.

U.S.  Environmental Protection Agency. 1981 Environmental risk assessment
     of tetrachloroethylene, draft report.  Office of Toxic Substances.  14
     September.
Veith, G. D., D. L. DeFoe, and B. V.
     ing the bioconcentration factor
     Board Canada 36:1040-0145.
Bergstedt.  1979.  Measuring and estimat-
of chemicals in fish.  J.  Fish. Res.
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.
                                  4-17

-------
                 5.   MAMMALIAN METABOLISM AND PHARMACOKINETICS

     The metabolism and  pharmacokinetics  of PCE are highly contingent on the
physicochemical properties of this  compound.   It  is  a  volatile  liquid  at  room
temperature with a relatively low vapor pressure  (19 torr  at  25°C),  is nearly
insoluble in water but is  highly lipophilic.   Table  5-1  compares the physical
properties of  PCE with other familiar chloro-substituted ethylenes.  It is of
interest to note that  with increasing chloro-substitution, water  solubility
decreases and  lipid  solubility  increases.  These properties  determine  to a
considerable extent the absorption,  distribution,  and routes of elimination of
PCE from the body,  i.e.,  the pharmacokinetics of PCE.  All  chloroethylenes are
thought to share as  common metabolic steps in the body a)  epoxidation  of the
ethylene double bond,  b)  rearrangement to the halogenated acetaldehyde  or acyl
halide, and c) partial transformation  of the latter to  a  halogenated acetv
acid (Bolt et al.,  1982).

5.1  ABSORPTION AND DISTRIBUTION
5.1.1  Dermal  Absorption
     Studies have shown  that  absorption  of PCE through  the skin,  from vapor
exposure or from partial body immersion,  is minimal  in comparison  to oral and
inhalation routes of  exposure.   Riihimaki and Pfaffli (1978)  exposed  human
volunteers (lightly clad  and with respirators to prevent  pulmonary  absorption)
to 600 ppm (4068 mg/m ) PCE for 3.5  hours and^estimated percutaneous absorption
of the  vapor to be about 1 percent  of pulmonary absorption.  Stewart and  Dodd
(1964) and Hake and  Stewart  (1977)  have  also  experimentally  estimated skin
absorption in volunteers  by noting PCE appearance  and concentration in  exhaled
air after immersion of the thumbs in liquid PCE.  The mean peak concentration
                                                                            3
of PCE in exhaled  air 40  minutes after immersion was only 0.31 ppm  (2.1 mg/m  ),
                                     3
and after 2 hours,  0.23 ppm (1.6 mg/m ),  indicating that  the rate of absorption
through the skin is very  slow,  even  after allowing for storage and  metabolism.
Jakobsen et al. (1982) experimentally  quantified  transport of PCE and other
chlorinated hydrocarbons across  guinea  pig skin.   Liquid contact (skin area,
      p
3.1 cm ) was maintained for up to 6  hours and solvent concentration monitored in
                                    5-1

-------
                    TABLE  5-1.   PHYSICAL PROPERTIES OF PCE
                           AND  OTHER CHLOROETHYLENES
                                      Vapor
                                   Pressure at
                                 25°C,  760 torr
             Ostwald Solubility, 37°C
           Water/  Blood/  Olive oil/
            air     air       air
Vinyl  chloride
  (1-chloroethylene)
Vinylidene chloride
  (1,1 dichloroethylene)
Trans,1,2 dichloroethylene
Cis 1,2 dichloroethylene
Trichloroethylene
  (1,1,2 trichloroethylene)
Tetrachloroethylene
  (1,1,2,2 tetrachloroethylene)
gas

600
350
250

 90

 19
2.1
2.9.

1.3
5.8
9.2

9.5
0.43     13.1
  69
 189
 270

 718

1917
Adapted from Sata and Nakajima, 1979.
Conversion factor:  1 ppm vapor in air = 6.78 mg/m3 at 25°C, 760 torr.

blood during and following dermal  applications.   Blood concentrations (reflect-
ing absorption  rate)  increased rapidly,  peaking at 0.5  hour  (1.1  ug/mL)  and
then decreased with time, suggesting that changes in dermal permeability occur
with time.  This pattern was found to be general and characteristic of chlori-
nated hydrocarbon  solvents  like  PCE.   Similar observations have been made by
Tsuruta (1975), who estimated  the percutaneous  absorption  rate  of  PCE through
                             2
mouse skin at  24  nmol/min/cm  of  skin.  Kronevi et al.  (1981)  have  observed
histological  changes in guinea pig derma within 15 minutes of topical exposure
to PCE.    Stewart  and Dodd (1964), Hake and Stewart  (1977),  Morgan  (1969),  and
Ling and  Lindsay  (1975)  have reported  severe  and prompt  erythema after dermal
contact  with  PCE in man.

5.1.2  Oral  Absorption
     PCE is rapidly and  virtually completely absorbed into the  body  from  the
gastrointestinal tract, presumably because of its high lipid solubility.   Pegg
et al.   (1979)  and Schumann  et al.  (1980)  have  administered 14C-PCE,  1  and
                                    5-2

-------
500 mg/kg (in  corn  oil)  to both rats  and  mice and recovered better than 90
percent of the  radioactivity  in  urine,  feces,  exhaled  air,  etc.   For the  rat,
peak blood concentration  occurred  within 1 hour or oral dosing (Figure 5-1).
Similar results have been found by Daniel (1963) after oral dosing of   Cl-PCE
to rats (1.1 to 6.1 g/kg).  Frantz and Watanabe (1983) found virtually complete
              14
absorption of   C-PCE in drinking water (saturated; approximately 150 ppm w/v;
150 ug/mL) by  rats  ingesting  ad 1 ibitum over  a 12-hr  period with an average
consumption equivalent to an 8 mg/kg dose.

5.1.3  Pulmonary Absorption
5.1.3.1  Man.   PCE  in vapor form in  air  is  readily  absorbed through the  lungs
into blood by first-order diffusion processes.   Pulmonary uptake of a volatile
compound  like  PCE during  inhalation  exposure  is  largely determined by the
ventilation rate (about 4 to 8 L/min for man at rest), duration of exposure at
a given air concentration, solubility in blood and other body tissues,  and its
metabolism.   When body tissue concentrations (body burden) are at steady-state
with inspired  air  concentration,  the rate  of  uptake  is  equal to the rate of
metabolism plus  nonpulmonary  excretion  of  PCE  (i.e.,  by  definition uptake per
unit time must equal  excretion by all  routes  at  steady  state).  Since there
are  no  known  significant  routes  of excretion of  PCE  except  pulmonary and
metabolism (Section 5.2),  the  steady-state  uptake  rate  approximates the meta-
bolism rate.
     The blood/gas  (air)  partition  coefficient (A. b/g) at 37°C expresses the
solubility of  a solvent  such  as PCE  as  the ratio of the  concentrations  in
blood (mg/L) and  in air (mg/L).  When  there  is  no impairment of pulmonary
diffusion, circulation,  or ventilation,  this  partition coefficient exists
between the  arterial blood concentration  (C.)  and  the  inspired air concentra-
tion (C ) (or  exhaled air  concentration  postexposure),  so  that C./C  = A  b/g.
       cl                                                        Ma
As illustrated  in  Figure  5-2,  blood  concentration  parallels air concentration
in a fixed ratio.   The blood/gas partition  coefficient  for PCE is about  13 to
15.  It is  considerably  higher than other  solvents such as trichloroethylene
(TCI),  9.5 and methylchloroform (MC), 5 (Monster,  1979;  Sato  and Nakajima,
1979; Table 5-1).   Thus  PCE  has the potential for a greater pulmonary uptake
rate than TCI or MC.
                                    5-3

-------
         100
        i10
        UJ
        O
          0.1
                                               • 600 ppm, 6 hr
                                               • 500 mo/kg GAVAGE
                        10
                                     20
                                   TIME, hours
                                                 30
                                                              40
 Figure  5-1.
First-order excretion  curves for  PCE  in blood  of rats after
exposure to 600 ppm  for  6 hr (—) or to 500 mg/kg gavage doses
(—).  The animals were  adult male Sprague-Dawley rats weighing
approximately 250 g.   The body burdens of PCE were about  77.5 mg
per animal  from inhalation exposure and 123.2 mg per animal after
oral  dosage.   (From Pegg et al.,  1979).
     The  rate  of uptake of PCE by body tissues from blood for a given concen-
tration  in  inhaled air is determined by tissue volume,  the relative solubili-
ties in  blood  and tissue  (tissue/blood  partition  coefficients), and  by  the
tissue  blood flow.  In accordance with  the physiological model  for PCE of
Guberan and  Fernandez  (1974),  uptake by  body tissues  is  a blood  flow-dependent
process which  can  be  grouped into four physiological  compartments for conven-
ience in analysis:  the blood  vessel-rich group of tissues (VRG) (corresponding
to  brain,  heart,  hepatoportal  system,  kidneys, and  endocrine  glands),  the
muscle group (MG)  (muscle and skin),  the  fat  group (FG) (adipose tissue and
yellow marrow), and the vessel-poor group (VPG), composed of  connective  tissue
(bone,  cartilage,  etc.).   The predicted uptake and  distribution  of PCE to
these tissue groups during and after 8-hr exposure to 100 ppm  (678  mg/m3)  is
shown in Figure 5-3.   For an exposure of this magnitude,  body burden is  calcu-
lated as about 1000 mg PCE for a  70-kg man,  with  half this amount distributed
into fatty tissue.
     With prolonged inhalation exposure  to PCE, when  a whole-body steady-state
condition is reached with inspired air concentration, the concentration  of  PCE
                                    5-4

-------
           =  _  1000
           § .5

             o
             0.
                  100
                  10
                  0.1
                 0.01
                     i A
i   1  1   |   I  I   I   1  |   I   T  I   I
 O  72 ppm PCE AT REST
 A 144 ppm PCE AT REST
 D 142 ppm PCE AT REST AND WORKLOAD
                                                   EXHALED AIR
                                    50           100
                                  POST EXPOSURE, hours
                                  150
Figure 5-2.  PCE  concentrations in  blood  and exhaled air following  inhalation
             exposure  for 4 hr  (means of 6  subjects).   From Monster et  al.,
             1979.
                                     5-5

-------
1000
          4      8

      • EXPOSURE -
  4      8     12

POST EXPOSURE	«-

               TIME, hr
20
24
26
Figure 5-3.   Predicted uptake  and  distribution of PCE to tissue groups during
             and after an  8-hr exposure to  100 ppm.   See text  for details.
             From Guberan and Fernandez, 1974.
                                    5-6

-------
in the individual tissue groups relative to blood is determined by the tissue/
blood partition  coefficients.   For  highly  1ipid-soluble  solvents  such  as  PCE,
the adipose tissue has the highest partition coefficient and hence the highest
tissue concentration  at  steady state.   For the three solvents, PCE, TCI, and
MC, the adipose tissue/blood partition coefficients at 37°C are about 105, 75,
and 50, respectively (Monster, 1979; Sato and Nakajima, 1979; Table 5-1).  The
capacity of a  tissue  to take up PCE is the product of volume and solubility.
Since the  volume  of  adipose tissue in a 70-kg man is about 10L, then adipose
tissue capacity  for  PCE is about 1500 mg;  but  because of the small rate of
perfusion  of  adipose tissue  (about 0.4L blood/min;  Eger,  1963),  the time
needed to  equilibrate  the  adipose tissue is  large  in comparison to that of
other tissues.   Figure 5-3 indicates that a steady-state plateau concentration
is not achieved  for  adipose tissue within  the  period  of  an  8-hr  exposure to
PCE,  although  it is  for muscle tissue and  other  vessel-rich tissues.  For a
given  concentration  in blood (or air),  the half-time (TjJ (i.e., the time
necessary  to equilibrate the adipose tissue to 50 percent  of  its final  con-
centration) is about 25  hours for PCE (Monster, 1979; Fernandez et al., 1976).
This  means  that  during a single 8-hr exposure, adipose tissue does not reach
steady-state equilibrium,  which requires about five times  TI ,  or about  125
                                                             ^
hours.
     At whole-body steady-state condition,  the  net pulmonary uptake  of PCE is
balanced with excretion  of PCE by other routes, principally by metabolism.  The
difference between the inspired air concentration and end alveolar air concen-
tration (exhaled air concentration) provides a measure of PCE uptake, so that:

                           Q =  (Cins - Caly) V  •  T                  (5-1)

where Q is the amount absorbed during time  T (min), C is the air concentration
in mg/L, and V is the ventilation rate in L/min.  The percent retention (R) is
defined as:

                             D _  (Cins - Ca1v)                       (5-2)
                             R -      c:
                                       ins
and therefore
                           Q = R (C.    • V  • T)/100                    (5-3)
                                    5-7

-------
Figure 5-4 shows the  change  in end alveolar concentration during exposure of
5 subjects for 8 hours at  100  ppm  (678 mg/m ) (Fernandez et al., 1976).  The
data show a rapid rise in alveolar  concentration during the first hour of expo-
sure, followed by a slower but  continuous rise during the subsequent 7 hours of
exposure  (retention value decreasing with exposure) towards a  fixed  retention
value at  whole-body equilibrium, although it  is apparent  from  Figure 5-4  that
equilibrium or steady-state retention  is not  reached during an 8-hr  exposure.
Thus, rate of uptake,  at first  rapid,  diminishes with continuing exposure to a
constant rate at steady state.   Retention values of 60 to 80 percent have been
reported  at the  end  of short exposures  (4  to 6 hours),  and lower retention
values would be expected with longer exposures (Fernandez et al., 1976; Monster
et al., 1979;  Bolanowska and Golacka,  1972).  These values indicate  pulmonary
uptake of  PCE  is rather high.   Since  the difference  in  retention, 20 to 40
percent,  represents  excretion  by  other  than pulmonary routes (principally
metabolism),  these  retention  values suggest  that  metabolism  of  PCE may  be
considerably greater  than  indicated by other measures of metabolism (Section
5.2.2).
     Monster et  al. (1979) have subjected volunteers to controlled  inhalation
exposures  for  4 hours and determined  experimentally  PCE  uptake kinetics.
Monster et al. found that net pulmonary uptake decreased with exposure duration,
with about 25 percent less PCE  absorbed in the fourth hour as compared with the
first hour of exposure.  Table  5-2  gives the amounts (mg) PCE uptake at 72 and
144 ppm (488 and 976 mg/m ) exposure for 4 hours.   For 6 subjects, the average
uptake was 455 mg and 945 mg for the two exposure concentrations, respectively,
showing a  direct proportionality of uptake  with inspired  air concentration  at
least up  to 144  ppm (976 mg/m  ).   Their  data  (Table 5-2)  also  demonstrate the
marked increase  in  uptake  rate (and total uptake) as a result of an increase
in  ventilation  rate provoked  by  exercise.    Interperson  variations in PCE
uptake were also influenced by  total body mass and adipose tissue mass.
     Fernandez and co-workers  (Guberan and  Fernandez,  1974;  Fernandez et  al.,
1979) and Stewart and co-workers  (Hake  and Stewart,  1977; Stewart  et al. ,
1961, 1970, 1974)  have also  conducted controlled exposures with human volun-
teers.  Their results are in accord with those of Monster and co-investigators.
Hake and Stewart (1977) exposed five subjects for 5 consecutive days to 100 ppm
(678 mg/m ) PCE  for 7  hours daily.   As shown  in Figure 5-5,  the  concentration
of PCE in  exhaled  breath of the 5 subjects increased  as  the 5-day week pro-
gressed,  indicating an increase in  the body burden with repeated daily  exposure.  , , ,
                                    5-8

-------
  60
  50
a
a

2 40
<
cc
LU

g 30

O
o

cc
o
LU

^ 20

<
  10
       V


                                   4       5


                                EXPOSURE T!ME,hr
Figure  5-4.   PCE alveolar  air concentration during exposure  of  5  subjects for

              8 hr at 100 ppm.  From  Fernandez et al.,  1976.
                                      5-9

-------
     TABLE  5-2.   ESTIMATED  UPTAKE  OF  SIX  INDIVIDUALS EXPOSED TO PCE AT REST
                               AND AFTER  EXERCISE
Subject
A
B
C
D
E
F
Average
Body
Mass
(kg)
70
82
82
86
67
77
77.3
Lean
body
mass
(kg)
62
71
71
74
61
61
66.7
Ventilation
Min volume
at rest
(L/m)
7.6
11.6
10.0
11.3
12.3
8.8
10.3
72 ppm
(at rest)
(mg)
370
490
530
500
390
450
455
Uptake*
144 ppm
(at rest)
(mg)
670
940
1000
1210
880
970
945
144 ppm
+ exercise
(mg)
1060
1500
1400
1510
1320
1120
1318
 ^Exposures  for 4  hr; exercise two 30-min periods during 4-hr  exposure.
 Source:   Monster  et al. , 1979.

 Fernandez et al. , using their  mathematical model,  predicted that  chronic
 occupational  exposure  to  100 ppm (678 mg/m3) PCE  (8 hr/day,  5 day/wk) would
 result  in cumulation  of PCE in  adipose  tissue.   As uptake by  fatty tissue
 during  the  working hours of the  week equals the  elimination during nights and
 weekends, an  equilibrium  is eventually established  but which  requires  a  time
 period  of 3 to 4 work weeks.
 5.1.3.2   Rodents.   Pegg et al.  (1979) and  Schumann  et al.  (1980)  have  deter-
 mined the overall  pulmonary uptake  of rats  and mice exposed to PCE by inhala-
 tion for  6 hours.   The animals were exposed in a 30-liter glass chamber to 10 or
 600 ppm (67.8 or 4068 mg/m3) 14C-PCE.  At termination of exposure  (6  hours),  the
 cumulative pulmonary uptake was  estimated by determining  the  radioactivity of
 carcass,  and  radioactivity  in  expired air,  urine,  and  feces.   The  data  for
 rats and  mice are given in Tables 5-3 and  5-4,  respectively.   For mice,  the
body burden resulting  from  a 10-ppm  exposure was 0.40 mg/animal  (16.5 rag/kg),
and for rats, 1.48 and 77.5 mg/animal (5.9 and 310 mg/kg)  for 10  and 600 ppm
(67.8 or 4068 mg/m ),  respectively.   These data indicate that pulmonary uptake
                                    5-10

-------
LJJ
O
CC
O

I
O
            §: 100
            z
            O
               10
            Z
            LLJ
            O
            Z
            O
            O
                    I   I   I    I   I   I
                  7 hr VAPOR EXPOSURES
                                   1    I
                         5   6   7  8

                            TIME, days
                                                 9  10  11  12  13  14
Figure 5-5.   Mean  exhaled  breath concentrations  of PCE for five  volunteers
             exposed to  100  ppm PCE for 7 hr  per day for 5 days.   Note  the
             rising concentration  of PCE toward equilibrium with inspired air
             concentration with daily  exposure,  and the long  decay of PCE
             concentration in  expired  air  after the  last  exposure.  (From
             Hake and Stewart, 1977; Stewart et al.,  1970).
                                    5-11

-------
   TABLE 5-3.   DISPOSITION IN RATS OF 14OPCE RADIOACTIVITY 72 HR  FOLLOWING
                       ORAL OR INHALATION ADMINISTRATION
                                       Recovery, pmol equ. PCE
                       1 mg/kg*
                              500 mg/kg
Expired air
Unchanged PCE
Metabol ized
14C02
Uri ne
Feces
Carcass

1.05 ± 0.00

0.04 ± 0.01
0.24 ± 0.00
0.09 ± 0.02
0.05 ± 0.00

71.5

2.5
16.5
6.2
3.3

667.31 ± 22.72

3.44 ± 0.36
34.48 ± 2.40
29.08 ± 4.01
8.50 ± 0.66

89.9

0.5
4.6
3.9
1.2
Total
Expired air
  Unchanged PCE
1.46 ± 0.02
10 ppm,  6 hr*

6.08 ± 0.44
68.1
742.79 ± 27.75
600 ppm, 6 hr

412.38 + 7.60
88.0
Metabol ized
14C02
Urine
Feces
Carcass
Total

0.
1.
0.
0.
8.

32 ± 0.
66 ± 0.
46 ± 0.
38 ± 0.
91 ± 0.

05
12
02
06
33

3.
18.
5.
4.


6
7
2
3


3.
27.
14.
10.
467.

25
40
24
07
05

± 0.
± 1.
± 0.
± 0.
± 8.

17
68
75
54
11

0.
6.
3.
2.


7
0
1
2

*Mean ± SE for 3 rats; gavage doses in corn oil in volume of 1.0 mL/kg;
 Sprague-Dawley rats.
Source:  Pegg et al. ,  1979.

of PCE by the rat at least is approximately proportional to exposure concentra-
tion.
     On a comparative  species  basis,  the pulmonary uptake (comparing 10 ppm,
6-hr exposures) for the  rat  and mouse more  closely  relates to the ratio of
                              2/3            ?/^
their surface areas (b.w.  rat  /b.w.  mouse '  = 4.7) than to their relative
body weights  (ratio  10.2).   Recently, Landry et al.  (1983) have  developed
methods to measure simultaneously  respiratory  frequency, tidal volume,  minute
volume, and net uptake  of an inhaled vapor in rats.   During steady state, if
metabolism is the only significant route of  elimination, then  net  uptake  rate
of the inhaled vapor  is  equal  to   the rate of  metabolism.   This  new  approach
                                    5-12

-------
   TABLE  5-4.   DISPOSITION IN MICE OF 14OPCE RADIOACTIVITY 72 HR FOLLOWING
                       ORAL OR INHALATION ADMINISTRATION

Expired air
Unchanged PCE
Metabol ized
14C02
Urine
Feces
Carcass
Cagewash
Total

*
10 ppm, 6 hr.

0.29 ± 0.06

0.19 ± 0.01
1.53 ± 0.20
0.16 ± 0.04
0.07 ± 0.01
0.19 ± 0.11
2.44 ± 0.27
Recovery
%

12.0

7.9
62.5
6.8
3.0
7.7

, pmol equ. PCE
500 mg/kg

53.69 ± 4.96

0.85 ± 0.67
6.58 ± 3.19
0.80 ± 0.05
0.32 ± 0.07
2.67 ± 1.53
64.92 ± 3.09

%

82.6

1.3
10.3
1.2
0.5
4.1

*Mean ± SD for 3 rats;  gavage doses in corn oil;  BgC^f^ mice.
Source:   Schumann et al.,  1980.

appears to  be useful  for inhalation  dosimetry and  evaluation of metabolic
rates in rats.

5.1.4  Tissue Distribution and Concentrations
     PCE is expected to  distribute,  by first-order diffusion processes,  into
all body tissues,  as  is  known for other chloroethylene compounds with lower
lipid solubility  (Table 5-1;  Waters  et al.,  1977).   PCE readily crosses the
blood-brain barrier and the placental barrier.   The partition coefficients  for
various tissues,  relative to  blood or air, have  not been firmly established,
although approximations can  be  calculated from the  rat  tissue distribution
data of Savolainen  et al. (1977)  (Table 5-5); rat tissue/blood:  adipose, 60;
liver, 5; brain,  4.  Absolute tissue  concentrations  are directly proportional
to the body burden or exposure dose.
5.1.4.1  Rodents.   Savolainen et al.  (1977) analyzed  tissue levels of PCE
         	                                                     3
after exposure of rats 6 hours daily for 5 days to 200 ppm (1356 mg/m ) PCE in
air.  Values  are  shown  in Table 5-5.  Seventeen  hours after  exposure  on day
four, blood and  other  tissues still  contained significant concentrations of
PCE with highest  concentrations  in adipose tissue  (103 pg/g)  reflecting the
high solubility of  PCE  in this tissue and  its  long storage or half-time of

                                    5-13

-------
     TABLE 5-5   RAT ORGAN CONTENT OF PCE AFTER DAILY INHALATION EXPOSURE
       OF 200 PPM FOR 6 HOURS PER DAY.   MEASUREMENTS MADE ON FIFTH DAY.
Time (hr of
exposure
on fifth day)

0
(17 hr after
day 4
exposure)
2
3
4
6
Cerebel

18.4
±3.7
89.7
+16.7
108.7
±4.9
101.5
±17.2
142.5
±0.2
1 urn Cerebrum
nmol/g (mean of
13.1
±4.4
62.0
±7.1
72.0
±4.5
68.1
±5.5
92.3
±1.3
Lungs
Liver
two determinations
9.5
±1.8
45.6
±10.1
52.4
+9.8
59.5
±13.4
74.0
+3.6
35.2
±9.3
107.4
±27.3
134.9
±1.0
133.8
±0.7
160.7
±24.2
Peri renal
Fat
+ range)
622.2
±20.0
977.6
±174.0
807.1
±38.0
1105.3
±109.7
1724.8
±422.4
Blood

4.3
±1.1
21.3
±4.0
25.3
±1.5
24.4
±3.4
30.3
+6.6
Source:  Savolainen et al.,  1977.

desaturation.  With exposure on day five, tissue cumulation occurred in brain,
lungs, liver,  fat,  and  blood,  and steady-state level was not achieved by the
end of the 6-hr exposure.
5.1.4.2  Man.   Except for blood concentrations  (Figure 5-1).  no  direct  analy-
tical  information of  tissue  levels after various exposure  concentrations of
PCE is available.   As  for  the rat, tissue concentrations are expected  to be
proportional to exposure concentrations  and to  duration  of  exposure.  Guberan
and Fernandez (1974) developed a mathematical  model  of PCE uptake, distribution,
and excretion (based on experimental data).   Their physiological model predic-
ted the distribution  of  PCE in the three major physiological 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.   Figure 5-2
shows that by  the end of exposure, 50 percent of the solvent taken up by the
body has been  distributed to  the fatty  tissue, due  to its  high  tissue-blood
partition coefficient.  The distribution of PCE to the other  tissue  groups  -is
                                    5-14

-------
related to their volume and partition coefficient and  therefore  is  higher for
the MG than for the VRG and the vessel-poor group (VPG).  After  exposure, the
depletion of  these  three  groups of tissues  is  almost complete after about
20 hours, and the continuing elimination in alveolar air is then related to the
slow release of PCE from adipose tissue.  Beyond about 20 hours of elimination,
the rate of discharge  of  PCE from the adipose tissue, and therefore from the
whole body, is  an  exponential  function of time with a predicted half-life of
about 71 hours.  Because  of this long  residence  of PCE in adipose tissue,
repeated daily  exposure (6  hr/day, 100  ppm) results  in  an accumulated concen-
tration,  as PCE from new exposures adds to residual  concentration from previous
exposures,  until steady state is reached.  Adipose  tissue levels of PCE  reach
a  steady-state  "plateau"  (dependent  on inhalation  concentration and daily
exposure duration)  and thereafter remain relatively constant.   For a 6-hr dc
100 ppm  (678  mg/m )  exposure,  Guberan  and Fernandez (1974) predicted that
plateau  concentrations  in adipose tissue  occur about 22 to  29  days  after
initial daily exposure.
5.2  EXCRETION
     The total  elimination  of an absorbed body  burden  of PCE involves two
major processes—pulmonary excretion of unchanged PCE and metabolism to urinary
metabolites.  The metabolism  of  PCE appears to  be  very limited  in man and
animals and  is  reviewed  extensively in Section  5.4.  PCE or  its metabolites
have not been  reported to be excreted by other routes than lung and urine in
any significant amounts.   Bolanowski and Golacka (1972)  have reported that the
rate of PCE  elimination  through  the skin  in man is  about 125 ug/hr during a
6-hr exposure  to 390 ppm  (694 mg/m  ),  or about 0.02  percent of the uptake/hr.
Thus, pulmonary elimination is the major route of PCE excretion.
     In rodents, the  principal  metabolites of PCE  are  trichloroacetic acid
(TCA) and  oxalic acid.   In  man,  TCA is  the  predominant metabolite although
oxalic acid may be  unrecognized.

5.2.1  Pulmonary Elimination in Man
     Pulmonary elimination  of unchanged  PCE after exposure occurs as a first-
order diffusion process across the  lungs from blood  to  alveolar air, depicted
in a manner as an inverted reproduction of its uptake as illustrated in
                                    5-15

-------
Figure 5-3.   From controlled experimental exposures in humans, Stewart and co-
workers (Stewart et al.,  1961,  1970,  1974; Hake and Stewart, 1977) have followed
the body  desaturation  of  PCE  after  exposure by  serial  breath analysis of
alveolar air.   Figure 5-5  shows  the  results following exposure of 5 subjects
for 5 consecutive days  to 100 ppm (678 mg/m3) PCE for 7 hours daily.  At least
two first-order  phases  of  pulmonary  elimination of  PCE  are apparent:   an
initial fast  phase,  followed by a slow  predominant excretion phase with  a
half-time of  about 65  hours.  Monster et al.  (1979)  have also studied the
kinetics of pulmonary elimination  of  PCE in experimental human exposures and
determined from  the  concentration curves  of PCE in  exhaled air  and blood
(Figure 5-2)  that  the compound was eliminated from  the body  through  the lungs
at three different first-order rate constants with corresponding half-lives of
12 to  16  hours,  30 to  40 hours, and  55  to  65 hours,  in accordance with the
desaturation of  three major  body compartments,  represented by  VRG,  MG,  and FG
compartments,  respectively (Monster  et  al. ,  1979;  Guberan  and  Fernandez,
1974).  The long half-time (55 to 65 hours)  of elimination of PCE from adipose
tissue (FG),  which is due to the high adipose tissue/blood partition coefficient
of PCE and the low rate of blood perfusion of the tissue (Eger, 1963), is inde-
pendent of the body burden of PCE as shown by the parallel blood and exhaled air
concentration decay curves of Figure 5-2 for exposure concentrations of 72 and
144 ppm (488 and 976 mg/m3).
     However, the  exhaled  air  (or  end alveolar air) concentrations and blood
concentrations  of  PCE after exposure  and throughout desaturation are propor-
tional to the acquired body burden or exposure concentration and duration, and
serve  as  a means for estimating body burdens,  as illustrated  in  Figures  5-2
and 5-6 (see  Section 5-3  for discussion).  Guberan and Fernandez (1974) have
developed a physiological  kinetic  model  for PCE  excretion  in  man,  based on
experimental  findings from controlled exposures of volunteers.  The alveolar
concentrations during and after exposure in the excretory phase were predicted
for use in estimating  mean exposure in occupational and environmental situa-
tions.  The  predicted  half-life of  PCE stores in  adipose  tissue was 71.5
hours, in good  agreement with the  experimental  observations  of Stewart  et al.
and of Monster  et  al.  The long  half-time of pulmonary excretion  of PCE shows
that a  long period is necessary  to completely eliminate  PCE, i.e.,  above  five
times the half-life or about 2 weeks.
                                    5-16

-------
  300
                                      DAY
Figure 5-6.
Daily (8-hr)  occupational  inhalation  exposure to  PCE.  Course of
the  relative  (first Monday morning concentration taken  as 100
percent) concentration of trichloroacetic acid (TCA) in blood and
urine (mean of  23 subjects ± S.D.) during the work week.  c	
Monster et al., 1983.
                                                                        From
                                    5-17

-------
5.2.2  Urinary Metabolite Excretion in Man
     In contrast to first-order pulmonary excretion of PCE, the urinary excre-
tion of TCA (or total trichloro-compounds by Fujiwara reaction) from metabolism
of  PCE  is dose-dependent, and presumably  follows Michaelis-Menten kinetics
[See Section 5.4.3 for discussion].  Most studies have quantitated PCE urinary
metabolite excretion by the Fujiwara reaction for which there is a question of
the true  nature of  the compound(s) measured, although  it  is  accepted  that TCA
is the principal compound.  Thus Ikeda and Imamura (1973) found the mean T^ of
total urinary  trichloro-compounds  for  13 subjects occupationally  exposed to
PCE to be 144 hours (range 123 to 190 hours).  In contrast, when TCA is directly
administered to men, the half-life has been found to be 51 to 82 hours (Paykoc
and Powell,  1945; Muller  et al., 1974).  Hence,  the  longer T^ of TCA  from PCE
metabolism  is  no  doubt  due to constant  metabolic formation  of TCA from PCE
cycling to  the  liver over the period of its long residence time in adipose
tissue (Tj , 17 to 55 hours).
         'i
     Monster  and  co-workers  (Monster et al. , 1979;  Monster  and Houtkooper,
1979; Monster  et  al. ,  1983),  using gas chromatographic methods for analysis,
                                                       3
exposed volunteers  to  72  and  144 ppm (488 and 976 mg/m ) PCE for 4 hours and
determined  TCA  in  blood  and  urine.  They  found  that urinary TCA represented
less than 1 percent of the estimated absorbed dose of PCE.  In blood, following
exposure, TCA  continued  to increase in  concentration  until  about 20  hours
(representing continued metabolism of PCE in the body to TCA) and then declined
exponentially as a  first-order elimination process (Figure 5-7).  The blood
concentration of TCA was  proportional  to exposure concentration (i.e., body
burden of PCE) and declined with a half-life of about 65 to 90 hours.   Figure
5-8 shows that urinary cumulation paralleled blood TCA disappearance, and also
that the  urinary TCA cumulation was proportional to the  inhalation exposure
concentration of PCE (and hence body burden).  Physical activity increases PCE
pulmonary uptake (body burden), and hence, increased TCA formation and urinary
excretion.  However, more recently Ohtsuki et al. (1983) correlated inhalation
exposure concentration with urinary  total  trichloro-compounds and found that
urinary metabolite  excretion was not a  linear function of exposure concentra-
tion but  instead a  hyperbolic function  approaching  a  "plateau" at about  400
ppm exposure (Figure 5-9).  Since urinary metabolite(s) excretion reflects PCE
metabolism,  these observations indicate  that PCE  metabolism  is  dose-dependent
and saturable with high PCE body burdens or exposure.
                                    5-18

-------
              O)
              E
              o
              o
              O
              O
              _l
              03
0.6 —
0.4 —

0.2 —

0.1 —
                             £144 ppm PCE AT REST
                             D142 ppm PCE AT REST AND WORKLOAD
                        I  I   I  I  I  I   I  I  I  I   I  I  I  I   I  I  I  I
                        0           50          100         150
                               TIME AFTER EXPOSURE, hours
Figure 5-7.  Trichloroacetic acid  blood concentrations  following inhalation
             exposure  to PCE  for  4  hr (means  of 6 subjects).   From Monster
             et al., 1979.
k.
QJ

•a 10
< 8
£ 6
LLJ
Z 4
i 2
n
I I I I
72 ppm PCE , 1
AT REST I
I
— I
:I^i
44 p
AT
rh

1 1
pm PCE
REST

¥1


*

! 1
142 ppm PCE
AT REST AND
WORKLOAD
"MT -
                     0  22 46 70     0 22 46 70     0 22 46 70
                            TIME AFTER START EXPOSURE, hours
Figure 5-8.  Urinary  excretion of  trichloroacetic acid  during  and following
             inhalation  exposure  to PCE  for 4  hr (means  of  6 subjects).
             From Monster  et al. ,  1979.
                                     5-19

-------
            150
                   100
                        200
                             300    400   500   600    700

                               PCE CONCENTRATION IN AIR, ppm
                                                        800
                                                              900
 Figure  5-9.   Relationship  between  PCE occupational inhalation exposure (time-
              weighted  ppm  average  for an 8-hr work shift) and urinary concen-
              tration of total trichloro-compounds  at  end  of  work shift (correc-
              ted  to a  urine specific  gravity  of 1.016) for  36  male (squares)
              and  25 female  workers  (circles).  From Ohtsuji et al. ,  1983.
 5.2.3   Chronic  Exposure
     The  long half-time of elimination of PCE  by the  pulmonary  route  (~  65 hr)
 and  by  urinary  metabolite(s) (^ 65 to 144 hr)  indicates that, with repetitive
 daily or  chronic  inhalation  exposure to  PCE,  sustained or "plateau" levels of
 PCE  and TCA in  blood and of TCA  in urine would occur  as body burden, metabolism,
 and  excretion approach a steady state with pulmonary uptake.   Monster et al.
 (1983)  determined  blood  levels  of PCE and  TCA and urine TCA concentrations
 during occupational exposure (8  to 50 ppm, 8 hours daily) during the course of
 the  work week.  Figures 5-6 and  5-10 show their data  expressed  in terms  relative
 to Monday morning concentrations (control; 100 percent) occurring from previous
 work exposure for  23 male  and female workers.   Figure 5-10  shows that exhaled
 air  concentration (a measure of  body burden) and blood concentration  (directly
 proportional  to exhaled  air  concentration and also a measure of body  burden)
 both rose 4-fold  with  the  3-day exposures of  Monday, Tuesday,  and Wednesday,
 and  tended  to "plateau"  with Thursday and Friday  exposures.  These "plateau"
 levels declined to  Monday  morning levels (100 percent)  during  exposure-free
 days, Saturday  and  Sunday.   Figure 5-6  shows  that the concentrations of the
metabolite TCA  increased during the work week in  blood as  well as in urine.
The  relative  concentrations  on  Wednesday and  Friday  after work were higher by
a factor  of two than on Monday  morning.   The  increase of  the concentration in
                                    5-20

-------
          800
                   MON   TUES   WED   THUR   FRI

                                      DAY
                                               SAT   SUN
                                                         MON
Figure 5-10.  Daily (8-hr) occupational inhalation exposure to PCE.  Course of
              the relative  concentration  (first Monday morning concentration
              taken as 100 percent) of PCE in blood and exhaled air during the
              work week  (mean  of 23 subjects ±  S.D.).   From  Monster et  al. ,
              1983.
urine was more  pronounced than that in blood.  At the start of the following
week the  relative concentrations had returned to  about  100 percent.   Thus,
during daily exposure  during the work week, the body burden and blood levels
of PCE increased with each exposure and then tended to plateau towards the end
of the week, but the metabolite TCA continued to accumulate in blood and urine
apparently  as a result of the longer half-life of TCA.  Similar experimental
observations have been reported by Tada and Nakaaki (1969), and predicted from
the mathematical model of Guberan and Fernandez (1974).
5.2.4  Excretion Kinetics in the Rodent
     As is the  case for man, PCE is excreted from the rodent (rat and mouse)
principally  by  two routes:   pulmonary  elimination  of unchanged PCE  and  by
metabolism and renal elimination of PCE metabolites.  For the rat, Pegg et  al.
(1979) and Frantz  and Watanabe  (1983)  found that the rate  of PCE excretion  in
expired air is described by apparent first-order  kinetics.   No significant  dif-
ference was observed in the elimination half-time (approximately 7 hours) with
                                    5-21

-------
either dose or route of administration (inhalation, gavage, or drinking water).
Figure 5-1  shows the  disappearance  of PCE from  whole  blood which  follows
apparent first-order kinetics for up to 36 hours with elimination  rate constant
of 0.10 to 0.12 hr"1.
     The percentage  of  the  total body burden excreted by the pulmonary route
in the rat and mouse depends on the dose.   As calculated from the  data of Pegg
et al. (1979) and Schumann et al. (1980) (Tables 5-3 and 5-4), the body burdens
of PCE  in  rats  after a 6-hr  inhalation exposure  to 10 and  600  ppm  (68  and
4068 mg/m3) are  1.48 and  77  5 mg per  animal,  respectively;  the percentages of
these body  burdens   excreted by the pulmonary route as unchanged  PCE were 68
                                                                       3
and  99  percent,  respectively.   For the mouse  exposed to 10  ppm (68 mg/m  )  PCE
for  6  hours,  the body  burden was 0.40 mg and the pulmonary excretion only 12
percent, whereas  for a body burden of 10.8  mg  from oral  administration, 83
percent was excreted by the pulmonary route.   Similar results can  be calculated
for  body  burdens  from  gavage and drinking water administration from the data
of Tables  5-3,  5-4,  and 5-6  (Pegg et  al.,  1979; Schumann et  al. ,  1980; Frantz
and Watanabe, 1983).  Hence, as the body burden of PCE is increased  in the rat
or mouse, the percentage excreted unchanged increases.   Conversely,  as metabo-
lism  is  the  other principal  route of  elimination  of PCE, when  the body burden
increases,  the  percentage of  the  burden  metabolized (urinary metabolites)
decreases (although the absolute amount increases) (Tables 5-3, 5-4, and 5-6).
These observations  suggest  that metabolism of PCE and  urinary excretion of
metabolites  in  the   rodent  are  rate-limited  and  dose-dependent,  following
Michael is-Menten  kinetics (whereas pulmonary excretion is a first-'order process
and dose-independent,  i.e.,  with half-time and rate constant  independent  of
dose).
     Other investigators have found evidence of dose-dependent Michaelis-Menten
metabolism kinetics.  Filser and Bolt (1979) exposed Wistar rats to  PCE in air
in a  closed system  and determined  the pharmacokinetics from  disappearance
rate of PCE  from the chamber (uptake  by rat).  In comparison  to  other halo-
genated ethylenes,  they found PCE to  be metabolized extremely  slowly in  rats.
Zero order Vmax  (saturation) was <1.16 mg/hr/kg b.w.; for a 250-g rat, about
3.0 mg PCE/hr is metabolized and excreted as urinary TCA and other metabolites.
The principal urinary metabolites in the rat are oxalic acid and TCA (Table 5-7)
Buben and O'Flaherty (1984), using Swiss-Cox mice, found urinary metabolite(s)
(TCA only in  these  mice)  excretion was limited by metabolic  capacity.   Mice
were chronically administered PCE by gavage at dose levels up to 2000 mg/kg/day.
                                    5-22

-------
  TABLE  5-6.   DISPOSITION  IN  RATS  OF  14OPCE RADIOACTIVITY 72 HR FOLLOWING
                           DRINKING WATER INGESTION*

Expired air
Unchanged PCE
Metabolized
14C02
Urine
Feces
Carcass
Cagewash
Total
Recovery, umol equ.
8.09 mg/kg*
12.47 ± 5.15**
0.29 ± 0.08
0.96 ± 0.26
0.23 ± 0.07
0.15 ± 0.04
0.02 ± 0.02
14.10 ± 5.51
PCE
%
87.9
2.2
7.2
1.7
0.9
0.2

 *Rats consumed an average dose of 8.09 ± 3.13 mg/kg,  which varied with the
  volume of drinking water ingested over a 12-hr exposure period.
**Each value represents mean ± SD from four Sprague-Dawley rats,
Source:   Frantz and Watanabe, 1983.

A plot  of amount  of  urinary  metabolite  (TCA)  per day versus dose  level fitted
the Michaelis-Menten function showing the metabolism of PCE is capacity-1imi ted
even  at  relatively  low doses  (below 100-200 mg/kg) (Figure 5-11), with an
estimated V    for urinary metabolite excretion of 136 mg/kg/day (5.7 mg/hr/kg
           max
b.w. ) and a K  of  660 mg/kg.  These  values  suggest  an  excretion by metabolism
in mice greater  than that estimated in  the  rat by  Filser  and Bolt (1979)  by
other experimental means.

5.3  MEASURES OF EXPOSURE AND BODY BURDEN
     The pharmacokinetic  parameters  of PCE uptake, metabolism,  and pulmonary
excretion obtained from experimental human and animal  exposures  have been used
to determine reliable estimates of body burdens occasioned by occupational  and
environmental  exposures (Stewart et al., 1961, 1970, 1974; Boettner and Muranko,
1969; Ikeda et al., 1972;  Essing et al., 1973; Ikeda and Imamura,  1973; Guberan
and Fernandez, 1974;  Fernandez  et al.,  1976;  May,  1976;  Monster et al., 1979;
Monster and  Houtkooper, 1979; Tada  and Nakaaki,  1969; Ogata et  al.,  1971;
                                    5-23

-------
           TABLE  5-7.   REPORTED  METABOLITES  OF PCE (OTHER THAN TCA)

       Metabolite                   System                Investigator
Inorganic chloride                Rat urine           Daniel,  1963
Carbon dioxide                   Rat,  mouse           Pegg et  al., 1979
                                   exhaled air       Schumann et al.,  1980
Oxalic acid                      Rat,  mouse           Yllner,  1961
                                   urine              Pegg et  al. , 1979
                                                     Dmitrieva, 1967
Dichloroacetic acid              Mouse urine         Yllner,  1961
Ethylene glycol                   Rat urine           Dmitrieva, 1967
              Production from PCE-oxide Spontaneous Decomposition
                      in Aqueous and Non-aqueous Solution
Trichloroacetic acid                                 Bonse et al.,  1975
Trichloroacetyl  chloride                             Kline et al.,  1978
Verberk and  Scheffers,  1980; Ziglio,  1981;  Ohtsuki  et  al. ,  1983;  Monster
et al., 1983; Monster and Smolders,  1984).   From these studies, three approaches
have been developed to estimate body burdens of PCE:  1) concentration of PCE
in exhaled air,  2) blood levels  of PCE,  and 3)  concentrations of trichloroacetic
acid (TCA) in blood  and urine.   Stewart et al.  (1961)  pioneered the use of
expired air  concentrations as a measure of body burden.   In  addition to the
advantage of a non-invasive methodology, alveolar PCE concentrations appear to
be a good index of the vapor exposure to which individuals  have  most recently
been subjected (Figure 5-12).  Urine analysis  for TCA excretion  appears to  be
of lesser value in  estimating  exposure to PCE.  The urinary excretion of TCA
increases slowly  and gains the maximal  level  only  after 1  to  4 days,  and
furthermore,  the level is not clearly proportional  to inhalation concentration
(Weiss, 1969; Tada and  Nakaaki,  1969;  Kundeg and Hogger, 1970;  Ikeda et al.,
1972; Fernandez et al.,  1976; Ohtsuji  et al.,  1983), owing to the long half-
life of PCE  metabolites  that appear in  urine  (144  hr,  Ikeda,  1977)  and  to  a
low  metabolic rate  which  is dose-dependent  and saturable  (Section  5.4.3).
Monster and co-workers (Monster et al., 1983; Monster and Smolders, 1984) find
                                    5-24

-------
           120
                  200
                          500
                                      1000
                                       DOSE, mg/kg
                                                    1500
                                                                2000
Figure 5-11.
Relationship between PCE and dose and the amount of total urinary
metabolite excreted per day by mice.  Data points are means of 9
to 11  mice ± SEM.   The  points  fit  the Michaelis-Menten  equation
with an r2 value of 0.996.  The calculated value from the curvi-
              linear fit  are V    = 136 mg/kg/day  and  K  = 660 mg/kg.
              Buben and 0'FlaheTty, 1984.                m
                                                           From
that  for  measuring environmental exposure  to  PCE,  biological monitoring of
exhaled air  is a  "simple,  efficient,  effective  and convenient method" of
assessing total ambient exposure of both young and aged  subjects.   For  measuring
occupational exposure, they find that the best parameter to estimate the time-
weighted average exposure to PCE over the whole work week  is  the exhaled concen-
tration of  PCE and TCA in blood  at the end of the  workday on Friday.   The
second best parameter is measurement of blood PCE and exhaled air concentration.
     These methods  for  estimating PCE body  burdens  after  single or chronic
exposure  are  subject to  high  inter-individual  variations that  limit  their
predictive value.   Some of the factors  known to contribute to these variabili-
ties are  pulmonary dysfunctional  states, individual differences in intrinsic
metabolic capacity for PCE, body mass and adipose tissue mass, modification of
metabolism by  drugs  and environmental  xenobiotics,  age, sex,  and exercise or
workload.   Guberan  and  Fernandez  (1974),  using  a  mathematical model  developed
to predict uptake  and distribution of PCE  in the body and its elimination in
alveolar air, have computer-simulated the effects of age, body weight,  height,
and body fat content, both at rest and during physical exercise.
                                    5-25

-------
 5000
 1755-
 1000

 805—



to
-f 500
 o

 a.

cc
<
Q
LU
-i
<
I
X
IU
z

m 100
   50
   10
         Ln PERC = -2.79+1.345 Ln TWA

         R2 = 0.927

          A= 1.43

          n = 63
 • WORKSHOP A

 • WORKSHOP B

 A WORKSHOP^

 • WORKSHOP D

	 REGRESSION LINE

	95% CONFIDENCE LIMIT
                      /"  -,
                 *   /*       *
                                                                            TT
           /     ,
               /
         I  I   KM  I
           50
 Figure 5-12.
          100
500     1000     2000

                  2050
                                                                    II
                                                                              10,000
                       PCE IN WORKROOM AIR AS 4-HOUR TWA, /zmol/m3
      Direct  linear relationship between  the time-weighted average
      occupational  exposure to  PCE  over the  last 4 hr of  a work day
      and the  concentration of  PCE  in  exhaled air 15-30 rain after the
      end of  exposure  (data for  32 subjects).   Inhalation exposure
      concentration  of 2,050 umol/m3 is equivalent to 50  ppm.   From
      Monster et al.,  1983.
                                      5-26

-------
5.4  METABOLISM
5.4.1  Known Metabolites
     Trichloroacetic acid  (TCA)  is  the most prominent  metabolite  of PCE in
animals and man.   The principal site of metabolism is the liver.  Bonse et al.
(1975) have perfused the isolated rat liver with PCE and directly demonstrated
the capacity of  this  organ to metabolize PCE to TCA; however, no other meta-
bolite (e.g., oxalic acid) was found.  Other tissues, such as lung and kidney,
may also metabolize  PCE,  as these tissues  are  known also to contain  P-450
metabolizing systems.   Nonetheless, other metabolites have also been described
(Table 5-7).  The metabolism of PCE was first systematically investigated with
14
  C-labelled compound  some 25  years  ago by  Yllner  (1961).  Yllner established
the principal  urinary  metabolites of mice to be  TCA  (52 percent), oxalic acid
(11 percent) with traces of dichloroacetic acid, and the remaining radioactivity
(37 percent) in unknown form.  More recently, Dmitrieva (1967) and Pegg et al.
(1979) have  confirmed  that oxalic acid  is a major  urinary metabolite of rats,
and, indeed, in this species, may be the principal  metabolite.   As yet, oxalic
acid has not been determined to be a metabolite of man, for whom TCA is pre-
sently assumed to be  the principal metabolite.  Buben  and 0'Flaherty  (1984)
observed no  excess  oxalic acid excreted by mice after chronic administration
of  large oral  doses of PCE,  and  these workers  suggest that oxalic acid is not
a significant metabolite in this species.
     The identification  of trichloroethanol  (TCE)  in the urine of humans has
been described by some authors (Ogata et al. ,  1962; Tanaka  and  Ikeda, 1968;
Ikeda and Ohtsuji, 1972; Ikeda et al., 1972).   However, in all of these studies,
the method  used  for TCA metabolite quantitation (Fujiwara reaction) is based
on  color  production before and  after  oxidation with chromium trioxide and
addition of pyridine.   The  difference  between the color production before
(TCA) and  after  oxidation  is  considered an  estimate  of  TCE content.  Sakamoto
(1974), from comparative  urine analysis by  GC  and  the  Fujiwara  reaction  under
different  pH and temperature  conditions,  expresses doubt  that the entire
fraction detected by Fujiwara reaction in the oxidation fraction is truly TCE.
Nonetheless, Monster et al. (1983) and also Weicherd and Lindner (1975) identi-
fied by GC  small  amounts  of TCE  (<4 |jmol/mmol  creatinine)  in the  urine  of
persons exposed to 10 to 30 ppm (68 and 204 mg/m ) PCE  in air, although others
using GC have  not in controlled experimental exposures to pure PCE (Hake and
Stewart, 1977; Fernandez et al., 1976; Monster et al.,  1979).  For mice, Yllner
                                    5-27

-------
(1961) reported that  by  his  chromatographic method  TCE  was not detected  in
urine as  a  PCE metabolite.   Daniel  (1963)  reported similarly for  the  rat,
-using steam distillation combined with isotopic dilution methodology.
     Buben and 0'Flaherty (1984)  also were unable to find  evidence  of TCE  in
the urine of chronically dosed mice as analyzed by gas chromatography; TCA  was
the only  metabolite found.  Furthermore, Costa  and  Ivanetich (1980) could not
detect TCE as  a product  of PCE metabolism by pre-induced rat liver microsomal
preparations i_n vitro.
     In mammalian  systems, oxalic  acid is known  to  be  converted  in part to
carbon dioxide, which has been identified in exhaled air of the  rat and  mouse
after exposure to  14C-labelled PCE  (Table 5-4).  Dmitrieva (1967) has also
reported  substantial  amounts  of  ethylene glycol  in  urine  of rats  exposed to
PCE,  although  this observation has not been confirmed  by others  (Buben and
0'Flaherty, 1984).   As no complete balance study of end-metabolites  of PCE  has
been  reported  for  the rodent  or man,  it  is  likely that all  metabolites of PCE
have  not  been  identified in these  species.  At  present,  there is insufficient
evidence  for  assuming a  qualitative species  difference in  PCE  metabolism.

5.4.2  Enzymic Pathways of Metabolism
     The  pathways  of  PCE metabolism are speculative.  The currently accepted
pathway for the  production  of TCA from  PCE  is  shown in Figure  5-13.  This
pathway was initially proposed by Powell  in 1945 for trichloroethylene and was
subsequently supported for PCE by the results of Yllner  (1961),  Daniel (1963),
Liebman and Ortiz  (1970,  1977),  Costa and  Ivanetich (1980) and  others.   The
first step in this  pathway appears to be catalyzed by hepatic cytochrome P-450
system to give 1,1,2,2-tetrachloroethylene  oxide, although  formation of this
epoxide has not been  rigorously verified by chemical  isolation or  identifica-
tion as an  intermediate  to  TCA formation in microsomal  reactions.   However,
the evidence is  strong as reviewed  below.  The epoxide  has been chemically
synthesized (Frankel  et  al.,  1957;  Bonse et al. , 1975;  Kline et al. ,  1978).
In sodium phosphate buffer  at pH 7.4, the oxide hydrolyzes  at a pseudo-first
order rate of  6 x  10    min'1  (T^-12 min) to TCA, and thermally  decomposes  in
nonaqueous solution to trichloroacetyl chloride (Table 5-4).   Sakamoto (1976)
also synthesized the  epoxide  in  milligram amounts,  essentially  by the method
of Frankel et al.  (1957), and  determined urinary metabolites after  i.p.
                                    5-28

-------
      ci          ci
        \      /
          c =c

      CI /         CI
         NADPH, O2
           P450
                        CI
                                            O
\/  \
           CI
                                      CI
                                  \
                            EPOXIDE

                           HYDRASE
                     H20,
            O            O

                 c - c
              /      \
            CI           OH
             DECARBOXYLATION
                     V2H20


                             O



                            'OH


                             O




                            ~OH



                      OXALIC ACID
           CO2 +
     HC
                         O
                        'OH
                                                   CI
                                         CHLORIDE

                                        MIGRATION
                                                                O
                                                                CI
                                                               H2O
                                                                   O
                                                                   OH
                                                                  TCA
                                                                        +H  CI
Figure 5-13.
Postulated  scheme for the metabolism of PCE.
1979; Costa and Ivanetich, 1980).
                       (After Pegg et al.,
                                      5-29

-------
 injection  into  guinea pigs.   By GC analysis, TCA was the only metabolite; the
 Fujiwara  reaction,  however,  indicated the  presence  of a small amount of TCE
 which  this  investigator believed to  be  an  artifact of  the  Fujiwara method.
     The  enzymatic  or spontaneous nonenzymatic steps in PCE conversion to TCA
 does  not  explain other reported  metabolites  of PCE, particularly  the  well-
 established  urinary  metabolite,  oxalic  acid  (Table 5-7).   Furthermore,  a
 significant portion of  PCE is completely metabolized to  C02  in a  dose-dependent
 manner (Tables  5-3, 5-4, and 5-6).
 5.4.2.1  P-450 oxidation.   That  PCE  is  metabolized by hepatic  microsomal
 cytochrome  P-450 system is strongly  supported  by  several  observations.   This
 enzymic system is known to metabolize analogs  of PCE, i.e., vinyl chloride,
•vinylidene  chloride,  and trichloroethylene  (Bolt et al., 1982).  PCE has been
 shown  to  bind to the  active  site of the enzyme system, as  evidenced by  the
 production  of a Type  I difference spectrum  in hepatic microsomes from rats jjn
 vitro  (Pelkonen and  Vaino,  1975;  Costa  and Ivanetich, 1980).   Costa and
 Ivanetich  (1980)  and  Liebman and  Ortiz (1977) demonstrated with rat microsomes
 j_n  vitro  that PCE also stimulated  nicotinamide adenine dinucleotide reduced
 (NADPH) oxidation which is characteristic of cytochrome P-450-dependent reac-
 tions.  Furthermore,  TCA was produced during  the aerobic incubation of  hepatic
 microsomes  (0?;  NADPH  required) in  a  time-dependent  manner (Figure  5-14).  The
 KM  (1.1 mmol)  and V    (0.046  nmol  TCA/  min/nmol  P-450) were respectively de-
  I I                  IllctX.
 creased and increased  with  microsomes  from  rats  treated  with inducers  of
 cytochrome  P-450 (Table 5-8).   With respect  to  phenobarbital  induction,  the
 data  of Table 5-8 suggest the  induction of one iso-form of  cytochrome P-450
 that  has  a greater PCE binding equilibrium and also  a  greater reactivity,
 i.e.,  a greater ability to metabolize PCE by both  increasing 2.5-fold hepatic
 microsomal content of  P-450 and also  increasing 4-fold average P-450  reactivity.
 Finally,  inhibitors of cytochrome  P-450 (SKF 525A,  metyrapone, CO) added to
 hepatic microsomes  with PCE  inhibited both  spectral  binding  and metabolism of
 PCE to TCA.
     The equilibrium  constant for the binding of PCE to  cytochrome  P-450  (1C =
 0.43 mmol)  is  similar to that  of trichloroethylene  (0.4 mmol).   However,  the
 maximum rates  of uninduced metabolism (V    ) of  PCE by the  cytochrome P-450
                                         ni3x
 system are  approximately  30-fold lower than  the  corresponding rate  for  the
 metabolism of  trichloroethylene (0.046  versus  1.55  nmol TCA/min/nmol P-450)
 (Table 5-8; Costa et  al., 1980).
                                    5-30

-------
                INDUCER

             PHENOBARBITAL
             PREGENOLONE-16a-CARBONITRILE
             UNINDUCED
             fl-NAPHTHOFLAVONE
                                     TIME, min
Figure 5-14.
    TABLE 5-8.
Production of TCA  from  PCE  by  hepatic  microsomes  from different-
ly pretreated  rats as  a  function  of time.   Incubation mixtures
contained  in  0.2 M  Tris-HCl  (pH 7.4),  tetrachloroethylene
(3.3 rtiM).  NADPH-generating systems,  EDTA (0.2 mM), and hepatic
microsomes (2 mg  protein/ml) from uninduced (triangle), p-naph-
thoflavone-induced   (diamond),   pregnenolone-16a-carbonitrile
(square)  and phenobarbital-induced  (circle)  male  Long-Evans
rats.  (From Costa et al.,  1980).

  METABOLISM OF PCE  BY  RAT  HEPATIC MICROSOMES AND THE EFFECT
                OF VARIOUS  INDUCERS
   carbonitrile
Phenobarbital
                           P-450  content
                             nmoles/mg
                           microsomal
           2.32 ±  0.32
0.2 ± 0.1
                                                        max
                                                 nmoles TCA/min/
Inducing Agent
None
p-naphthoflavone
Pregnenolone-16 a
fjr u ie i ii
0.89 ± 0.06*
1.03 ± 0.01
1.50 ± 0.11
(mM)
1.1 ± 0.8
0.5 ± 0.5
2.4 ± 0.8
nmole P-450
0.046 ± 0.004
0.055 ± 0.014
0.16 ± 0.09
0.19 ±  0.02
*Mean ±S.D.; measured in presence  of  3.3  mM PCE.

Source:  Costa et al., 1980.
                                    5-31

-------
5.4.2.2  Other Pathways.   The occurrence  of oxalic acid and  of  C02  as  major
metabolites of PCE, at least  in rodents (Table 5-7),  indicates the existence
of pathway(s) of metabolism other than the primary TCA pathway.  Assuming the
primacy of P-450 oxidation to PCE-oxide as the initial step in the metabolism
of PCE, as outlined above,  it has been proposed  by  Pegg et al.  (1979) that
PCE-oxide, after conversion  to chloroethylene glycol  by microsomal  hydrase,  is
dehydrohalogenated to tetrachlorodiacetyl chloride and then to oxalic acid or
CCL and formic acid (Figure 5-13).  Liebman and Ortiz (1977)  attempted  a  test
of the  postulated PCE-oxide  to  diol  step by adding  inhibitors  of hydrase
activity  to  microsomal preparations metabolizing  PCE.   These  investigators,
however, did not observe  the expected  increase in  the  PCE-oxide to TCA pathway
by the addition of cyclohexene to rat  microsome-PCE incubations to inhibit the
postulated epoxide-diol  pathway.   Nonetheless, these  experiments were inconclu-
sive, particularly as the data did not include a control  demonstrating hydrase
activity  in  their  microsomal  preparations,  and the effect of the cyclohexene
inhibitor with a  known  substrate.   According to Oesch (1972), there are  two
types of  epoxide  hydrase, one of which is tightly coupled to  the cytochrome
P-450 system  and  one  of  which is not; the coupled form is thought to be more
important  in detoxifying epoxide formed  by  mixed-function  oxidases and  is
relatively resistant to  inhibitors.
     The  postulated pathways  of  PCE  metabolism (Figure 5-13)  do not include
conjugation  reactions as possible metabolic detoxification steps.  Presently,
there  is  little evidence that  PCE-oxide might conjugate with glutathione
(GSH), for example, as a detoxification pathway as postulated for vinylidene
chloride-oxide (Jones and Hathway, 1978;  Reichert et al. , 1979).  Pegg  et al.
(1979) found  that  inhalation  exposure of  rats to 600 ppm (4068 mg/m3) PCE for
6 hours had  no  significant  effect on  total  hepatic nonprotein sulfhydryl, an
estimate of  hepatic GSH  content.   GSH concentration  was estimated as 5.06 ±
0.02 and  4.93 ± 0.21 umol/mg  liver for  control  and exposed animals, respec-
tively.   The  pathways for the origin  of  major metabolites of  PCE other than
TCA (Table 5-7) are deserving of further attention and investigation.

5-4.3  Magnitude and Dose-Dependency  of Metabolism
     Experimental  human  and  animal exposures to PCE have been carried out with
an overall finding that PCE metabolism is very limited,  i.e.,  that saturation
of mammalian  capacity to  PCE occurs   at relatively  low levels of exposure.
                                    5-32

-------
5.4.3.1  Man.   Estimates  of the extent of metabolism in man, as a percentage
of retained  PCE, have  been made from balance  studies  by  accounting for a
retained dose  after  inhalation exposure by measuring metabolites (TCA, total
trichloro-compounds  by  Fujiwara reaction)  excreted in  the  urine (Stewart
et al., 1961, 1970; Monster et al. , 1979; Monster and Houtkooper, 1970; Monster
et al., 1983;  Boettner  and Muranko, 1969; Ikeda et al., 1972; Essing et al.,
1973;  Fernandez  et al.,  1979;  May,  1976).  Only  about  1 to  3 percent  of  the
estimated amounts  absorbed are  metabolized  into  TCA and  other  chlorinated
metabolites.    Ikeda  and co-workers  (Ikeda  et  al.,  1972;  Ohtsuki  et al. , 1983)
have explored the relationship of inhalation exposure concentration in occupa-
tional settings (time-weighted 8-hr average) metabolite concentration in urine
(total trichloro-compounds;  Fujiwara reaction) at  the  end of an 8-hr work
shift. Under these  conditions,  these  investigators  found a saturation  of
                                                           •3
metabolism occurs between  100  to 400 ppm (678 to 2112 mg/m ) PCE in inspired
air  (Figure  5-9).  They estimated that  at  the  end  of  an  8-hr  shift with expo-
                                3
sure  to PCE  at 50 ppm (339 mg/m ) (below saturation; Figure 5-9), 38 percent
of PCE pulmonary  uptake would  be exhaled  unchanged,  and less than 2 percent
would  be metabolized to urinary chlorinated compounds,  while the rest would
remain in the body to be eliminated in succeeding hours.
     The difficulties associated with balance  studies in man are the problems
encountered  with  (1)  accurate  measurement of the retained dose  of  PCE from
inhalation exposure,  (2) the imprecision  of the older methodologies using the
Fujiwara reaction for metabolite quantification,  (3) the  possibilities that
metabolites  other than TCA  (e.g., oxalic acid, TCE-glucuronide)  may be excreted
in urine or.by biliary  excretion, and (4)  the  very  long  half-life (144 hours)
of PCE urinary metabolites  which necessitates  extended collection of samples.
For  example, Bolanowski  and Golacka (1972) in  their  series  of  experimental
exposures to PCE  in  man found  that  the  elimination  of TCA  in the urine repre-
sents  only a few percent of the dose while the alveolar retention of PCE was
60 to 80 percent (Fernandez et al., 1976; Monster et al.,  1979).   This suggests
a total metabolism of 20 to 40  percent  of  the dose.   Thus they assume that
another unrecognized  pathway exists  for  PCE which  has not  yet been taken  into
consideration.   Conversion of PCE to COp could constitute  a significant portion
of this metabolism.
5.4.3.2  Animals.  For rodents (rat and mouse), the extent of metabolism after
single doses of PCE  has been estimated by balance studies using isotopically
labelled PCE.   Yllner  (1961)  carried  out the earliest  study  in mice.   By
                                    5-33

-------
exposing the mice by inhalation to   C-PCE vapor,  he  found  that  of  the  radio-
active PCE absorbed  (2  hr-exposure;  0.9 mg/kg b.w.), 70 percent was excreted
unchanged in expired air,  2 percent was metabolized and excreted in the urine,
and 0.5 percent  in  the feces  over a four day  post-dosage  period.   Daniel
(1963) gavaged rats  with  36C1-PCE  (about 1.12  and 8.29 mg/g  b.w.)  and  found
that 98 percent of  these  doses were excreted unchanged by the lungs and only
2 percent of radioactivity was metabolized and  excreted in  the urine over  a 2
to 18-day period (Table 5-9).   These early observations, which indicate a very
limited metabolism  of  PCE by  rodents,  are  in accord with  the  more recent
kinetic studies of  Filser and  Bolt (1979).   These investigators exposed rats
to PCE in a closed system  and measured the rate of disappearance of PCE vapor.
The kinetics of metabolism  of  PCE  was dose-dependent and the estimated M
                                                                         (TldX
was less  than  1.2 mg/kg/hr.   Filser and Bolt noted that PCE was metabolized
extremely slowly in  rats  (in comparison to other  halogenated ethylenes) and
they were unable to  experimentally differentiate between zero order and first
order  kinetics and  hence  to  determine the saturation point, i.e.,  ppm PCE in
air at which metabolism "saturation" occurs.
     Other investigators have also  found evidence of dose-dependent Michaelis-
Menten kinetics for  metabolism of PCE in mice and rats following both oral  and
inhalation exposure.  Pegg et  al.  (1979) exposed rats to-single oral doses of
14
  C-PCE  (1  and 50 mg/kg  b.w.) and to single inhalation exposures  (10  and
600 ppm for 6  hours) and  followed the disposition  of  radioactivity  in exhaled
air, urine,  and feces for  3 days post administration.   Their data are given in
Table 5-3.  For these  low doses  (1 mg/kg and 10 ppm, 6 hours),  71.5 and 68.1
percent, respectively,  of the  body burdens  were excreted  as unchanged PCE in
exhaled air.   The  remainder,  28.5 and 31.9 percent, was metabolized to C02 and
fecal  and urinary  metabolites.   In  contrast,  for the higher oral  and inhalation
exposure doses (500  mg/kg and  600  ppm,  6 hours), a greater percentage of the
body burdens was  exhaled as unchanged PCE (89.9 and 88.0 percent, respectively)
Hence, these data  demonstrate that  even at very low body burdens of PCE in the
rat (approximately 0.25 to 2.5 mg per animal) only 30 percent is metabolized,
and this  percentage  decreased  with increasing  body  burden.   This  indicates
that metabolic disposition of  PCE  is a saturable, dose-dependent process  in
this species after  either  oral (gavage) or inhalation exposures.   Recently,
Frantz and Watanabe  (1983) have demonstrated that PCE from saturated solutions
of drinking  water,  allowed ad  libitum over a 12-hr period  to rats, provided
                                    5-34

-------
       TABLE  5-9.   DISPOSITION OF 36C1-PCE AFTER ORAL ADMINISTRATION TO
                                  WISTAR RATS
Dose
(uc)***
1.75
13
Mg/Kg
Equivalents
1116
6160
No.
Animal s
1
4
Expired air*
(unchanged)
PCE
97.9
--
Metabol
Urine**
2.1
1.6
ized
Feces
0
0
  "Collected for 48 hr.
 **Collected for 18 days.
***Specific activity of 36C1-PCE was 1.3 uc/mmol.
Source:   Daniel, 1963.

average body burdens of  8.09 ± 3.13 mg/kg (2.3 mg per animal),  and gave com-
parable results to  low  dose (1 mg/kg) gavage dosage, i.e.,  88 percent of the
body burden of PCE was  excreted unchanged in exhaled air and only approximately
12 percent was metabolized.   Their results are shown in Table 5-6.
     Schumann et al. (1980)  have investigated the extent of metabolism in mice
using   Olabelled  PCE.   The  experimental  conditions  were  very similar to
those of  Pegg et  al.  (1979) and Frantz  and Watanabe (1983)  in  rats as these
experiments  originated  from the  same laboratory.   Schumann and co-workers
determined the disposition in mice of body burdens of PCE of 0.04 mg per mouse
(from 10  ppm,  6 hours  inhalation exposure) and  of 10.8 mg per mouse  (from
500 mg/kg gavage  dose)  (Table  5-4).   For the smaller body burden,  12 percent
was excreted  unchanged  in exhaled air and  the  remainder (about 88 percent)
metabolized; for  the  larger body burden, the proportions were  reversed  with
83 percent of the body burden excreted unchanged and about 17 percent disposed
of by metabolism.   These  data indicate that for mice also, metabolism of PCE
is dose-dependent and  saturable  for this  species.  When  compared to rats
(Tables 5-3  and  5-6 versus  Table 5-4), mice were found to metabolize 8.5 and
1.6 times more PCE per kg body weight following inhalation of 10 ppm (68 mg/m )
for 6 hours or a  single oral dose of 500 mg/kg respectively.  However, on a body
burden per animal basis, the rat metabolized 6.8 and 1.3 times more PCE  than the
mouse following  inhalation  of 10 ppm (68 mg/m ) for 6 hours or a single oral
                                    5-35

-------
dose of 500 mg/kg respectively.   Thus, from these observations, the metabolism
of PCE by these two species is more consistent with  a  metabolism  proportional
-to body surface area than to kg body weight.
     Buben and  0'Flaherty (1984) have also  investigated the extent  of  PCE
metabolism in male mice  (Swiss-Cox strain) using  urinary metabolite  excretion
as an index of metabolism.  These investigators daily dosed the mice  by gavage
with PCE  in corn oil, 5  days per week for 6 weeks, at  dose  levels of  20,  100,
200, 500, 1000,  1500,  and 2000 mg/kg/day.   TCA was the only metabolite found
in the urine  of  this  strain with less than 5 percent of the dose excreted  in
the feces.  Nonetheless,  total  urinary metabolite as an index of metabolism is
only a dose approximation since  it is known that  some  metabolized compound  is
bound to  cellular macromolecules  and  some is metabolized to CO^  (Tables 5-3,
5-4, and 5-6).  However,  urinary excretion is by far the major route  of disposi-
tion of  metabolized PCE.  Therefore,  total  urinary  metabolite should be a
reasonable approximation  of the amount  of  metabolism.  Also, with  chronic
administration, body burden of  PCE  is expected to reach steady-state, and  in
fact these investigators  found  no  week-to-week trends in daily urinary meta-
bolite output although urinary  TCA  tended to increase during the week with
plateauing towards the end  of  the week; urine collections  were made at the
weekly plateau.  Figure  5-12  shows  the relationship found between the amount
of metabolism and the administered dose.   The data clearly show a dose-dependent
metabolism consistent with capacity limited kinetics in accordance with Michaelis-
Menten equation, Y = M    - X/K  +  X,  where X is  the  dose  (mg/kg) and Y is
                      maX      ITS
mg/day of metabolite resulting.  From the computer-fitted equation,  Buben and
0'Flaherty estimated Mmo^ (maximum  rate  of metabolite formed and excreted  in
                      INclX
24 hr) to be  136 mg/kg/day  and Km (the dose at which the amount of metabolite
excreted  in a 24-hr  period  is half the  apparent  maximum amount) at  660  mg
PCE/kg.   These workers also estimated that at a  very  low chronic oral dose
of PCE only 25  percent was  metabolized (indicating significant hepatic flow-
limited or first pass  effect)  and less  than  5  percent of a very high dose.
These observations also  indicate that,  in this strain of mice  at least,  the
metabolism of PCE deviates  from linearity (i.e.,  from first order)  at oral
doses above 100 to 200 mg/kg.  Of considerable importance also  is the finding
of a clear relationship  between toxicity and metabolism of PCE.  Figure 5-15
shows a direct linear relationship between hepatotoxicity (serum  SGPT activity)
and urinary metabolite excretion.   Similar relationships were  observed  with
                                    5-36

-------
other indices  of  hepatotoxicity such as liver weight increase, liver trigly-
ceride accumulation,  and  glucose-6-phosphatase activity.  These observations
are consistent with  the formation of a  toxic  or  reactive intermediate(s)  in
the primary metabolic pathway of PCE (Figure 5-13).

5.4.4  Covalent Binding
       PCE is known to bind to the substrate binding site of hepatic microsomal
cytochrome P-450  i_n  vitro (Costa et al.,  1980).   A direct  relation  also  has
been  demonstrated among  the  level  of  hepatic microsomal cytochrome  P-450,
the extent of  metabolism  of PCE i_n  vivo,  and  cellular  damage  (Bonse  et al.,
1975; Moslen et al., 1977; Pegg et al., 1979; Schumann et al. , 1980; Buben and
0'Flaherty, 1984).   For example, Moslen  et al.  (1977),  in a study  designed to
correlate metabolism  with hepatocellular toxicity,  observed that induction of
microsomal systems in rats by phenobarbital or chlorinated biphenyls increased
metabolism of  orally administered  PCE  (5.7-fold) and increased indices of
toxicity  (2-fold).   Buben and 0'Flaherty  (1984)  have also presented evidence
for a strong  correlation  between hepatocel1ular  toxicity and metabolism.   As
shown in Figure 5-15, these workers found  in mice linear relationships between
the amount  of metabolism  of PCE (as measured  by urinary metabolites)  and
hepatotoxicity as  measured by several  indices  of toxicity.   In terms of the
proposed  metabolic pathway(s)  for  PCE  (Figure 5-13),  covalent  binding  to
cellular macromolecules,  leading  to cellular damage, may occur with putative
reactive intermediates of  PCE metabolism:  PCE-oxide, trichloroacetyl chloride,
and chlorodiacetyl  chloride.   Bonse  et al.  (1975) and Bolt  et  al.  (1982) have
proposed that  the symmetrical  PCE  oxide  metabolite  is much  less reactive than
the epoxides of unsymmetrical chlorinated  ethylenes such as trichloroethylene,
vinylidene chloride,  and  vinyl  chloride.    Consequently,  covalent  binding  of
PCE metabolites to cellular macromolecules may result in less severe hepatic
toxicity and genotoxicity, even when taking into  account the large differences
in the  hepatic abilities  to  metabolize these  compounds  (Section 5.4.3).  This
correlation between  biological  activity  and chemical stability has been con-
firmed by Jonas and  Mackrodt  (1982), who studied  the  relationship  between  the
mutagenic activity  of chlorinated  ethylenes and  the C-0 bond energy of their
epoxides.
     Bonse et  al.  (1975)  perfused isolated rat  liver  preparations with  PCE
(180 ppm  concentration)  and observed that  only  3 to 5 percent of the total
uptake bound to  liver tissue; this portion was  extractable only after acid
                                    5-37

-------
                                40            60

                         TOTAL URINARY METABOLITE, mg/kg
                                            80
100
Figure 5-15.
Relationship between  hepatotoxicity parameters  from PCE oral
administration  (shown  as serum  SGPT activity increases)  and
total urinary metabolite excreted  per  day by mice at increasing
dose levels  of  PCE.  The slope of  the  linear regression line is
              0.436 with  a coefficient of
              Buben and 0'Flaherty, 1984.
                              regression, r2, of  0.949.   From
                                    5-38

-------
hydrolysis.  They proposed that the acyl chloride intermediate in the conversion
to TCA (10 to 15 percent of uptake) reacted with cell functional groups, e.g.,
-OH, -5H,  and -NH2-   Pegg  et  al.  (1979)  have  demonstrated  i_n  vivo  rat  hepatic
binding after exposure  to  PCE by  both  inhalation  and oral  routes.   Their data
are shown  in Table  5-10.   Radioactivity irreversibly associated with  tissue
macromolecules  was  assayed 72 hr after termination  of inhalation exposure to
10 or  200  ppm (68 or  1356  mg/m3)  for 6  hr  and 72  hr  after  oral  administration
   14
of   C-PCE.  Irreversible  binding (primarily to cellular protein and nucleic
acid)  was  dose-dependent  and  metabolism-dependent,  but independent of the
route  of  exposure.   The ratio of  amount  bound to total amount metabolized
(about 1 percent; Table 5-10) was not significantly  different between high and
low doses  of PCE by either route  of exposure,  suggesting  that irreversible
binding was solely  a  function of metabolism and the availability of reactive
species.    The  persistence of  irreversible  bound  metabolites  of PCE (72 hr
after  exposure)  indicates  that turnover rates of the bound metabolite(s)  are
slow enough that accumulation could potentially occur  in  man with repeated
daily exposures.

         TABLE  5-10.  IRREVERSIBLE HEPATIC BINDING OF 14C-PCE IN RATS
                             72 HR AFTER EXPOSURE

Route
Oral
Inhalation
(6 hr)

Exposure
1 mg/kg
500 mg/kg
10 ppm
600 ppm
Amount
Metabol ized
umol eq
0.41 ± 0.03*
73.92 ± 4.98
2.82 ± 0.16
55.46 ± 0.44
Bound
(jmol eq/
1 i ver protein
0.0035 ± 0.0003
0.4321 ± 0.0477
0.0245 ± 0.0022
0.4029 ± 0.0222
Bound/
Metabolized
x 100
0.83
0.59
0.87
0.72
*Mean ± S.E.
Source:  Pegg et al.,  1979.
     Schumann et al.   (1980)  have compared metabolism of PCE and irreversible
hepatic binding in the  two species, rat (Sprague-Dawley) and mouse (BgC^F^).
Their data are  summarized  in Table 5-11.  The  time  course  of  hepatic macro-
molecular binding  differs  between these two species; the peak binding occurs
in mice  significantly more  rapidly than in rats  after  exposure from both
                                    5-39

-------
  TABLE 5-11   COMPARISON OF IRREVERSIBLE HEPATIC BINDING OF  14C-PCE  IN SPRAGUE-DAWLEY
                    RATS AND BgCsFi MICE AFTER  INHALATION (6  HR)  AND  ORAL EXPOSURES
|jg mol-eq bound/ g hepatic protein
Time,

hr post-
exposure

0
1
6
12
24
48
72
A.
1.604

1.355

1.027
0.729
0.481







600 ppm
Mice
± 0.251*
-
± 0.120
-
± 0.138
± 0.232
± 0.127
B.
0.174

0.227

0.279
0.246
0.171
Rats
± 0.064
-
± 0.026
-
± 0.017
± 0.062
± 0.050
A/B
9.2
-
6.0
-
3.7
3.0
2.8
A. Mice

0.
2.
1.
1.
1.
0.
-
964 ±
171 ±
687 ±
386 ±
081 ±
472 ±

0.115
0.523
0.112
0.262
0.290
0.131



500 rag/kg
B.

0.170
0.331
0.402
0.612
0.347
0.321
Rats
-
± 0.025
± 0.027
± 0.086
± 0.027
± 0.060
± 0.060
A/B
-
5.7
8.2
4.2
2.3
3.1
1.5
*Mean ± S.D.
Source:   Schumann  et  al. ,  1980.

      inhalation and  oral  routes of  administration.   There is 3 to  5 times more
      binding per  gram  hepatic  protein  in mice  than  rats.   Since  liver weight of  these
      two species  is  proportional to species  body weight, the total  hepatic-bound
      PCE is  comparable to  the metabolism of  PCE by  the mouse and  rat  (Section
      5.4.3.2).
           Schumann et  al.  (1980) also  have tried to evaluate the extent of binding
      to DNA and  its relation  with the  probability  of genotoxicity.   These workers
      utilized the mouse in  their experiments  because  of greater  binding  per gram of
      liver in this  species, and because  of the susceptibility of this species to
      hepatic tumors  (NCI, 1977).  These  investigators did not find  a detectable
      binding of high specific activity 14C-PCE (1.3  x 105 dpm/nmol) to liver DNA
      when  mice were  exposed by  inhalation  (600 ppm)  or by gastric  intubation (500
      mg/kg).   However,  the limit of detection was  estimated as  10  to 14.5 alkyla-
      tions  per 10   nucleotides.   Therefore,  levels  of DNA binding  below their
      detection limit may  have occurred.   In  parallel experiments, these  investi-
      gators  (Schumann  et  al.,  1980;  Watanabe et al.,  1980) also chronically admin-
      istered PCE  (11 to 12 daily doses)  by gavage  at increasing dose levels up  to
      1000  mg/kg (Table 5-12).   Mice, but not rats,  showed a 20-25 percent  increase
      in liver/body  weight ratio.  DNA/g  liver protein decreased, indicating that

                                          5-40

-------
       TABLE 5-12.  EFFECTS OF CHRONIC ORAL ADMINISTRATION OF  PCE ON  HEPATIC DNA
                    CONTENT AND DNA SYNTHESIS  IN MICE  AND RATS3
Liver DNA content
(mg DNA/g liver)
Dose
(mg/kg/day)
12 doses in 16 days
0
500
11 doses in 11 days
0
100
250
500
1000


4.00
3.31

2.51
2.29
2.25
2.23
2.30
Mice

± 0.42
± 0.22*

± 0.11
± 0.10*
± 0.17*
± 0.15*
± 0.10*
Rats

2.22 ± 0.29
2.03 ± 0.17

2.33 ± 0.19
2.24 ± 0.22
2.51 ± 0.16
2.36 ± 0.24
2.18 ± 0.27
Liver DNA synthesis
(dpm [3H]thymidine/Mg DNA)
Mice

65 ± 34
118 + 22*

44 ± 15
56 ± 27
63 ± 26
75 ± 23
82 ± 31*
Rats

63 ± 11
65 ± 17

88 ± 24
46 ± 26
137 ± 103
105 ± 45
108 ± 45
aMean ± SD  (n = 3-7);  [3H]thymidine  1000 pCi/kg  i.p. 6  hr prior to sacrifice.
*
 p <0.05.
Source:  Schumann et al.  (1980).

    organ weight gain  resulted  from  hypertrophy  rather  than  hyperplasia.  Nonethe-
    less,  DNA  synthesis (  C-thymidine incorporation)  tended  to  increase in both
    mice and rats in a  dose-related  manner,  suggesting  some  activation  of "repair"
    mechanisms.

    5.4.5   Interactions with Metabolism
    5.4.5.1  Induction and Inhibition.   The  metabolism of  PCE  is known to be
    increased  by  inducers of the  cytochrome P-450 oxidative  system  (microsomal
    mixed-function  monooxygenases).   Moslen et al. (1977) demonstrated that the
    metabolism  of  PCE  in  intact  rats,  as  measured by  urinary metabolites,  was
    enhanced  (4 to 7-fold)  by  pretreatment with  inducers of  cytochrome P-450,
    phenobarbital  and  polychlorinated  biphenyls, which  increased the  hepatic
    cytochrome  content 2.4 and  3.3-fold,  respectively.   Similar results have been
    reported by Ikeda and Imamura (1973)  for both intact rats and hamsters.  Costa
    and Ivanetich (1980)  found  that  pretreatment of rats with  the inducers,  pheno-
    barbital and pregnenolone-16orcarbonitrile,  increased cytochrome  P-450 content
    of microsomes  and markedly enhanced the metabolism of  PCE to TCA  by these
                                        5-41

-------
microsomes i_n vitro (Figure 5-14).   Inhibitors  of  cytochrome  P-450,  SKF 525A,
metyrapone, and  CO,  added to the microsomal  reaction medium, significantly
decreased the production of TCA.
     Chronic exposure  to PCE appears  to  produce minimal  self-induction  of
metabolism.  Kaemmerer et al. (1982) report that feeding  rats 25 and 200 ppm
PCE in their diet resulted in a small  but statistically  significant  induction
of cytochrome P-450  in  rat liver homogenates after  14 and 7  days  treatment,
respectively.   Vainio and co-workers  (Savolainen et  al.,  1977; Vainio  et  al.,
1976) found a slight  depressing  effect on  rat  liver P-450 content 24  hours
after a single high oral  dose of PCE (1700 mg/kg) but reported slight increases
                                                     3
after PCE  inhalation  exposure at  600 ppm (4068 mg/m ),  6 hours  daily  for 5
days.   Plevova et al.  (1975)  and also Kaemmerer et al. (1982) observed  that in
                                    3
rats,  exposure to 177 ppm (1200  mg/m ) PCE for 6 hours, 20 hours prior  to i.p.
administration of  pentobarbital  (44  mg/kg),  lengthened  by  30  percent the
pentobarbital  sleeping time.   These observations suggest that substrate competi-
tion for  cytochrome  P-450  system  for the metabolism of pentobarbital and for
PCE may occur,  with a consequent increase of rate of metabolism of phenobarbital
and prolongation of its half-life.
5.4.5.2   Drugs.   In  addition  to the potential for PCE interactions with
clinical  drugs,   some of which are known to  induce  P-450  system or  compete for
P-450 oxidative  metabolism,  other types of interactions  are  possible.   For
example,   Hake and  Stewart (1977)  evaluated the effect  of diazepam  and of
alcohol,  CNS depressants,  on  human  exposure to PCE.   During an 11-week study
with inhalation  exposure to PCE, 5.5  hours  per  day at 25  and  100 ppm  (169 and
678 mg/m ), concomitant administration  of  ethanol  (blood  levels 30 to  100 mg
percent)  increased PCE blood  levels in the 25  ppm exposure but not with  100
ppm; diazepam (blood  levels 7 to 30  ug percent) had  little effect.  Addition-
ally,  behavioral and neurological tests revealed no  interactive effects under
these conditions.  However, May  (1976)  found  that  urinary excretion  of  TCA
was completely inhibited by a blood  ethanol concentration of  0.65  mg percent.
5.4.5.3  Other Interactions.   PCE has been observed to have a significant
effect on  intermediary metabolism.   Takano  and Miyazaki  (1982) and Miyazaki
and Takano (1983) reported that  the  PCE inhibited glutamate, succinate,  and
malate oxidation by rat liver mitochondria.   The lesser inhibition of succinate
oxidation suggested to the authors  that PCE  may act as an uncoupler in the
electron  transport system  between  NADH dehydrogenase and  coenzyme Q.   Ogata
                                    5-42

-------
and Hasegawa  (1981)  have  reported  similar  findings  for PCE  and  other halocar-
bons for succinate oxidation by rat liver mitochondria.  Since both  the K  and
Vmax of the reactions were affected by PCE, it was suggested that PCE represents
a nonspecific  inhibitor (possible  mitochondrial  membrane perturbation by  this
lipid solvent)  rather  than a specific inhibitor like  cyanide or antimycin A.

5.5 SUMMARY
     Perchloroethylene is  rapidly  and virtually completely absorbed  from the
gastrointestinal tract of rats  and mice after  administration of doses  up to
500 mg/kg  by  intragastric intubation, and of doses up to 8 mg/kg in drinking
water, with peak  blood concentrations occurring within one  hour  of dosing.
Pulmonary uptake of PCE during inhalation exposure by  experimental  animals and
man is linearly proportional to exposure duration and  concentration  in air;  it
is also influenced by physical activity and body mass.  For rats exposed to  10
                             3
and 600 ppm (68 and 4068 mg/m ) for 6 hours, net uptake of PCE has  been measured
as 5.9 and 310 mg/kg b.w.; for man exposed to 72 and 144 ppm (488 and 976 mg/m )
for 4  hours,  the  net  uptake  averages  6.5 and 13.5 mg/kg.  Steady-state  uptake
in  man  is not  established within short (8 hours)  exposure  periods.  With
chronic exposure  (100  ppm;  8 hr/day;  5 days/wk), 3  to  4 weeks is required for
steady state.   Absorption of PCE during vapor  or liquid contact with the  skin
of  experimental animals  or man is very slow and adds  less than 1 percent to
the body burden.
     PCE distributes  widely  into body tissues,  and  readily crosses  the  blood-
brain barrier  and  the placental  barrier.   Highest  tissue concentrations  are
found in  the  adipose  tissue  (60 times blood level),  and in brain and liver (4
and  5  times  blood level,  respectively).   Tissue concentrations increase in
direct proportion  to  the  body burden.  With chronic exposure in man, adipose
tissue concentrations do  not achieve  plateau concentrations until 3  to 4 weeks
of exposure.
     PCE is  eliminated from the body principally by pulmonary  excretion of
unchanged  compound and also by a  rate-limited  metabolism of  PCE.   Pulmonary
excretion  in  man  and experimental  animals occurs in three first-order phases
of  desaturation of blood  vessel-rich  tissue group  (VRG), muscle tissues (MG),
and adipose  tissues  (FG).  For man,  the half-times of elimination  from these
groups are 12-16 hr, 30-40 hr, and 55-65 hr, respectively.   For rat,  the domi-
nant half-time of pulmonary elimination is about 7 hours.
                                    5-43

-------
     The capacity of man and rodents to eliminate PCE by metabolism is  limited.
For man, as assessed by urinary metabolite excretion, about 1-3 percent of the
                                                                            2
absorbed amounts of PCE from exposures of 100 to 400  ppm  (678  and 2112 mg/m  )
exposure for 8 hours are metabolized, with a half-time of renal elimination  of
the metabolites of about 144 hours.   Urinary metabolite excretion  increases  as
PCE body burden  increases  until  a plateau is reached at about 400 ppm  (2112
mg/m3) exposure  (500 to 600 mg body  burden).  Metabolism  is, therefore,  dose-
dependent  and  saturable in man.   Similar observations have  been made for
rodents.   For  the rat and mouse, the percentage of the administered dose  that
is metabolized decreases with  increase of dose.  Metabolism in both  the  rat
and mouse is consistent with saturable, dose-dependent Michaelis-Menten kinetics
with a  maximum rate  (V    )  in  the order of 1.6 mg/hr/kg b.w.  for the rat and
                       max
5.7 mg/hr/kg for the mouse.
     Indices of  hepatotoxicity of PCE  in the rat and  mouse  are highly correl-
ative with the dose-dependent  nature of the metabolism of PCE in these species.
The principal  site  of  metabolism appears to be  hepatic  P-450  monooxygenase
system where PCE is  oxidized to PCE oxide,  which rearranges to trichloroacetic
acid (TCA).  TCA  has  been  identified as a  major  urinary  metabolite  in man,
rat, and mouse.   Metabolism to TCA, i_n  vitro  and  i_n vivo, is  known  to be
increased  by inducers of  the  P-450 oxidative system.  Secondary  pathways of
metabolism may include hydration of PCE oxide with subsequent dehydrohalogena-
tion of the  glycol  to  chlorodiacetyl chloride and further  rearrangement to
yield oxalic acid or formic acid and carbon dioxide.   In  rodents,  oxalic  acid
is a urinary metabolite while  carbon dioxide is a pulmonary metabolite.  There
is no evidence that  conjugation  reactions with glutathione or cysteine are
important  in the  metabolism of PCE.   While the metabolite profiles of PCE in
man, rat,  and  mouse  are not as yet fully established, there is no convincing
evidence of  qualitative or quantitative  differences of  pathways in  these
species.
     PCE metabolite(s)  are  known to  covalently bind,  i_n vitro  and in  vivo, to
cellular macromolecules such as protein and lipids.   Tissue binding appears to
be solely  a  function  of metabolism and availability  of reactive metabolites.
Since tissue-bound metabolites  have  a  slow rate of  turnover and  accumulate,
cumulative cellular damage  may occur with chronic exposure.  Covalent binding
to hepatic DNA i_n vivo  is  minimal;  however, as  the  limit of  detection for
these experiments was only  10  to 14.5 alkylations/10  nucleotides, very  low
levels of DNA binding cannot be completely excluded.
                                    5-44

-------
5.6  REFERENCES
Barchet, et  al.  1972.  Chemical  investigations  for environmental protection.
     Deutsch  Lebensm - Rundsch.  68:  69.  Cited  in:  Chem.  Abstr. 77: 32869.

Boettner, E.  A.  and Muranko,  H.  J.  1969.  Animal  breath  data for estimating
     the exposure  of humans  to chlorinated hydrocarbons.  Am. Ind. Hyg.  Assoc.
     J.  30:  437-442.

Bolanowska,  W. and Golacka, J.  1972.  Absorption  and  elimination  of tetrachloro-
     ethylene in humans under  experimental conditions.  Mdcyncyna Pracy  23:  109-
     119. (English translation)

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

Bonse, G. ;  Urban,  Th. ;  Reichert, D.  and  Henschler,  D. 1975.  Chemical reac-
     tivity, metabolic oxiraine  formation and biological  reactivity  of chlori-
     nated ethylenes  in  the  isolated perfused rat liver preparation. Biochem.
     Pharmacol. 24: 1829-1834.

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

Costa, A. K.  and Ivanetich,  IK.  M.  1980.  Tetrachloroethylene metabolism  by
     the  hepatic microsomal  cytochrome  P-450  system.  Biochem.   Pharmacol.
     29: 2863-2869.
                                     oc
Daniel,  J.  1963.  The metabolism  of   Cl-labeled  trichloroethylene and tetra-
     chloroethylene in the rat.  Biochem. Pharmacol.  12: 795-802.

Dmitrieva, N.  V.  1967. Contribution to  the  metabolism of tetrachloroethylene.
     Gig. Tr. Prof. Zabol. 11:  54-56.  (English  translation).

Eger, E. I.  A mathematical model of uptake  and distribution. In: Uptake and
     distribution  of  anesthetic agents.  Ed.  , Popper,  E.  M. and  Ketz, R.  J.
     pp. 72-103, McGraw-Hill,  New York.

Essing,  H.  G. ;  Schacke,  G. ;  Schaller,  K.-H.  and Valentin, H. 1973.  Arbeits-
     Medizinische Untersuchungen  zur  Dynamik  des  Perchlorathylens im Organismus.
     Med. Welt. 24: 242-244.

Fernandez,  J. ;  Guberan,  E.;  Caperos, J.  1976. Experimental human exposures to
     tetrachloroethylene vapor  and  elimination  in breath  after inhalation.  Am.
     Ind. Hyg. Assoc. J.  37:  143-150.

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

Frankel, D.  M.; Johnson,  C. E.  and  Pitt, H.  M.  1957.  Preparation and properties
     of  tetrachloroethylene oxide.  J. Org. Chem.  22:  1119-1124.


                                     5-45

-------
Frantz, S.  W.  and  Watanabe,  P. G.  1983.  Tetrachloroethylene: balance  and
     tissue distribution  in male Sprague-Dawley rats by drinking water admin-
     istration.  Toxicol. Appl. Pharmacol. 69: 66-72.

Guberan, E. and  Fernandez,  J. 1974. Control of  industrial  exposure to tetra-
     chloroethylene by  measuring alveolar  concentration:  theoretical approach
     using a mathematical model. Brit. J. Ind. Med.  31: 159.

Hake, C. L. and  Stewart, R.   D. 1977.  Human exposure to tetrachloroethylene:
     inhalation and skin contact.  Environ.   Health  Persp.  21:  231-238.

Ikeda,  M.  and  Ohtsuji,  H.  1972.  A  comparative study on the excretion of
     Fujiwara  reaction  positive substances  in  the  urine of  humans and rodents
     given trichloro or tetrachloro derivatives of  ethane  or ethylene.  Brit.
     J. Ind.  Med. 29:  94-104.

Ikeda,  M.  and Imamura,  T. 1973. Biological  half-life of trichloroethylene and
     tetrachloroethylene in human subjects.  Int. Arch. Arbeitsmed.  31:  209-223.

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

Jones,  B.  K.  and  Hathway,  D. E.  1978.  The biological fate  of vinylidene
     chloride  in rats.  Chem.  Biol.  Interact. 20: 27-41.

Jones,  R.  B. and Mackrodt,  W. C.  1982.  Structure-mutagenicity relationships
     for chlorinated ethylenes:  a  model  based on  the  stability of  the metab-
     olically  derived epoxides. Biochem.  Pharmacol.  31: 3710-3713.

Kaemmerer,  K. ;  Fink, J.  and Kietzmann, M. 1982. Studies on  the  pharmacodynamics
     of perchloroethylene. Der prakitsch Tierarzt  2: 171-182.

Kline,  S.  A.; Solomojn,  J.  J.  and Van Duuren, B. L.  1978. Synthesis and  reac-
     tions of  chloroalkene epoxides. J. Org. Chem.  43: 3596-3600.

Kronevi, T. ; Wahlberg,  J.  F.  and Holmberg,  B. 1981. Skin pathology following
     epicutaneous  exposure  to seven organic solvents.  Int. J. Tiss.  React.
     3: 21-30.

Kundig, S.  and Hogger,  D.  1970. Die  Bedeutung  der  tri-  und  Perchlorathlen-
     mentaboliten im Urin. Int. Arch. Arbeitsmed.  26:  306-315.

Landry, T.  D. ; Ramsey,  J.  C.  and  McKenna,  M.  I.  1983. Pulmonary physiology
     and inhalation dosimetry in rats: development of  a method  and  two  examples.
     Toxicol.  Appl. Pharmacol. 71:  72-83.

Liebman, K. C.  and Ortiz,  E.  1970.  Epoxide intermediates in microsomal oxi-
     dation of  olefins  to  glycols.  J.  Pharmcol.  Explt.  Therap. 713:  242.

Liebman, K. C.  and Ortiz,  E.  1977.  Metabolism of  halogenated ethylenes.
     Environ.  Health Persp.  21: 91-97.

Ling, S.  and  Lindsay,  W.  A.  1971.  Perchloroethylene  burns.  Brit.   Med.  J.
     3:  115-116.

                                    5-46

-------
May, R.  Aufnahme,  Metabolisierung und Ausscheidung  von Tetrachlorathylen  und
     Beeinf1ussung  durch  Gleichzeitige  Aufnahme  von Athanol beim  Menschen.
     Dissertation, Universitat Wurzburg,  F.  R. G.

Miyazaki, Y. and Takano, T. 1983. Impairment  of mitochondria! electron  transport
     by tetrachloroethylene. Toxicol.  Letters 18:  163-166.

Monster, A. C. ;  Boersma,  G. and  Steenweg,  H.  1979.  Kinetics of tetrachloro-
     ethylene  in  volunteers;  influence  of exposure  concentration  and  work
     load.  Int. Arch. Occup. Environ.  Health  42:  303-309.

Monster, A. C.  and  Houtkooper,  J.  M.  1979.  Estimation of  individual uptake of
     trichloroethylene,  1,1,1-trichloroethane  and tetrachloroethylene  from
     biological parameters.  Int. Arch.  Occup.  Environ.  Health  42:  319-323.

Monster, A.;  Regouin-Peeters, W. ;  van Schigndel, A.  and  van der  Tuin, J.
     1983.  Biological  monitoring of occupational  exposure  to tetrachloroethy-
     lene.  Scand.  J. Work  Environ. Health 9:  273-281.

Monster, A. C.  and  Smolders, J.  E.  J.  1984. Tetrachloroethylene in exhaled
     air of persons  living near  pollution sources. Int. Arch. Occup.  Environ.
     Health 53: 331-336.

Morgan,  B.  1969.  Dangers  of perchloroethylene.  Brit.  Med.  J.   2:  513-516.

Moslen,  M.   T. ;  Reynolds,  E.  S.  and  Szabo, S.  1977. Enhancement  of the metabo-
     lism  and  hepatotoxicity  of trichloroethylene  and perchloroethylene.
     Biochem.  Pharmacol. 26: 369-375.

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

Oesch,  R.  1972.  Mammalian epoxide hydrases:  inducible enzymes  catalyzing  the
     inactivation  of carcinogenic  and cytotoxic  metabolites  derived  from
     aromatic and olefinic  compounds.  Xenobiotica 3:  305-340.

Ogata, M. and Hasagawa, T.  1981.  Effects  of  chlorinated aliphatic hydrocarbons
     on  mitochondrial  oxidative  phosphorylation  in  the rat  with reference to
     the effects  of  chlorinated  aromatic hydrocarbons.  Ind.  Health 19:  71-75.

Ogata,  M. ;  Sugiyama,  K.  and Kuroda,  Yl.  1962.  Investigation of  a dry-cleaning
     shop  using tetrachloroethylene,  with  special  reference to Fujiwara's
     substance  in the urine of the employees. Okayama Igakkai Zasshi 74: 247-253.

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

Ohtsuji, T.;  Sato,  K. ; Koizumi,  A.;  Kumai,  M.  and  Ikeda, M. 1983. Limited
     capacity  of  humans to metabolize tetrachloroethylene.  Int.  Arch.  Occup.
     Environ.  Health 51: 381-390.
                                     5-47

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

Pegg, D.  G.;  Zemple,  J.  A.; Braun, W.  H.  andWatanabe, P. G. 1979.  Disposition
     of  (14C) tetrachloroethylene  following oral and  inhalation exposure in
     rats.  Toxicol.  Appl.  Pharmacol.  51:  465-474.

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

Plevova,  J. ;  Frantik,  E.   and  Chladkova,  D.  1975.  Problem  of  interaction
     between  drugs  and some air pollutants. Proc. Eur. Soc.  Toxicol.  16:  303-306.

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

Reichert,  D. ; Werner,  H.  W. ;  Metzler,  M.  and  Henschler, D. 1979.  Molecular
     mechanism of 1,1-dichloroethylene toxicity:  excreted metabolites  reveal
     different pathways of reactive intermediates. Arch.  Toxicol.  42: 259-269.

Riihimaki, V. and Pfaffli, P. 1978. Percutaneous  absorption of  solvent  vapors
     in  man.  Scand.  J.  Work Environ.  Health 4:  73-85.

Sakamoto,  N.  1976.  Metabolism  of  tetrachloroethylene  in  guinea  pigs. Jap. J.
     Ind.  Health 18:  11-16.

Savolainen, H. ;  Pfaffli,  P.;  Tengen,  M.  and Vainio, H.  1977. Biochemical  and
     behavioral   effects  of inhalation exposure  to tetrachloroethylene and
     dichloromethane.  J.  Neuropath. Exp.  Neurology 36: 941-949.

Schumann,  A.  M. ; Quast,  J.  F.  and Watanabe, P. G. 1980.  The pharmacokinetics
     and macromolecular interaction of perchloroethylene in mice  and rats as
     related  to oncogenicity.  Toxicol.  Appl. Pharmacol.  55:  207-219.

Stewart,   R.  D.  and Dodd,  H.  C.  1964.  Absorption  of carbon tetrachloride,
     trichloroethylene, tetrachloroethylene, methylene chloride, and  1,1,1-tri-
     chloroethane through the human skin. J. Ind. Hyg. 25: 429-446.

Stewart,  R. D. ;  Gay,  H.  H.; Erley, D. S.; Hake, C.  L.  and Schaffer, A.  W.
     1961.  Human exposure  to tetrachloroethylene vapor.  Arch.  Environ.  Health
     2:  40-46.

Stewart,  R.  D. ; Baretta, E. D.; Dodd,  H.  C. and Torkelson, T. R. 1970.  Experi-
     mental  human  exposure to  tetrachloroethylene.  Arch.  Environ.  Health
     20:  224-229.

Stewart,  R. D.;  Hake,  C.  L. ;  Forster, H.  V.;  Lebrun,  A. J. ; Peterson,  J. E.
     and Wu,  A.  1974. Tetrachloroethylene--development of a biologic standard
     for  the  industrial  worker  by  breath  analysis.   Report  no.  NIOSH
     NCOW-ENVM-PCE-74-6, 172 pp.

Tada, 0.  and Nakaaki,  K.  1970. Urinary excretion of alkali-pyridine  reactants
     in  humans  exposed to  perchloroethylene vapor.  Rodo Kagaku 45: 588. Chem.
     Abstr.  72:  47126.

                                    5-48

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

Tanaka, S.  and Ikeda,  M.  1968. A method for determination of trichloroethanol
     and trichloroacetic  acid  in urine.  Br.  J.  Ind.  Med.  25:  214-219.

Tsururata, H.  1975. Percutaneous absorption  of  organic solvents.  I.  Comparative
     study  of  the  in vivo percutaneous  absorption  of  chlorinated  solvents in
     mice. Ind. Health 13: 227-236.

U.S. Department of Health, Education  and Welfare. 1977.  Public  Health  Service,
     National  Institute   of  Health,  National Cancer  Institute.  Bioassay of
     tetrachloroethylene  for  possible  carcinogenicity.  DHEW pub.   no.  (NIH)
     77-813.

Vainio, H.  M. ; Parkki,  G. and Marniemi, J.  1976.  Effects of aliphatic chloro-
     hydrocarbons on drug metabolizing  enzymes  in rat  liver j_n  vitro.  Xenobio-
     tica 6: 559-604.

Verberk,  M.  M. Scheffers, T.  M.  L.  1980. Tetrachloroethylene in exhaled air
     of residents near dry-cleaning shops.  Environ.  Res.  21:  432-437.

Watanabe, P.  G. ;  Reitz,   R.  H. ;  Shumann, Z.  M. ; McKenna, M.  J. ; Quast, J.  F.
     and Gehring, P. J. 1980.  Implications  of the mechanisms  of tumorigenicity
     for  risk  assessment. Dev.  Toxicol.  Environ. Sci.  6:  69-89.

Waters, E.  M. ; Geistner,  H.   B.  and  Huff, J. E. 1977.  Trichloroethylene:  an
     overview. J. Toxicol. Environ. Health  2: 670-707.

Weichard, H.  and Lindner, J.  1975. Gesundheitsgefahren durch Perchlorathlen
     in  Chemisch-Reinigungsbetrieben  aus arbeit medizinischtoxikologischer
     Sicht.  Staub Reinhalt.  Luft 35:  416-420.

Weiss,  G.  1969.  A  course  of  observations of trichloroacetic acid excretion in
     occupationally  caused  perchloroethylene  poisoning.  Zentr. Arbeitsmed.
     Arbeitschutz. 19: 143-146.

Yllner,  S.   1961.  Urinary metabolites  of C14-tetrachloroethylene  in  mice.
     Nature  191: 820-821.

Ziglio, G.  1981.  Human exposure to environmental  trichloroethylene and tetra-
     chloroethylene:  preliminary data on population  groups  in Milan, Italy.
     Bull.  Environ. Contam.  Toxicol.  26:  131-136.
                                     5-49

-------
                               6.  TOXIC EFFECTS

6.1  HUMANS
     The  known  effects of  tetrachloroethylene (perchloroethylene,  PCE)  on
humans have been established primarily from Individuals accidentally or occupa-
tional^ exposed to very high (in nearly all cases, unknown) concentrations of
PCE.  Exposure to high concentrations of PCE causes a variety of toxicological
effects in humans.  Effects upon the central nervous system (CNS) are generally
the  most  noticeable  following  acute or  excessive occupational exposures.
Effects upon the liver and kidney usually are observed after an elapsed period
of exposure to high concentrations.

6.1.1  Effects on the  Liver
6.1.1.1   Acute Exposure Situations—Associations   between  liver damage and
acute exposures  of  humans to high  air  levels  of  PCE have been  reported  by
several  investigators (Stewart et  al. ,  1961a;  Stewart,  1969;  Sal and,  1967;
Hake and Stewart, 1977; Levine,  1981).
     Clinical  evidence of impaired liver function was  found by  Stewart  et al.
(1961a) in a worker rendered semi-comatose by overexposure to a petroleum-based
solvent estimated to  contain about  50 percent PCE.  Simulation of the exposure
suggested that during  the 3.5-hr exposure, the  last 30 minutes probably averaged
                          •2
about 1000  ppm (6780 mg/m ) PCE;  the  average  estimated  concentration during
the  initial 3  hours was about 275 ppm (1864 mg/m3).  At such levels, metabolic
saturation was  exceeded.   A neurological examination  conducted  1  hour  after
the  worker's  collapse indicated no  abnormality of function;  also,  the  liver
was  not  palpable.   During a 6-week period subsequent to  the  incident,  neither
clinical  jaundice  nor neurological  deficits were observed.  Beginning  9  days
postexposure,  urinary urobilinogen and total  serum  bilirubin  were elevated.
Serum  glutamic-pyruvic transaminase  (SGPT)  and  serum  glutamic-oxaloacetic
transaminase  (SCOT) were  in  the  normal  range throughout the postexposure
period.  On the  18th  day, SGPT was  slightly elevated.  On  the 16th  postexposure
day,  the  PCE  level  in  expired  air was  sharply elevated.  The  investigators
suggested that such an acute overexposure may  represent a  continuing insult to
the  liver  since  PCE had an exceedingly long exponential  decay in expired  air,
indicating slow  release from body  tissues.
      In  another  case  report,  Stewart  (1969)  diagnosed transient  and  mild
hepatitis  in  a worker exposed  to  anesthetic  levels of PCE for  30 minutes.
                                     6-1

-------
Urinary urobilinogen was  elevated on  the  9th postexposure day.   SCOT was
slightly elevated on the  3rd  and 4th postexposure days.   Stewart concluded
this patient had experienced  marked  depression of the central  nervous system
followed by transient,  minimal liver  injury.
     Hake and  Stewart  (1977)  reported  mild  liver injury, as  indicated by
elevated serum  enzymes,  in a 60-yr-old  male  overcome by  PCE  vapors.   The
individual  was  reported to have  fully recovered.
     In the reports by Levine et al.  (1981) and Sal and (1967), a cause-effect
relationship between PCE  and the effects  observed  cannot be made.
     An enlarged liver and obstructive jaundice were diagnosed by Bagnell and
Ellenberger (1977)   in a 6-wk-old, breast-fed infant.   While this situation is
not uncommon in  infants,  the  infant  had  been  indirectly exposed to PCE.   The
child's father worked as a leather and suede cleaner in a dry-cleaning estab-
lishment where PCE  vapors  were  present.   During regular lunchtime visits to
the exposure site,   the mother had been exposed to the  same  vapors.   These
visits lasted between 30  and 60  minutes.  The  concentration of PCE in the work
place was unknown,   although it was believed to be excessively high because of
reported episodes of dizziness.   In  the  infant, bilirubin, SGOT,  and serum
alkaline phosphatase were elevated;  other blood  and  urinary  parameters of
liver function were normal.  Normal liver function was found in both parents.
Analysis of  the mother's blood  2  hours  after one of her  lunchtime  visits
indicated a PCE  level  of  0.3  mg per 100 ml.   One hour after a  visit,  her
breast milk contained 1.0 mg PCE per  100  mL.   After 24 hours,  the concentration
of PCE  in the breast milk decreased to 0.3 mg per 100 ml.  Chlorinated hydro-
carbons were not found  in the mother's urine.   One week after breast feeding
was discontinued, serum bilirubin and serum alkaline  phosphatase levels in the
infant returned to  a  normal range.   The findings  suggest  that the neonatal
liver may be sensitive to toxicological effects of PCE, although other causal
factors in this instance cannot be ruled out.
6.1.1.2  Chronic Exposure Situations—Alterations  in liver function in persons
exposed to  unknown  concentrations of  PCE  over extended periods  have been
reported by a number of investigators  (Coler and Rossmiller, 1953; Franke and
Eggeling, 1969; Hughes, 1954;  Trense and Zimmerman, 1969;  Meckler and  Phelps,
1966; Larsen et  al.,  1977;  Moeschlin,  1965;  Dumortier et  al., 1964).   Liver
function parameters  that have  been altered as  a result of excessive PCE exposure
include sulfobromophthalein retention time, thymol turbidity,  serum bilirubin,
serum protein  patterns,  cephalin-cholesterol   flocculation,  serum alkaline
                                    6-2

-------
phosphatase, SCOT,  and  serum lactic  acid dehydrogenase  (LDH).   However,  these
parameters may  result  from other causes that are completely dissociated with
PCE.
     Effects attributed  to exposure  to  PCE  at  high  or unknown  levels  included
cirrhosis of the  liver  (Coler and  Rossmiller,  1953),  toxic  hepatitis  (Hughes,
1954; Meckler  and Phelps,  1966), liver  cell necrosis (Trense  and Zimmerman,
1969; Meckler  and  Phelps,  1966),  and  enlarged  liver (Meckler and Phelps,
1966).  In some cases, liver  dysfunction parameters returned to normal follow-
ing cessation of  exposure  (Hughes, 1954).   In one case, the liver was enlarged
6 months after cessation of exposure (Meckler and Phelps, 1966).
     Chielewski et al. (1976) found that SCOT and SGPT  levels were significantly
elevated in a group of 16  of  25 workers compared to non-exposed controls.  The
group of 16  had been  exposed to  PCE  vapors  in  the range of  59  to  442  ppm (400
to  3000 mg/m )  for periods ranging  from  2  months to  27 years.  Serum enzyme
levels in the  remaining group of 9  who were exposed to levels of PCE at or
                      3
below 29 ppm (200 mg/m ) were normal.   Low  urinary excretion of 17-ketosteroids
and abnormal EEG  tracings  was reported  for  members of both exposed groups,  but
no  cause-effect relationship  to PCE exposure can be inferred.

6.1.2  Effects on Kidneys
     Associations between  high air levels of PCE and  symptoms of renal dysfunc-
tion  have been  reported by Larsen et al. (1977) and  Hake and Stewart (1977).
Symptoms included diminished  excretion  of urine, uremia, elevated serum creati-
nine  (Larsen et al.,  1977),  and proteinuria and hematuria (Hake and Stewart,
1977).  A cause-effect  relationship  with PCE exposure cannot be inferred from
the  report  by  Larsen  et al.  (1977).   Similarly, other  factors that may  have
been  responsible  for  the observation of  renal  damage in the case report by
Hake and Stewart  (1977) cannot be precluded.

6.1.3  Effects on Other Organs/Tissues
6.1.3.1  Effects  Upon the  Pulmonary  System  and  Skin--0verexposure  to  high  but
unknown concentrations  of  PCE  have  been associated  with pulmonary damage
(Patel et al.,  1977; Levine et al., 1981).
     Effects of PCE upon the  skin range from a mild to  moderate burning sensa-
tion  upon  direct contact  for 5  to 10  minutes, to  a  marked erythema after
prolonged exposure, and  finally.,  blistering if PCE is  trapped  under  clothing
                                    6-3

-------
or in  shoes  (Hake and Stewart, 1977).  Observations of this nature  have  been
reported also  by  Stewart  and  Dodd  (1964),  Morgan (1969),  and Ling and Lindsay
(1971).
6.1.3.2  Effects  Upon the Heart—In  the  case report of Abedin et al.  (1980),
work-related exposure of  an  individual  to PCE  was  apparently the principal
factor  leading to dizziness,  headaches,  and  premature  ventricular beats.   The
individual's prior history showed no other possible contributing factors.  The
individual's symptoms were more pronounced when the  PCE plasma  level was  high
(3.8 ppm).   Removal  from  exposure  to PCE was reported  to  relieve  the symptoms
as well as the premature ventricular beats.

6.1.4  Effects on CNS and Behavior
6.1.4.1   Effects  of  Short-Term Exposure—Reports of  accidental   short-term
exposure to  PCE  have implied  that such  exposure produces temporary central
nervous system (CNS)  effects (Coler and Rossmiller, 1953;  Lob, 1957; Eberhardt
and Freundt, 1966; Gold,  1969; Stewart,  1969; Bagnell  and Ellenberger,  1977).
Effects most prominently mentioned were dizziness, confusion, headache, nausea,
and irritation of eyes and mucous  tissue.  Higher-level exposures  intensified
the above  symptoms and, at sufficiently  high levels, produced unconsciousness
and eventually death.
     An early  report of  short-term  experimental exposures  to  PCE  of four
subjects (Carpenter,  1937)  described  subjective effects   similar  to those
listed above.  He also observed signs of increased autonomic activity  such as
palmar sweating,  salivation,  and nasal secretion.  Such symptoms were observed
                      3
at 500  ppm  (3390  mg/m )  PCE exposure for  about 2. hours.   Concentrations of
1000 to 1500 ppm  (6780 to 10,170 mg/m3)  PCE  for up to  2 hours produced faint-
ness.   Higher concentrations produced faintness more quickly.
     Rowe et al.   (1952)  collected subjective reports from six subjects exposed
to PCE  concentrations ranging  from about 100 to 1000 ppm  (678 to  6780  mg/m3).
No effects were observed  at 100 ppm  (678 mg/m ),  but the  odor was noticeable.
The threshold  for eye irritation,  dizziness,  and sleepiness  was less than 200
               3                                                          3
ppm (1356  mg/m )  for 20  to 30 minutes.   Exposure to  280 ppm  (1898 mg/m )
intensified the  symptoms  and  also produced "lightheadedness" and numbness
about the mouth.  Subjects  reported feelings of  motor incoordination  at  600
               q
ppm (4068 mg/m )  exposure levels after 10  minutes.  One minute of exposure at
1000 ppm (6780 mg/m  ) produced all of the above  symptoms and was considered
intolerable by the subjects.
                                    6-4

-------
     Stewart et al.  (1970)  collected subjective reports  and  a few objective
measures of  CMS  function in 11 subjects exposed to 100 ppm (678 mg/m3) for  7
hours.   This single-day  study was part  of  a long-term exposure experiment.
Symptoms reported  were  headache,  mucous tissue and eye irritation, autonomic
symptoms such  as  flushed skin, dizziness, sleepiness, and speech  difficulty.
Objective data collected were  performance  of the  Romberg  test of  balance,  the
Flanagan test of coordination, an arithmetic test, a visual inspection task, a
visual  acuity  test,  and  a test of depth perception.  Of  the  objective tests,
none showed any abnormality except for the Romberg test,  which was abnormal  in
three of the 11 subjects.  No  control group was used and  data were compared  to
known norms.  Stewart et al. (1977) executed another long-term repeated exposure
study of PCE using 12 subjects exposed  to  0,  25,  or  100  ppm  (0,  169,  or 678
    3
mg/m ) PCE.  Data  collected concerned the Romberg test of balance, the Michigan
eye-hand coordination  test,  rotary  pursuit  performance, Flanagan  coordination
test, eye  saccade velocity measurement,  the Lorr-McNair  mood scale,  and  a
divided attention  task involving the simultaneous detection of one of an array
of peripheral  red lights while simultaneously  counting the  occurrence of  a
white central light.  EEC spectra were also  collected.  No effect was noted  on
any of  these tests on the  first  day of  exposure  to 100 ppm  (678  mg/m  ) PCE.
In this  study,  data were compared  to  control  (0  or 25 ppm PCE) days.   This
should  have  made  the tests more  sensitive.   It is difficult to explain why
                                         3
this experiment  using 100 ppm (678  mg/m )  PCE did not show  effects on the
Romberg  test while in Stewart et al. (1970) effects were reported.  The fact
that only  three  of the  11 subjects  showed an abnormal  Romberg test in  Stewart
et al.  (1970)  might indicate  abnormally sensitive subjects or the occurrence
of chance results.
     Only three  of the  above  publications  (Rowe et al., 1952; Stewart  et al.,
1970; Stewart  et  al. , 1977) reported sufficiently quantitative data at lower
exposure limits  so that CNS threshold effects  can be estimated.   It appears
that such  effects as dizziness, eye  and mucous tissue irritation, headache,
and sleepiness all appear in  mild form when humans are exposed to between  100
and 200  ppm (678  and 1356 mg/m )  PCE.   Higher exposure concentrations  produce
intensified  symptoms leading  to  confusion  and  nausea.   Feelings of  motor
incoordination and faintness  occur between  600 and 1000  ppm (4068 and 6780
mg/m ).   The data  are too scant to  hazard a  generalization about  the durations
required to produce  symptoms.   The above  symptoms were  reported at  times.
ranging from 10 minutes  to  2 hours.  All symptoms  due  to  experimental  exposures
                                     6-5

-------
were reversed in a matter of hours after cessation of exposure.  These conclu-
sions should be treated with extreme skepticism, however, due to the extremely
small data base, the few subjects studied, and the largely qualitative methods
employed.
6.1.4.2  Effects of Long-Term Exposures—Data about  long-term  (greater  than  a
few  weeks)  exposure  to PCE in humans  originate entirely from case  studies
(Grossdorfer, 1952; Coler and Rossmiller, 1953;  Lob, 1957; Gold, 1969; Weichardt
et al., 1975).   Extended exposure to PCE appears to make many of the temporary
symptoms produced  by short-term  exposure  more continuous  or  of  a  longer  dura-
tion.  Reports of frequent dizziness, headache,  nausea, fatigue, and disorien-
tation are common even for extended periods of time after cessation of exposure.
New symptoms which have no clear analogs in the short-term case appear, however,
during long-term exposure.   Long-term  exposed subjects are  reported  to  have
short-term memory deficits, ataxia, irritability, disorientation, sleep distur-
bances and  decreased alcohol tolerance.   Such symptoms  are sometimes  reported
to be irreversible.
     Stewart et al. (1970) and Stewart et al.  (1977) performed repeated experi-
mental exposures to  100  ppm (678 mg/m ) PCE.   While these were not long-term
studies with  respect  to  frequent  industrial exposures,  these  studies could
provide data about cumulative effects and adaptation over a moderate period of
time.  Subjective  reports  indicated that perceived odor  intensity  decreased
over  repeated  exposures  as well  as within  days.   Subjective complaints  also
decreased over the course of repeated  exposures.   No systematic trend  was
found  in objective test  results  (see Section 6.1.4.1  for  details),  and  indeed
it may be argued that no effects on objective test results could be demonstrated
due to PCE exposure at all.
     Stewart et al.  (1977)  conjectured  that  repeated PCE exposure  effects
might  be  exacerbated by  simultaneous  administration  of  either alcohol or
Diazepam.   While both  alcohol and  Diazepam  produced  decrements  in the various
objective tests  (see  Section 6.1.4.1 for details), the effects were no worse
when  these  substances  were combined with PCE exposure.   It would appear that
repeated 100-ppm (678  mg/m ) PCE  exposure  is not  close to the threshold for
objective test effects.
     There  is  only limited useful  information concerning  the effects of long-
term  exposure  to PCE.  Case  studies provide poor exposure data and are poten-
tially confounded  with  exposure  to other similar  substances which would  be
expected to  produce similar results,  e.g., Tuttle et al. (1977).   Control
                                    6-6

-------
groups are frequently absent In such studies as well.  From repeated exposures
in a  laboratory  setting (Stewart et al.,  1970;  Stewart et al. , 1977) it is
apparent that  some  sensory and symptom  adaptation  to  low-level  PCE  exposure
occurs, but  the  extent  to which such  adaptation  occurs  at  higher  levels  and
the sequelae of such adaptation are not  known.

6.2  LABORATORY ANIMAL STUDIES
     Reported effects associated with  PCE exposure of laboratory animals include
effects on  the CMS,  cardiovascular  system,  liver,  kidney,  skin,  eyes,  and the
immune system.   A  summary of  effects and dose  data  are  provided in Tables 6-1
and 6-2.

6.2.1  Lethality and Anesthesia
     The lethality  and  anesthesia levels for single exposures to PCE in rats
were studied by Carpenter  (1937) and Rowe et al. (1951).  Data reported by the
above articles were  used to construct  the curves of Figure 6-1.  Linear regres-
sion  analyses  were  used to predict  the time required to  produce an effect for
various concentrations  of PCE in a  log-log plot.   The  line labeled "LC100"
represents  the time required  to just  produce death at various concentrations
of PCE.  The line  labeled "ANESTHESIA"  is  the  time  required  to just produce
anesthesia  at various concentrations of  PCE.  The "LCD" line is the prediction
line  for the time  required to produce effects just short of death at various
concentrations of PCE.   In order to depict  the time by concentration curve for
anesthesia  in linear space, Figure 6-2 was  plotted.  It appears that anesthesia
may be produced  in  rats by sufficiently long single  exposures  even by  concen-
trations of  PCE as  low  as  2500 ppm (16,950  mg/m ), even though  nearly 16 hours
of exposure  is required.  While the curves  in Figures 6-1 and 6-2 are probably
of the correct form, the  exact values are likely to be in  error.  Data  were
fitted to  reported  means from groups  of unequal size.  There is likely to be
variation due to strain  of rats.  Log  transformed data were antilog transformed
to produce  Figure  6-2,  a process which  does  not  necessarily  produce a least
squares fit  in linear space.
     Repeated  exposure  to higher  concentrations of PCE  appear to produce con-
siderable tolerance  in  rats.  Carpenter  (1937)  reported that 2,750 ppm (18,645
    o
mg/m  ) PCE  produced anesthesia in rats  on  the  first exposure day.  After six
such  exposures,  however, rats did not become anesthetized  even  though 10,000
ppm (67,800 mg/m )  exposures  were tried.  This finding  corresponds  to reports
                                    6-7

-------
                                            TABLE 6-1.   SUMMARY OF  THE  EFFECTS  OF  PCE ON ANIMALS
   Animal           Dose
   species     concentration
                   Route of
                administration
                 Exposure variables
                                Effects
                                        Reference
   Rabbit        N/A
    (female)
   Rabbit
   Mouse
i
CO
   Mouse
   Guinea
    pig

   Guinea
    pig
   Guinea
    pig
13 mmol/kg
200 ppm
200 ppm
100 ppm
200 ppm
400 ppm
                  skin
oral
inhalation
inhalation
inhalation
inhalation
inhalation
single application
  single instillation
  (eye)

single dose
4 hours
single exposure
4 hours/day
6 days/week
1-8 weeks

7 hours/day
5 days/week
132 exposures

7 hours/day
5 days/week
7 hours/day
5 days/week
169 exposures
                                             primary  eye  and
                                               skin  irritant
marked increase in
  serum enzymes, I.e.,
  alkaline phosphatase,
  SCOT, and SGPT within
  24 hours

moderate fatty infiltration
  of the liver 1 day
  after exposure but not
  3 days after

fatty degeneration of
  the liver
increased liver weights
  in females
increased liver weights
  with some fatty degenera-
  tion in both males and
  females - slight increase
  in lipid content and
  several small fat vacuoles
  in liver

more pronounced liver
  changes than at 200 ppm,
  slight cirrhosis was
  observed - increased liver
  weight, increase in neutral
  fat and esterified choles-
  terol in the liver, moderate
  central fatty degeneration,
  cirrhosis
                                    Duprat  et  al.,  1976
Fujii  et al.,  1975
Kyi in et al.,  1963
Kyi in et al., 1965
Rowe et al., 1952
Rowe et al. , 1952
 Rowe et al.,  1952

-------
                                                          TABLE 6-1.  (continued)
  Animal
  species
    Dose
concentration
   Route of
administration
Exposure variables
     Effects
    Reference
  Guinea
   pig
   Rabbit
   Rat
   Monkey

   Rabbit
01
   Rat
   Rat
   Rat
   Rabbit
   Rabbit
  2500 ppm
  inhalation
  100-400 ppm       inhalation
  2500 ppm
  2500 ppm
  1600 ppm
   3000-6000
  inhalation
  inhalation
  inhalation
   inhalation
   15  ppm
   2212 ppm
   (15 mg/L)
   inhalation
   inhalation
 18 7-hour
 exposures
 7 hours/day
 5 days/week
 6 months

 28 7-hour
 exposures
 1-13 7-hour
 exposures

 18 7-hour
 exposures
 single exposure
 up to 8 hours
 3-4  hours/day
 7-11 months

 45 days
 4 hrs/day
 5 days/week
loss of equilibrium,
  coordination, and strength,
  increase in weights of liver
  and kidney, fatty degenera-
  tion of the liver, cloudy
  swelling of tubular epithe-
  1ium of the kidney

no abnormal growth,
  organ function, or
  histopathologic findings

central nervous system
  depression without
  unconsciousness

loss of consciousness
  and death

drowsiness, stupor, increased
  salivation, extreme restless-
  ness, disturbance of equili-
  brium and coordination, biting
  and scratching reflex

increase in liyer weight, in-
  crease in total lipid content
  of liver accompanied by a few
  .diffusely distributed fat
  globules

depressed agglutinin
  formation

1iver damage
  indicated by elevated
  SGPT, SGOT, SGLDH:
  marked reduction of
  Schmidt index
Rowe et al., 1952
Rowe et al., 1952



Rowe et al., 1952



Rowe et al., 1952


Rowe et al. , 1952





Rowe et al. , 1952





Mazza, 1972


Mazza, 1972

-------
                                                        TABLE 6-1.   (continued)
Animal Dose Route of
species concentration administration Exposure variables
Rat 70 ppm inhalation 8 hours/day
5 days/week
150 exposures
(7 months)
Rat 230 ppm inhalation 8 hours/day
5 days/week
150 exposures
(7 months)
Rat 470 ppm inhalation 8 hours/day
5 days/week
150 exposures
(7 months)
en
o
Rat 2750-9000 inhalation single exposure
Rat 19000 ppm inhalation 30-60 minutes
Rabbit 15 ppm inhalation 3-4 hours/day
7-11 months
Effects Reference
no pathological findings Carpenter, 1937
similar, but less severe, Carpenter, 1937
pathological findings as
with 470 ppm - congestion
and light granular swelling
of kidneys
congested livers with cloudy Carpenter, 1937
swelling, no evidence of
fatty degeneration or
necrosis: evidence of kid-
ney injury - increased
secretion, cloudy swelling,
and desquamation of kidneys:
congestion of spleen
no deaths Carpenter, 1937
congested livers with granular Carpenter, 1937
swelling, some deaths
moderately increased Navrotskii et al . , 1971
urinary urobilinogen,
Rabbit
2211 ppm
(15 mg/L)
inhalation
45 days
  pathomorphologi cal
  changes in the
  parenchyma of liver
  and kidneys

significant reduction
  of glomerular filtration
  rate and the  renal
  plasma flow;  decrease
  of highest excretory
  tubular capacity
  (kidney damage)
Brancaccio et al., 1971

-------
      TABLE 6-1.   (continued)
Animal Dose
species concentration
Mouse 2.5 ml/kg
(Swiss
male,
10 animal's)
Mouse 5. 0 mL/kg
(10
Animals)

Rabbit 2211 ppm
(15 rng/L)
Mouse N/A
Mouse N/A
Dog
Dog
Dog
Rat 300 ppm
Route of
administration Exposure variables
intraperitoneal
intraperitoneal
(urine samples were collected 24 hours
inhalation 45 days
intraperitoneal
intraperitoneal
intraperitoneal
intraperitoneal
intraperitoneal
inhalation 7 hours/day
Effects
100 mg percent or more
protein found in one of
six mice - proximal con-
voluted tubules were
swollen in all animals
and necrotic in one
two of four mice had
100 mg percent or
more protein in urine
post- injection)
increased plasma and urine
levels of adrenal cortical
and adrenal medullar hor-
mones; increased excretion
of principal catecholamine
metabolite (not statistically
significant)
liver dysfunction
LD50
elevated SGPT
caused phenolsulfo-
nephthalein retention
indicating kidney dysfunction
LD50
decreased maternal
Reference
Plaa and Larson, 1965


Mazza and Brancaccio,
1971
Klaassen and Plaa, 1966
Klaassen and Plaa, 1966
Klaassen and Plaa, 1967
Klaassen and Plaa, 1967
Klaassen and Plaa, 1967
Schwetz et al . ,
days 6-15 of
gestation
weight gains,
increased fetal
reabsorptions
1975

-------
                                                           TABLE 6-1.  (continued)
Animal
species
Mouse
Dose Route of
concentration administration Exposure variables
300 ppm inhalation '-7 hours/day
days 6-15 of
gestation
Effects Reference
maternal liver weights Schwetz et al.,
increased relative to 1975
body weight; increased
incidences of fetal
subcutaneous edema,
delayed ossification of
skull bones, and split
sternebrae
en
i
   Rat
   Mouse
   Rat
    Rat
    Dog
    (male
    beagles)
44.2 ppm
15-74 ppm
15 ppm
73 and
147 ppm
0.5-1.0%
v/v
5000 &
10000 ppm
inhalation        entire gestation
                  period
inhalation        5 hours/day
                  3 months
inhalation        4 hours/day
                  5 months
inhalation        4 hours/day
                  4 weeks
inhalation        7 minutes house air
                  followed by 10
                  minutes of tetrachloro-
                  ethylene 8 ug/kg
                  epinephrine given I.V.
                  (1) a control dose
                  after 2 minutes of
                  breathing air (2) chal-
                  lenge dose after 5 min-
                  utes of breathing test
                  compound
decreased levels of DNA
  and total  nucleic acids
  in the liver, brain,
  ovaries, and placenta

decreased electroconductance
  of muscle  and "amplitude"
  of muscular contraction

EEG changes  and proto-
  plasmal swelling of
  cerebral cortical cells,
  some vacuolated cells and
  signs of karyolysis

EEG and electromyogram
  changes; decreased
  acetylcholi nesterase
  activity

cardiac sensitization
  (development of serious
  arrhythmia or cardiac
  arrest) was not induced
  at the concentrations
  tested (other similar
  compounds gave positive
   results at  same  concen-
   tration,
Aninina, 1972
Dmitrieva, 1968
Dmitrieva, 1966
Dmitrieva, 1966
Reinhardt et al., 1973

-------
                                                            TABLE 6-1.   (continued)
cr>
i
Animal
species
Cat
Cat
Mouse
Mouse
Rabbit
Cat
Dog
Dose
concentration
3000 ppm
14,600 ppm
40 mg/L
5,900 ppm
4-5 mL/kg
5 ml/kg
4 mg/kg
9000 ppm
Route of
administration Exposure variables
inhalation 4 hours
inhalation 1-2 hours
inhalation
oral
oral (in oil)
oral (in oil)
inhalation
Effects
no anesthesia
anesthesia
minimal fatal concentration
death in 2-9 hours from
central nervous system
depression
death in 17-24 hours
death within hours
narcosis, marked
Reference
Lehmann, 1911


Lehmann and Schmidt-
Kehl, 1936
Lamson et al . ,
Lamson et al . ,
Lamson et al . ,
Lamson et al . ,
Lamson et al . ,
1929
1929
1929
1929
1929
                                                                                   salivation, "narrow

                                                                                   margin of safety"
    Dog
4-25 mL/kg
oral (in oil)
death in 5-48 hours
Lamson et al.,  1929

-------
                                                   TABLE 6-2.  TOXIC DOSE DATA
01
I
Description
of exposure3
LDEO
LD50
ED50
EDso
LD50
ED50
ED50
LD50
ED50
LD50
LD50
LD50
LCLo
LD50
Species
mouse (male)
mouse
mouse
mouse
mouse
dog
dog
dog
mouse
mouse
mouse
mouse
mouse
mouse
rat
Route of
administration
oral
interperitoneal
interperitoneal
interperitoneal
interperitoneal
interperitoneal
interperitoneal
interperitoneal
subcutaneous
subcutaneous
oral
(undiluted)
oral
(in oil)
oral
inhalation

Dose
concentration
8100 mg/kg
2.9 ml/kg
28 mM/kg
4700 mg/kg
2.9 ml/kg
28-32 mM/kg
34 mM/kg
24 mM/kg
2.1 ml/kg
21 mM/kg
3400 mg/kg
0.74 ml/kg
7.2 mM/kg
1.4 mL/kg
390 mM/kg
27 mM/kg
0.109 ml
0.134 ml
8850 mg/kg
23000 mg/m3

Toxic effect
endpoint
death
death
1 iver
dysfunction
death
liver toxicity
death
liver damage
ki d'fie'y
dysfunction
death
liver toxicity
death
death
death
death
death
Time Reference
36 hr Wenzel and Gibson, 1951
24 hr Klaassen and Plaa, 1966

24 hr Gehring, 1968
24 hr Klaassen and Plaa, 1967
24 hr Klaassen and Plaa, 1967
24 hr Klaassen and Plaa, 1967
10 days Plaa et al. , 1958
Plaa et al. , 1958
unknown Dybing and Dybing, 1946
unknown Dybing and Dybing, 1946
unknown Handbook of Toxicology, 1959
2 hr
Withey and Hall, 1975

-------
                                                     TABLE  6-2.   (continued)
Description
of exposure3
LCLo
LCLo
LDLo
LDLo
LDLo
LDLo
Species
rat
rat
dog
dog
cat
rabbit
Route of
administration
inhalation
inhalation
oral
intravenous
oral
oral
Dose
concentration
4000 ppm
4000 ppm
4000 mg/kg
85 mg/kg
4000 mg/kg
5000 mg/kg
Toxic effect
endpoint
death
death
death
death
death
death
Time
4 hr
4 hr
unknown
unknown
unknown
unknown
Reference
Handbook of Toxicology, 1959

Arch. Hyg. Bakteriol.
116:131, 1936
Carpenter et al . , 1949
Clayton, 1962
Lamson et al . , 1929
LCL  - the lowest concentration of a substance,  other than an LC50,  in  air  which  has  been  reported to  have caused death  in
humans or animals.
LDL  - the lowest dose of a substance, other than an LD50, that is  introduced  by  any  route  other than  inhalation over any
givin period of time and reported to have caused death in humans or animals introduced  in  one or more  divided
portions.

-------
1000
                                                            O  LCD (ROWE etal. 1951)
                                                            D  ANESTHESIA (CARPENTER. 1937)
                                                            A  LC100 (Rows et al. 1951)
  10
                            345              10
                               PCE CONCENTRATION (x 103), ppm

               Figure 6-1. Concentration-time curves for various effects of PCE

               Source: Carpenter (1937) and Rowe et al (1951).
30    40
                                          6-}6

-------
1000
                  5                     10                     15



                               PCE CONCENTRATION (x103), ppm





              Figure 6-2. Concentration-time curve for PCE-induced anesthesia.
20
                                        6-17

-------
of tolerance in humans  (Stewart et al.,  1970;  Stewart et al., 1977).   Behavioral
adaptation,  although in a less obviously systematic way, has also been reported
in rats by  Rowe et al.  (1951) and Goldberg et al.   (1964) (see Section 2.2.2).
     Effects of high-level  PCE  exposure in rats appeared  to be similar to
those of other anesthetics.   When death  occurred due to a single PCE exposure,
the process  was highly  similar to narcotic overdose (Carpenter,  1937).   Although
rats dying from repeated exposures to  high-level PCE showed some liver involve-
ment, the death was still attributable to CNS impairment (Rowe et al., 1951).

6.2.2  Behavioral  Effects
     Behavioral effects  of  high-level  and repeated  exposures were  reported
qualitatively by Carpenter  (1937)  and  Rowe et  al.   (1952).   Most often noted
were ataxia, somnolence,  and eventually  anesthesia.  A group of 18 rats exposed
to 1600 ppm  (10,848  mg/m )  PCE for 7 hr/day for 18  days were reported to be
stuporous during the first week and to have shown  increased  salivation, motor
activity, "reflexive"  aggression  and some ataxia,  but the  latter symptoms
subsided toward the end  of the second week (Rowe et  al., 1951).  Brief recur-
rence of the symptoms was reported at the beginning  of  each  week, but overall
reduction was  a function of repeated exposures.   Some  of  the symptoms are
reminiscent of those reported in high-level, long-term  industrial exposure of
humans (see Section 2.1.4.2).
     Goldberg et al. (1964)  studied  the effects of 0, 1150, and 2300 ppm (0,
                     3
7797, and 15,594 mg/m ) PCE  exposure  for 4 hr/day,  5 days/week for 2 weeks,  in
groups of 8  to 10  rats.   Rats were tested  before  and  after exposure on an
escape/avoidance task in which the response was pole  climbing.  An 80 percent
loss of both  escape  and avoidance responses due to ataxia was reported after
                                3
the  first 2300 ppm  (15,594  mg/m ) exposure period.  Upon repeated exposures,
however, the effect was  diminished.  No data were  presented  on the adaptation
effect other than  that  it occurred.
     Low-level  repeated  PCE exposure (200  ppm, 6 hr/day,  for 4  days) was
studied in  groups of 10 rats by Savolainen et  al. (1977).  Many dependent
measures were  collected, including brain  protein,  RNA,  glutathione,  acid
proteinase,  and nonspecific cholinesterase and six  aspects of  open field
behavior.   The behavioral variables were collected  and  a separate significance
test was done  on each measure after the last exposure day  and 17 hours  later.
Of the 12 t-tests  performed,  only the preening rate and preening time were
significant and only immediately  after  the last exposure.   The authors  also
                                    6-18

-------
discussed reduced RNA and increased nonspecific cholinesterase content in brain
due to  PCE  exposure.   If these results are replicable, they represent a near
threshold effect of PCE in rats.
     Ataxia and  somnolence  due to PCE ingestion  have  been reported in dogs
given 0.22  mg/kg for  worm  treatment  (Snow,  1973;  Myer and Jones,  1954).
However, such  side  effects  are not universally found at this dosage (Miller,
1966; Lamson et al., 1929; Schlingman, 1925).   Occurrence of neurotoxic effects
probably depends  upon  the extent  of absorption  of PCE  via  ingestions which  is
in turn likely to be affected by diet (Snow, 1973) as well as by the strain of
dog.  In the absence of tissue  levels, ingestion data are difficult to interpret,
     Extremely  low  levels of  PCE exposure were  reported  to have affected
electrophysiological responses  in rats (Dmitrieva, 1966, 1968, 1973; Dmitrieva
et  al.,  1968).   Documentation  is  sparse,  and  the  effective dose  levels are  so
low compared to the rest of the literature that the credibility of the reports
must be questioned until independently replicated.

6.2.3  Effects on the Liver and Kidney
     PCE is generally  regarded as being  both hepatotoxic  and nephrotoxic in
laboratory animals when exposure is excessive and prolonged.
     Carpenter (1937) exposed three groups of 24 albino rats each to PCE vapor
concentrations averaging  70, 230,  or  470  ppm  (475,  1560,  or  3188  mg/m3) for 8
hr/day,  5  days/week,  for  up  to 7  months.   The maximum  exposure for  any animal
during  the  7-mo  period was  150 days  (1200 hours).   A group of 18 unexposed
animals served as controls.
     The rats  exposed to 470  ppm (3188  mg/m3)  for 150  days followed  by  a
46-day  rest period  developed cloudy and congested livers  with swelling; there
was no  evidence  of  fatty degeneration or necrosis.  These rats  also had  in-
creased  renal  secretion  with cloudy  swelling and desquamation of kidneys,  as
well as congested spleens with  increased  pigment.  The pathologic changes were
similar  but  less  severe in  the rats  exposed  to 230 ppm PCE  (1560 mg/m3).    In
some instances, there was congestion  and  light granular swelling of the kidneys
after 21 exposure days.   After 150 days  of exposure,  followed by  a 20-day
rest, congestion was found in the kidney  and  spleen.  The  livers  showed reduced
glycogen storage.   Microscopic evidence of damage to  liver,  kidney, or  spleen
in rats exposed at 70 ppm (475  mg/m3) for 150 days was not observed.  In addi-
tion, microscopic examination  of heart,   brain, eye,  or  nerve tissue did  not
reveal any damaging effects in  any of the chronically exposed rats.  Functional
                                    6-19

-------
parameters,  including icteric  index,  Van den Bergh test  for  bilirubin,  and
blood and  urine  analysis, were  normal  after the  exposures.   Fertility  of
female rats,  as measured by a fertility  index (actual  number of litters/possible
number of litters), was increased slightly after repeated exposures to 230 or
470 ppm PCE (1560 or 3188 mg/m3).   No deaths or signs  of anorexia or diminished
activity were observed during the chronic exposures.
     Carpenter also tried to determine the highest concentration of PCE vapor
that would not anesthetize  rats  exposed for 8 hours.   Exposure to 31,000 ppm
(210,273 mg/m3) was  lethal  within  a few minutes.   Rats exposed to 19,000 ppm
(128,871 mg/m3) died  after  30  to 60 minutes.  Animals  that were exposed to
19,000 ppm (128,877 mg/m3) and removed from the inhalation  chamber just prior
to unconsciousness developed congestion  and  granular  swelling of the liver.
Similar liver effects were  seen  after exposure at 9000  ppm (61,047 mg/m3).
There was also marked granular swelling of the kidneys.  A  single exposure at
9000, 4500, or 2750 ppm (61,047, 30,523, or 15,261 mg/m3) did not cause death
to any of the  rats  in this study;  however,  post-mortem examinations  of  the
rats exposed to those  concentrations  revealed only a slight increase in the
prominence of liver and kidney  markings.
     Rowe et al.  (1952) exposed rabbits,  monkeys,  rats, and guinea pigs  to PCE
vapor for 7  hours,  5 days/week,  for up  to 6 months.   Exposure concentrations
ranged from 100 to 2500 ppm (678 to 16,957 mg/m3).  Three of the four species
tested -- rabbits, monkeys,  and rats --  showed no  effects of repeated  exposures
to concentrations up  to  400  ppm (2713 mg/m3).   There  were no adverse  effects
on growth, liver  weight,  or  lipid content, or gross  or microscopic anatomy
observed in any animal.   In contrast,  guinea pigs  showed marked susceptibility
to PCE  in this study.  The  liver  weights of female  guinea pigs increased
significantly after 132 7-hr exposures  at 100 ppm (678  mg/m3).   At 200  ppm
(1356 mg/m3), there  was  a slight depression of growth in female guinea pigs
and  increased liver  weights  in both males and females.   Slight  to moderate
fatty degeneration of the  liver  also was observed.  These  effects were more
pronounced in guinea  pigs  that received 169 7-hr exposures at 400 ppm (2713
mg/m3).   At this  concentration,  there also were increased amounts of neutral
fat  and esterified cholesterol  in  livers.   Gross  and  microscopic examination
of the tissues revealed  slight to moderate fatty  degeneration  in the liver
with slight  cirrhosis.   Rowe  et al.  stated that  at  395  ppm  (2680 mg/m3),
increased kidney weights  also  were  observed in guinea pigs but  not in other
species.   Guinea pigs  have  been  shown to be extremely sensitive to toxicity
testing.
                                    6-20

-------
     Klaassen and  Plaa  (1967)  showed that near-lethal  i.p.  doses of PCE in
mongrel male and female dogs can produce damage to the kidney and liver.  They
estimated the ED50  (effective  dose in 50  percent  of  the  animals  tested)  for
liver and kidney  damage as well as  the  24-hr LD50 value (lethal dose in 50
percent of the animals treated).  The ED50 values were measured by sulfobromo-
phthalein, SGPT, glucose,  protein,  and phenolsulfonephthalein (PSP), indica-
tors of  liver or  kidney dysfunction.  The LD50  was  2.1 ml/kg (21 mmol/kg)
and the ED50 for elevation of SGPT was 0.74 mL/kg (7.4 mmol/kg).   The ED50 for
diminution of PSP excretion was 1.4 ml/kg (14 mmol/kg).  After administration,
effects on the  liver and kidneys were determined by microscopic examination.
At near-lethal  doses, PCE  produced moderate  neutrophilic  infiltrations  in the
sinusoids and portal areas; necrosis was not observed.  Near the ED5Q, vacuoli-
zation of centralobular hepatocytes  in about half the animals was observed.
After a single i.p.  injection at 0.75 x LD50, SGPT peaked at 2 days and rapidly
declined  throughout the 9-day  measurement period.   Kidney dysfunction was
deemed significant  when PSP  excretion was  less than 39 percent (determined 24
hr after  interperitoneal  injection).  At  near PSP  excretion--ED5Q doses, only
mild dilation of the collecting ducts was seen in some of the kidneys.
     In a similar  study using Swiss-Webster mice, Klaassen  and  Plaa (1966)
determined that the  24-hour LD5Q was 2.9 ml/kg (28 mmol/kg).   The ED5Q for BSP
retention was 2.9  mL/kg (32  mmol/kg).  The ED5Q  for elevation of  SGPT was 2.9
mL/kg (28 mmol/kg).   When ethanol  (5 g/kg)  was  administered by gavage  for 3
days prior to injection of 1.0 mL/kg PCE, neither PSP excretion nor BSP reten-
tion was  significantly altered.  At ESP-ED™, mice  exhibited predominantly
inflammatory changes with  trace to marked quantities  of  lipid accumulation.
     A subcutaneous  injection  (0.4 mL daily,  for 8 weeks) of PCE  to  groups of
rats that were  fed  a diet containing  high protein had  less  of  an adverse
effect upon  the liver  than  PCE alone (Dumitrache et  al.,  1975).   PCE was
observed  to  induce  liver  hypertrophy compared to the  reaction of  the controls
(p <0.001).   Hypertrophy  was most pronounced in the  low-protein  group.  In
treating  rats fed  a diet containing low  protein,  cholesterol (p  <0.001)  and
total liver  lipids  (p < 0.001)  were elevated compared to the reaction of the
controls.   Twelve rats were used in each of the four groups.
     Effects upon  the livers of  rats exposed continuously for 3 months  to 0.7
                             o
and 2.7 ppm  (4.5  and 19 mg/m )  were  reported by Bonashevskaya (1977).   No
significant  histomorphological  changes  in the liver were reported at either
                                    6-21

-------
exposure level.  Slight  alterations  in succinic dehydrogenase and glucose-6-
phosphate dehydrogenase were reported.
     Kyi in et al.  (1963) noted moderate fatty degeneration of the liver with a
                                                3
single 4-hr exposure  to  200 ppm PCE (1356 mg/m )  in  female  albino mice  that
were sacrificed 1  day after exposure.   Degeneration was not observed in mice
sacrificed 3 days  after exposure.   The  mice were exposed to PCE concentrations
of 200, 400,  800,  or  1600 ppm  (1356,  2713,  5426,  or  10,852  mg/m3) for 4  hr.
Tissues were  studied  microscopically to assess  the extent  of necrosis and  fat
infiltration of the liver.   Moderate to massive infiltration was observed in
                                                            3
mice killed 1 or 3 days after exposure  at 400 ppm  (2713 mg/m ) or more,  but no
cell necrosis  was  observed  even after  4 hours  exposure up to 1600  ppm  PCE
            3
(10,852 mg/m ).  Kyi in et al.  (1965) exposed four  groups of  20 albino mice to
200 ppm PCE (1356  mg/m  ).   Each group  of 20 mice was exposed for 4  hr/day, 6
days per week, for 1,  2, 4,  or 8 weeks.  Microscopic examinations were performed
on  livers  and  kidneys of the exposed  mice and  controls.   Fatty  degeneration
was particularly marked  and tended to  be more severe with longer exposure to
PCE.  Chemical determination of the liver fat content was performed  in addition
to the histologic  examination.  Correlation between the histologically evaluated
degree of fatty degeneration and the concentration of extracted fat was +0.74.
Liver fat  content  of  the exposed animals was between 4 and  5 mg/g  of body
weight, as  compared to 2 to  2.5 mg/g for the control  animals.  The actual  fat
content of  the livers did not  increase with duration of exposure as did  the
extent of  the  fatty  infiltration.   No  liver cell  necrosis was observed.   No
effect on the  kidneys was reported upon histologic examination.
     Mazza  (1972)  exposed 15  male  rabbits  for 4 hr/day, 5  days a week, for 45
days, to  2790  ppm  PCE (18,924  mg/m3).   Mazza looked  at the  effect of PCE  on
serum enzyme levels in an attempt to determine the specific  location of initial
liver injury  as well  as  the  severity of the  damage to the  liver.  The Schmidt
Index, which  is the  sum of SCOT and the SGPT divided by the serum glutamate
dehydrogenase  (GDH),  was used as an indication  of  hepatic  disorders.   Enzymatic
determinations were made before exposure and 15, 30,  and 45  days after exposure
to  PCE.   All  three of the enzymes  showed an  increase  in activity, but the GDH
increased  the  most; GDH  reduced the Schmidt Index from 6.70 to  1.79.  Mazza
concluded that this reduction indicates the  prevalence of  mitochondrial injury
over cytoplasm!c injury in the  liver.
                                    6-22

-------
     Mazza and  Brancaccio  (1971)  exposed 10 rabbits for 4 hr/day, 5 days per
week, for  45  days,  to  2790  ppm  PCE  (18,924  mg/m  ).   These  investigators  found
a moderate, but not statistically significant, increase in levels of adrenal
cortical and  medullar  hormones—plasma and urinary corticosteroids and cate-
cholamines—including increased excretion of 3-methoxy-l-hydroxymandelic acid,
the principal  catecholamine metabolite.
     In another  study,  Brancaccio et  al.  (1971)  exposed  12 male  rabbits  for  4
hr/day, 6  days/week,  for 45 days, to  2280  ppm PCE  (15,465 mg/m3)  to  look at
effects on  kidney  function.   They noted a  reduction in glomerular filtration
and  renal  plasma flow and a decrease  in  the  maximum tubular excretion when
measured upon cessation of  the  exposure  regimen.  Brancaccio  et  al. concluded
that PCE causes kidney  damage,  primarily  in the  renal tubule.  These  findings
agreed with earlier histological  findings  of  Pennarola and Brancaccio (1968)
in which kidney  injury,  following exposure  to  PCE, appeared to be primarily in
the  renal tubule.
     Plaa  and Larson  (1965) dosed mice with PCE by i.p.  injection.  Ten mice
received 2.5  mg/kg  and  10 others  received  5.0  mg/kg.  Urine samples were col-
lected  from surviving  mice  24 hours after  the injection of PCE.   Protein was
found in the  urine  of  one  of the six surviving  mice  injected with the lower
dose and in two of  four survivors of  the  higher  dose  at  levels of  100 or more
mg percent.   None of  the survivors  had greater than  150  mg percent glucose in
the  urine.  The kidneys of  the  mice given  the  lower  dose were examined micro-
scopically.   The proximal  convoluted  tubules  were swollen in all animals and
necrotic in one.
     Cornish  et al.  (1973)  administered from  0.3 to  2.0 ml   PCE/kg i.p. to
rats.  SGOT levels were  elevated at all dose levels; phenobarbital pretreatment
did  not alter the response.
     Fujii  (1975)  observed an  increase  in serum enzyme activities (i.e.,
alkaline phosphatase,  SGOT,  and SGPT) within  24 hours after  a single dose of
13 mmol PCE/kg  given orally to rabbits.  These changes in serum enzyme activi-
ties, indicative of liver damage, were mild and transient.

6.2.4  Effects  on the Heart
     The effects of PCE upon the heart have not been systematically investi-
gated.   Kobayashi et  al. (1982) demonstrated  in rabbits under urethane anes-
thesia and in cats  and  dogs under pentobarbital anesthesia that PCE can sensi-
tize the  myocardium to  epinephrine,  leading  to  ventricular  arrhythmias and
                                    6-23

-------
premature contractions, bigeminal  rhythm,  and tachycardia.  Rabbits were the
most sensitive; doses of 0.6 |jg epinephrine per  kg and  5 mg  PCE per  kg  led  to
tachycardia.   In dogs,  20  to  40 mg PCE  per  kg,  i.v.,  produced significant
depressions in the  rate of rise of left  intraventricular  pressure with the
absence of significant effect upon arterial pressure.   The mean lethal dose in
cats was 81.4 ± 14.4 mg PCE per kg, i.v.   As  the authors noted, and as described
in the  preceding chapter,  the  absence of  cardiac  effects  in humans  in  the
literature may be  largely  due  to the  limitations on absorption of PCE by the
lungs.   Additionally, the  levels  to which individuals are ordinarily exposed
both environmentally and occupationally are obviously below  the threshold for
arrhythmia production via  endogenously  produced  epinephrine.  Experiments on
unanesthetized laboratory animals involving tasks designed to elicit endogenous
epinephrine in combination with vapor  levels  of PCE should produce more useful
information on the  cardiotoxicity of PCE.
     A  lack of cardiotoxicity of PCE  was reported by  Reinhardt et  al. (1973).
Unanesthetized dogs  (17) were  exposed to PCE  in air  at levels of 5,000 or
                                 3
10,000 ppm (33,915  or 67,830 mg/m ).  No positive responses were observed.   In
this study, a response  considered  indicative of cardiac sensitization was the
development of a life-threatening  arrhythmia or cardiac arrest  following a
challenge of epinephrine.
     Christensen  and Lynch  (1933)  observed depression of the heart in 5 dogs
at near-LDgQ oral  doses.

6.2.5  Effects on  the Skin and Eye
     Duprat et al.  (1976) have shown PCE to be a primary eye and skin irritant
in rabbits.   Instillation of the chemical into the  eye produced conjunctivitis
with epithelial abrasion.   However, healing of the ocular  mucosa was  complete
within 2 weeks.  PCE had a severe  irritant effect when a  single application
was made to the skin of the rabbit.

6.3  ADVERSE EFFECTS OF SECONDARY POLLUTANTS
     The  level of  phosgene derived from photodecomposition  of  PCE  is  not
likely to result in  acute  or long-term effects.   In occupational settings  or
under certain conditions  in which  phosgene is directly formed at high tem-
peratures from halocarbons (cigarette  smoke,  welding), sufficient warning from
the generation of extremely irritating hydrogen chloride vapors would prevent
exposure to harmful  concentrations of  phosgene.
                                    6-24

-------
6.4  SUMMARY OF ADVERSE HEALTH EFFECTS AND ASSOCIATED LOWEST OBSERVABLE
     EFFECT CONCENTRATIONS
6.4.1  Inhalation Exposure
     A number  of  case  reports  describe  accidental  or  occupational  exposure  to
PCE.   However,  the  duration and extent  of  exposure either were unknown  or
involved excessively high  concentrations.   The few controlled  human  studies
available generally provide  subjective  information on effects resulting from
short-term exposures at  levels near  100  ppm  (678 mg/m3).   The observations  in
these studies follow the expected pattern for a non-specific anesthetic effect.
Effects associated with  chronic exposures,  on the  other  hand,  are available
from animal experiments.
     In humans, short-term effects of PCE exposure  seem to  occur beginning  at
100 to 200 ppm (678 to 1356 mg/m3).   Such effects include dizziness, confusion,
nausea, headache,  and irritation of the eyes and  mucuous tissue.  Higher
levels of  exposure increase  the  symptoms and  eventually produce faintness and
unconsciousness.  Long-term  exposures are reported  to make  the  above  symptoms
more pervasive and also to affect short-term memory and produce disorientation,
irritability, ataxia,  and  sleep disturbance.   These conclusions are based on
largely anecdotal  data in field studies or  limited laboratory experiments.
Due to the paucity  of experiments,   the  poor  quality  of the extant data base
and the age  of  the data  base,  these  conclusions should be viewed with  consid-
erable skepticism.  Newer, more sensitive methods,  better experimental control,
more quantitative analysis,  and a wider range of  dependent variables could
demonstrate effects at lower levels.
     While suffering from  the  same deficiencies as those  outlined  for the data
on humans and PCE exposure, data from laboratory animals  are remarkably conso-
nant with human effects.    Effects on open field activity  have been  reported at
200 ppm (1356  mg/m3)  exposure levels.  Effects of higher level exposure seem
to relate  well  to the findings in humans.   It appears that the rat is a good
behavioral model for PCE effects in  man.
     Investigations in rats have likened the effects  of PCE to those of narcotic
drugs and anesthetics.   The lowest effects levels for PCE seem to  be  in CNS or
behavioral variables.  This  seems  to be true  for  either single or repeated
long-term exposures.   Similar  conclusions  are reached in a review by Annau
(1981).
     Effects upon the  human kidney  and/or  liver are  not  well  documented  and
appear to  be confined to  those  exposure scenarios in which PCE  levels  are
                                     6-25

-------
exceedingly high and exposures are short in duration.   Studies of experimental
animals suggest that reversible changes (congestion, fatty infiltration) occur
at exposure levels  of about 200 ppm (Carpenter,  1937;  Rowe et al. ,  1952; Kylin
et al.,  1963).   The relevance  of  these data to the  likelihood of similar
effects in humans at similar levels is unknown.
     Limited data involving anesthetized laboratory animals  indicate that  low
doses of PCE,  administered i.v., can sensitize the heart to i.v. challenges of
epinephrine.   However, it  is not considered  likely that PCE will exhibit this
action in humans under exposure conditions  normally experienced in  the environ-
ment and the workplace.

6.4.2  Oral Exposure
     The acute oral  toxicity  of PCE has been determined in rats,  mice, cats,
rabbits, and dogs.   There are  no subchronic oral  exposure studies  and only one
chronic study—the  NCI bioassay.

6.4.3  Dermal  Exposure
     Although PCE can be absorbed through unbroken skin, absorption was esti-
mated to be minor  (Stewart and Dodd,  1964).   Toxic quantities would probably
not be absorbed through this route.
                                    6-26

-------
6.5  REFERENCES FOR CHAPTER 6


Abedin, Z. ,  R. C.  Cook, Jr., and R. M. Milberg. 1980. Cardiac  toxicity  of
     perch!oroethylene (a dry cleaning agent).  Southern Med.  J.  73(8):
     1081-1083.

Aninina, T.  1972.  Effect of aliphatic hydrocarbons and  fluorinated  and  chlori-
     nated derivatives on the content of nucleic acids  in  animal  tissues
     during embryogenesis.  Tr. Permsk. Cas. Med. Inst.  110:69-71.

Archiv Hyg.  Bakteriol.  1936. (Munchen)  116:131.

Bagnell, P.  C., and H. C. Ellenberger. 1977. Obstructive jaundice due to a
     chlorinated hydrocarbon in breast milk.  J. Can. Med. Assoc. 117:
     1047-1048.

Blair, A.,  P.  Decoutle, and D. Graumen. 1979. Causes of death  among laundry
     and dry-cleaning workers.  Am. J. Public Health 69(5):  508-511.

Bonashevskaya, T.  I. 1977. Certain Results  of a morphological  and functional
     investigation of the lungs in a hygienic assessment of  atmospheric pollu-
     tion.   Gig. Sanit. 2:15-20.   (Translation)

Brancaccio,  A., V. Mazza, and R. DiPaola. 1971. Renal function in experimental
     tetrachloroethylene poisoning.  Folia  Med.  54:233-237.   (English  trans-
     lation)

Brown, D. P. 1978. Memorandum.  Division of Surveillance,  Hazard  Evaluations,
     and Field Studies, National Institute  for Occupational  Safety  and  Health.
     August 24.

Carpenter,  C.  P. 1937. The chronic toxicity of tetrachloroethylene.  J. Ind.
     Hyg. Toxicol. 19:323-336.

Carpenter,  C.  P., H. F. Smyth, Jr., and V.  C. Pozzani.  1949. The  assay  of
     acute vapor toxicity and the  grading and interpretation of results on 96
     chemical  compounds.  J. Ind.  Hyg. Toxicol.  31:434.

Chmielewski, J., R. Tomaszewski, P. Glombiowski, W.  Kowalewski, S.  R.
     Viwiatkowski, W. Szozekocki,  and A. Winnicka. 1976. Clinical observa-
     tions of the occupational exposure to  tetrachloroethylene.   Bull.  Inst.
     Marit.  Trop.  Med. Gdynia 27(2):197-295.

Christensen, B. V., and H. J. Lynch. 1933.  The effect of anthelmintics  on  the
     host.  I.  Tetrachloroethylene  II.  Hexylresorcinol.  J.  Pharmacol.  Exp.
     Ther.  48:311-316.

Clayton, J.  W. 1962. The toxicity  of fluorocarbons with special reference  to
     chemical  constitution.  J. Occup. Med.  4:262-273.

Coler, H. R.,  and H. R. Rossmiller. 1953. Tetrachloroethylene  exposure  in  a
     small  industry.  Ind. Hyg. Occup. Med.  8:227.
                                     6-27

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

Dmitrieva, N.  V.  1966.  Maximum permissible concentration of tetrachloroethylene
     in factory  air.   Hyg.  Sanit.   31:387-393.  (English translation).

Dmitrieva, N.  V.  1968.  Bioelectric activity and electric conducting properties
     of muscles  exposed to chlorinated hydrocarbons.  Farmakologiya i
     Toksikologiya  31(2):228-230,  (English translation).

Dmitrieva, N.  V.  1973.  Changes in the bioelectrical activity in the cerebral
     cortex of rats with the narcotic effect of substances with different
     polarization properties.   Experimental' naya Khirurgiya i Anestezidogiya
     6:72-75.  (English translation)

Dmitrieva, N.  V., and E. V. Kuleshov.  1971. Changes in the bioelectric activity
     and electric conductivity of the brain in rats chronically poisoned with
     certain chlorinated hydrocarbons.  Hyg. Sanit.  36:23-29.  (English
     translation)

Dmitrieva, N.  V., E.  V.  Kuleshov,  and E.  K. Orjonikidze. 1968. Changes in the
     impedance and bioelectrical  activity of the cerebral cortex of rats under
     the action  of anesthetic drugs.   Zhur. vysshei Nervnoi Deyatel 'nosti
     18(3):463-468.  (English translation)

Dumitrache,  S.,  I.  Gontea, and V.  Stanciu. 1975. Role of proteins in the
     resistance  of the organism to tetrachloroethylene.   Revista Igiena Bact.
     Virus Parazit. Epidem. Pneum.  Igiena 24(3):147-151.  (translation).

Dumortier, L.,  G. Nicolas, and F.  Nicolas. 1964. A case of hepato-nephritis
     syndrome due to perchloroethylene.   Arch. Mai. Prof. 25:519-522 (English
     translation)

Duprat, P.,  L.  Delsaut,  and D. Gradiski.  1976. Irritant potency of the princi-
     pal aliphatic chloride solvents on the skin and ocular mucous membranes
     of rabbits.   Europ. J. Toxicol.   3:171-177.

Dybing, F.,  and  0.  Dybing.  1946.  The toxic effect of tetrachloroethane and
     tetrachloroethylene in oily solution.  Acta Pharmacol.  2:223-226, 1946.

Eberhardt, H.,  and K.  J. Freundt.  1966.  Tetrachloroethylene poisoning.  Arch.
     Toxikol.  (Berlin) 21:338-351.   (English translation)

Fishbein, L.  1976.  Industrial mutagens and potential mutagens.  I.  Halogenated
     aliphatic derivatives.  Mutat. Res.   32:267-308.

Franke, W.,  and  F.  Eggeling. 1969.  Clinical and statistical studies on em-
     ployees of  chemical cleaning plants exposed to perchloroethylene.  Med.
     Welt 9:453-460.   (English translation)

Friborska, A.  1969. The phosphatases of peripheral white blood cells in
     workers exposed to trichloroethylene and perchloroethylene.  Br. Med.  J.
     Ind. Med.  26:159-161.
                                    6-28

-------
Fujii, T. 1975. The variation in the liver function of rabbits after admini-
     stration of chlorinated hydrocarbons.  Jpn. J. Ind. Health  17:81-88.
     (English translation)

Fuller, B. B. 1976. Air pollution "assessment of tetrachloroethylene."  Mitre
     Technical Report - 7143.  February.

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

Gold, J. H.  1969.  Chronic perch!oroethylene poisoning.  Can.  Psychiatric
     Assoc.  J. 14:627-630.

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

Hake, C. L.,  and R. D. Stewart. 1977. Human exposure to tetrachloroethylene:
     inhalation and skin contact.  Environ. Health Persp. 21:231-238.

Handbook of Toxicology, Volumes II-V, Philadelphia: 1959. W.  B. Saunders Co.
     Volume V., p. 76.

Hughes, J. P. 1954. Hazardous exposure  to a so-called safe  solvent.  J. Am.
     Med. Assoc. 156:234-237.

Klaassen, C.  D., and G. L.  Plaa. 1966.  Relative effects of  various chlorinated
     hydrocarbons on liver  and kidney function in mice.  Toxicol. Appl. Phar-
     macol.   9:139-151.

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.

Kylin, B., I. Sumegi, and S. Yllner. 1965. Hepatotoxicity of  inhaled tri-
     chloroethylene and tetrachloroethylene - long-term exposure.  Acta
     Pharmacol. Toxicol. (Kbh).  22:379-385.

Kylin, B., H. Reichard, I.  Sumegi, and  S. Yllner. 1963. Hepatotoxicity of
     inhaled trichloroethylene, tetrachloroethylene, and chloroform—single
     exposure.  Acta Pharmacol. Toxicol.  20:16-26.

Lamson, P. D. ,  B. H. Robbins, and C. B. Ward. 1929. The pharmacology and
     toxicology of tetrachloroethylene.  Am. J. Hyg.  9:430-444.

Larson, N. A., B.  Nielsen,  and A. Ravin-Nielsen. 1977. Perch!oroethylene
     intoxication—a hazard in the use  of coin laundries.   Ugeskr. Laeg.
     39(5):270-275.  (English translation)

Lehmann, K.  B. 1911. Experimental studies on the influence  of technically  and
     hygienically important gases and vapors on the organism.  Arch. Hyg.
     74:1-60.  (German)
                                    6-29

-------
Lehmann,  K.  B.,  and L.  Schmidt-Kehl.  1936.  The 13 most Important chlorinated
     hydrocarbons of the aliphatic series from the standpoint of occupational
     hygiene.   Arch.  Hyg.  116:131-268.   (German)

Levine  BMP  Fierro,  S.  W.  Goza, and J. C. Valentour. 1981. A tetrachloro-
     ethylene fatality.   J.  Forensic Sci. 26(1):206-209.

Ling, S. ,  and W. A. Lindsay.  1971. Perchloroethylene burns.  Brit. Med. J.
     3(5766):115.

Lob, M.  1957.  Dangers of perchloroethylene.   Arch. Gewerbepath. Gewerbehyg.
     16:45-52.   (English translation)

Mazza, V.  1972.  Enzyme changes in experimental tetrachloroethylene intoxica-
     tion.  Folia Med.   55(9-10):373-381.  (English translation)

Mazza, V., and A. Brancaccio. 1971. Adrenal  cortical and medullar hormones  in
     experimental tetrachloroethylene poisoning.  Folia Med.  54:204-207.
     (English translation)

Meckler,  L.  C.,  and D.  K.  Phelps. 1966.  Liver disease secondary to tetra-
     chloroethylene exposure.  J. Am. Med.  Assoc. 197(8):144-145.

Method,  H. C.  1946 Toxicity of tetrachloroethylene.  J. Am. Med. Assoc.
     131:1468.

Miller,  T. A.  1966. Anthelmintic activity of tetrachloroethylene against
     various stages of Ancylostoma cam'urn in young dogs.  Am. J. Vet.  Res.
     27(119):  1037-1040.

Moeschlin, S.  1965. Poisons,  diagnosis and treatment.  New York, N.Y.:
     Gruner and Stratton.

Morgan,  B. 1969. Dangers of perchloroethylene.  Brit. Med. J. 2:513.

National  Institute for Occupational Safety and Health. 1976. Criteria  for  a
     recommended standard...occupational exposure to tetrachloroethylene
     (perchloroethylene).   HEW publication no. (NIOSH) 76-185.  U.S. Department
     of HEW, PHS, CDC, NIOSH.  July.

Navrotskii,  V.  K., Kashin, I. L. Kulinskaya, L. F. Mikhaylovskaya, L.  M.
     Shuster,  I. N. Burlaka-Vouk, and B. V.  Zadorozhniy. 1971.  Comparative
     assessment of the toxicity of a number of industrial poisons when inhaled
     in low concentrations for prolonged periods.  Trudy S'ezda Gigen  Vkramkoi:
     SSR 8:224-226.  (translation)

Parker,  J. C.,  L. J. Bahlman, N. A. Feidel, H. P. Stein, A. W.  Thomas, B.  S.
     Woolf,  and E. J. Baier.  1978. Tetrachloroethylene  (perchloroethylene).
     Current NIOSH Intelligence Bulletin #20.  Am. Ind.  Hyg. Assoc.  J.  39:3.

Patel, R., N.  Janakiraman, and W. D. Towne.  1977. Pulmonary  edema due  to
     tetrachloroethylene.   Environ. Health  Persp. 21:247-249.
                                    6-30

-------
Plaa, G. L., and  R.  E.  Larson.  1965.  Relative nephrotoxic properties of
     chlorinated  methane,  ethane,  and ethylene derivatives in mice.   Toxicol.
     Appl.  Pharmacol.   7:37-44.

"Plaa, G. L., E. A.  Evans,  and  C.  H.  Mine.  1958.  Relative hepatotoxicity of
     seven  halogenated  hydrocarbons.   J.  Pharmacol.  Expt.  Ther.   123:224-229.

Reinhardt,  C.  F.,  L.  S.  Mullin,  and  M.  B.  Maxfield.  1973.  Epinephrine-induced
     cardiac arrhythmia potential  of some common industrial  solvents.   J.
     Occup. Med.  15:953-955.

Rowe, V. K.,   D.  D.  McCollister,  H.  C.  Spencer,  E.  M.  Adams,  and D.  D.  Irish.
     1952.  Vapor  toxicity  of tetrachloroethylene for laboratory animals and
     human  subjects.  Arch. Ind.  Hyg.  Occup.  Med.   5:566-579.

Saland, G.  1967.  Accidental exposure to perchloroethylene.   N.  Y.  State J.
     Med. 67:2359-2361.

Savolainen, H.,   P.  Pfaffli, M.  Tegen,  and H.  Vainio.  1977.  Biochemical and
     behavioral effects of inhalation exposure to tetrachloroethylene and
     dichloromethane.   J.  Neuropathol.  Exp.  Neurol.   36:941-949.

Schlingman, A.  S.  1925.  Critical  tests of tetrachloroethylene:  a new anthel-
     mintic with  special reference to its use in puppies.   J.  Am.  Vet.  Med.
     Assoc. 68(2):225-231.

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

Snow, D. H. 1973.  The effects  of pyrantel  pamoate and tetrachloroethylene on
     several blood enzyme  levels in  the greyhound.   Aust.  Vet.  J.   49:269-272.

Stewart, R. D.  1969.  Acute tetrachloroethylene intoxication.   J. Am. Med.
     Assoc. 208(8):1490-1492.

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., D.  S. Erley, A.  W. Schaffer,  and H.  H. Gay.  1961a. Accidental
     vapor  exposure to  anesthetic concentrations of a solvent containing
     tetrachloroethylene.   Ind.  Med. Surg.  30:327-330.

Stewart, R. D., H.  H. Gay, D.  S.  Erley, C. L.  Hake, and A.  W.  Schaffer. 1961b.
     Human  exposure to  tetrachloroethylene vapor.   Arch. Env.  Health.  2:40-46.

Stewart, R. D.,  E.  D. Baretta, H.  C. Dodd, and T.  R. Torkelson.  1970.  Experi-
     mental human exposure to  tetrachloroethylene.   Arch.  Environ. Health
     20:224-229.
                                     6-31

-------
Stewart, R. D.,  C. L. Hake, H. V. Forster, A. J. Lebrum, J. E.  Peterson,  and
     A. Wu. 1974.  Tetrachloroethylene:  development of a biological  standard
     for the industrial worker by breath analysis.  Report no.  NIOSH-MCOW-ENVM-
     PCE-74-6.  National Institute of Occupational Safety and Health.

Stewart, R. D.,  C. L. Hake, A. Wu, J. Kalbfleisch, P. E. Newton, S.  K.  Marloro,
     and M. V. Salama. 1977. Effects of perch!oroethylene/drug  interaction  on
     behavior and neurological function.  Final report, National Institute  for
     Occupational  Safety and Health, April.

Trense, E., and H. Zimmerman.  1969.  Fatal inhalation poisoning  with  chroni-
     cally-acting tetrachloroethylene vapors.  Zbl. Arbeitsmed. 19:  131-137.
     (English translation)

U.S. Environmental Protection Agency. 1977.  An assessment of the need  for
     limitations on trichloroethylene, methyl chloroform, and perchloroethylene.
     Draft final report, Volumes I,  II, III.  Office of Toxic Substances.   EPA
     contract no.  68-01-4121.

U. S. Environmental Protection Agency. 1980. Ambient water quality criteria
     for tetrachloroethylene.   EPA 440/5-80-073.

Walter, P., A. Craigmill, J. Villaume, S. Sweeney, and G. L. Miller. 1976.
     "Chlorinated hydrocarbon toxicity (1,1,1-trichloroethane,  trichloroethylene
     and tetrachloroethylene)" - a monograph.  Nat. Tech. Infor. Serv.
     Springfield,  Va.  PB-257 185/9, May.

Weichardt,  H, and J.  Under. 1975. Health hazards caused by perchloroethylene
     in dry-cleaning plants from the point of view of occupational medicine
     and toxicology.   Staub-Reinhalt.  Luft 35(11):416-420.  (English  transla-
     tion)

Wenzel, D.  G., and R. D. Gibson. 1951. Toxicity and anthelmintic activity of
     n-butylidine chloride.  J.  Pharm. Pharmacol.  3:169-176.

Withey, R.  J., and J. W. Hall. 1975. The joint action of perchloroethylene
     with benzene or toluene in rats.  Toxicol.  4:5-15.
                                    6-32

-------
         7.  TERATOGENICITY, EMBRYOTOXICITY, AND REPRODUCTIVE EFFECTS


     Because of  its widespread  use,  PCE has been  studied for teratogenic

potential.   Teratology studies have been performed in rats, mice, and rabbits,

using doses of PCE  which,  in some studies,  produced slight signs of maternal

toxicity.   Other  studies in  chicken  embryos (Elovaara  et al.,  1979) have

indicated that PCE  disrupts  embryogenesis in a  dose-related manner.   However,

since administration of  PCE  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 PCE 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 "teratogenic11 is  defined  as the tendency to
     produce physical  and/or functional  defects in offspring HI 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 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 cate-
     gories.   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 neces-
     sarily 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 subcate-
     gory 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 carcinogenidty)  which are attributable to i_n utero
     exposure.   The  third category would be embryo or fetal toxicity as com-
     prised 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.


                                    7-1

-------
          Two major problems  with  a definitional  scheme of this nature must be
     pointed out,  however.  The  first is  that the reversibility of any phenom-
     enon 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 irreversible manner.
7.1  ANIMAL STUDIES
7.1.1  Mice
     Schwetz et al.  (1975) reported  a  study of Swiss-Webster mice exposed to
                                                          3
PCE via inhalation  at  concentrations  of 300  ppm (2034 mg/m ) for 7 hours daily
on days 6 through 15 of presumed gestation.   Day 0 of gestation was designated
the day a  vaginal plug was  observed.  This concentration was cited as twice
the maximum allowable exclusion limit for human industrial exposure, with the
                          ®                      3
Threshold Limit Value  (TLV)   of 100 ppm (678 mg/m ).   Concurrent controls were
exposed to filtered  air.
     Following maternal  Caesarean-sectioning  on  day  18 of  gestation,  all
fetuses were  examined for external  anomalies.   One-half  the  fetuses  were
examined for soft tissue malformations  using a free-hand sectioning technique,
and the other one-half of the fetuses were cleared, stained, and examined for
skeletal malformations.  One fetus in each litter was processed and examined,
using  histopathological  techniques.   Seventeen litters were  examined.   The
pups in the exposed  group were  significantly smaller, as measured by decreases
in fetal body weight.   Also,  slight but not  statistically significant increases
in the  number  of runts were  observed,  as well  as sporadic increases in the
numbers of fetuses  with subcutaneous  edema,  delayed ossification of the skull,
delayed ossification  of  the  sternebrae,  and splits  in  sternebrae.  No other
remarkable malformations were reported in fetuses.   Increases in the absolute
and relative mean maternal liver weights were reported.  No evidence of  tera-
togenicity of PCE was  found  at  the concentration tested.

7.1.2  Rats
     Schwetz et al.  (1975) also administered PCE to  Sprague-Dawley  rats by
                              3
inhalation (300 ppm;  2034 mg/m  ) for 7 hours daily, on  days 6 to 15 of gesta-
tion.   Control rats were exposed to  filtered air.   Day 0  of gestation  was
                                    7-2

-------
designated  as  the day when spermatozoa were  observed in smears of  vaginal
contents.   Rats were  sacrificed on day 21  of  gestation.  Caesarean  sections
were performed, and  fetuses  were examined  for  external  malformations.   Half
of the  fetuses  in  each  litter were  examined for soft  tissue  malformations  and
the remaining  half of the fetuses were examined for  skeletal  malformations.
One fetus  from  each  litter was  randomly  selected for serial sectioning and
histological evaluation.   Average maternal body weight  gain  was slightly
reduced  in  the  rats  exposed to  PCE.   A slight but  statistically  significant
increase in  resorption  was  reported  in 9  of 17 PCE-exposed  litters evaluated.
Exposure of  dams  to  PCE produced no  effect  on  the average number  of  implanta-
tions  per  litter, fetal  sex ratios,  or  fetal  body measurements.   No soft
tissue or skeletal anomalies were reported  in the offspring of rats exposed to
PCE.   No evidence of teratogenicity of PCE was found at the  concentration
tested.
     Bellies et al.  (1980) exposed  Sprague-Dawley  rats  to  PCE at 300 ppm
(2034 mg/m3) for  7 hours daily, 5 days per week.   Controls were administered
filtered air.   Nineteen to 24  rats were  examined in this study.  One-half of
the rats were  exposed for 3 weeks  prior  to mating.  All rats  were  exposed
during  gestation  either between days  0 through 18  or between  days 6  through
18, with  day 0 designated as  the day when spermatozoa were  observed  in  smears
of vaginal contents.  Three rats (days 6-18) died on the  second day of preges-
tational treatment.   Signs  of  ataxia and  loss  of balance were  observed  in  all
of  the  other rats of this group on the   same day.  The  authors thought this
response was most likely due to the high levels in the  inhalation  chamber
during  the  last 2 hours of the  day.  Measurements taken  15  minutes before  the
end of  the exposure showed 568 ppm  (4061 mg/m3) but  could  have been higher
previously.  The  maternal body weight gain in the  PCE-treated rats was not
statistically different  from that  in controls  during the  pregestational period.
Inhibition  of  maternal  body weight gain  occurred during the first week,  as
well as  increases  in mean absolute, but not relative,  kidney and liver weights.
No  embryotoxic effects  were  observed  which  were attributable to maternal
exposure to  PCE,  except  for delays  in skeletal ossification.   This effect,
however, is thought to  be a reversible effect  and is  not considered a malforma-
tion as  such.   No  teratogenic effects  were  observed.
                                    7-3

-------
     Nelson et al.  (1980) evaluated the ability of PCE to elicit a behavioral
teratogenic response in Sprague-Dawley  rats.   In a pilot dose-range  finding
study, groups of 3  rats were exposed by inhalation to 1800 or 3600 ppm (12,204
              3
or 24,408 mg/m ) PCE.   Narcotization was  observed in dams.   Therefore, a
concentration of 900 ppm (6102 mg/m ) PCE was used in the pilot study.   Signs
of maternal  toxicity  such  as severe reductions  in  average  food intake  and
decreased average body  weight  gain  were observed.   Food consumption and body
weight gain were also  reduced  but were not statistically significant in rats
                             3
exposed to 900 ppm  (6102 mg/m  )  PCE on days 14 to 20 of gestation.   Dams and
pups  exposed  to  100 ppm (678 mg/3)  PCE on days 14 to 20 of gestation did not
show any adverse effects as compared to the controls.
     In the behavioral  testing  study,  102 pregnant rats  were exposed to PCE by
inhalation as follows:

     (1)  900 ppm,  days 7-13 of gestation (N=19)
     (2)  900 ppm,  days 14-20 of gestation (N=21)
     (3)  sham-exposed, days 7-13 of gestation (N=13)
     (4)  sham-exposed, days 14-20 of gestation (N=19)
     (5)  100 ppm,  days 14-20 of gestation (N=15)
     (6)  sham exposed, days 14-20 of gestation (N=15)

     Seven behavioral  tests were performed on days 4 through 46 postparturition,
using one male and one  female per litter.  One group of  rats,  consisting of a
male/female pair from each  litter, was tested for ascent on a wire mesh screen
and rotorod balancing.   One male and one female rat per  litter were tested for
open-field activity, activity wheel, and avoidance conditioning.   A third pair
was  tested  for operant  conditioning.   Catecholamines  (norepinephrine  and
dopamine), acetylcholine,  and  protein, measured  in  the  brain tissue, were
evaluated in 10 pups (no more than 2 per litter per treatment group) at birth,
or at  21  days postparturition.   Histopathological evaluation  of  brains  for
neuropathology was  performed in an unreported number of  pups.
                                                  3
     Rats from dams exposed to 900 ppm (6102 mg/m ) PCE  on days  7 to 13 of
gestation, performed  less  well  on  discrete testing of  ascent and rotorod
tests, but only  on  certain days  of testing.   Offspring  exposed to 900  ppm
          3
(6102 mg/m )  on days 14 to  20 of gestation performed less well on one test day
in the ascent  test,  but later  performed better than controls  in the  rotorod
test, and were  relatively  more active than controls in  the open-field tests.
Acetylcholine levels were reduced in 21-day-old rats of  dams exposed to 900 ppm

                                    7-4

-------
          3
(6102 mg/m ) PCE.  Dopamine  levels  were reduced in  rats  of dams exposed on
days 7 to 13 of gestation.
                                                      3
     In the group of rats exposed to 100 ppm (678 mg/m ) PCE during days 14-20
of gestation,  no  significant differences were observed between the offspring
of these  animals and their  controls  on any of  the  behavioral  tests.   The
authors summarized that  "there were generally  few behavioral or  neurochemical
                                                                               3
differences observed between offspring of animals exposed to 900 ppm (6102 mg/m )
PCE during  either  days  7-13 or 14-20 of gestation.   When significant differ-
ences did  appear,  they occurred more often when  the  group exposed  during days
14-20 of gestation was compared with its control group."
     It should be  noted  that the field  of  behavioral  teratology is in its
early stages of development (Buelke-Sam and Kimmel,  1979), and such behavioral
alterations cannot at this  time  be interpreted in terms of human effect.   It
also must  be  noted that these observed  changes  may  be related to  maternal
toxicity and thus do not represent a direct toxic effect.
     Tepe  et al.   (unpublished, 1982) exposed Long-Evans hooded female rats to
                               3
1000 ± 125 ppm (6780 ± 847 mg/m ) PCE  vapors  to ascertain if exposure before
mating and  during  pregnancy was  more detrimental to the embryo than exposure
during pregnancy alone.   Four treatment groups (30 rats each) were utilized in
a two-by-two factorial design:  exposure to PCE  for  two weeks (6 hours daily,
5 days per week)  before mating, through day 20 of pregnancy; PCE before mating
and filtered air during pregnancy; filtered air before and PCE during pregnancy;
and filtered air  throughout.   One-half the dams in each treatment group were
sacrificed  on  day  21 by ether anesthesia.   Elevations  1n relative maternal
liver weights and reduction in fetal body weight were observed without altera-
tion in maternal  weight  gains in groups exposed during pregnancy.   An excess
of skeletal variations was  seen in  the  group exposed before  mating  and during
pregnancy; and excessive soft tissue variations (e.g. kidney dysplasia) occurred
in the group  exposed during pregnancy alone.   These effects are consistent
with embryotoxlcity.   Elevation   in ethoxycoumarein  dealkylase  activity (an
indicator  of  P.™ activity) was   observed  in maternal  livers but not fetal
livers with pregnancy exposure.   Ethoxyresorufin dealkylase  activity  (an
indicator  of  P,4g activity)  was  not elevated  in maternal  livers  with PCE
exposure and was  not detectable   in fetal  livers.  No  other  effects were ob-
served in pregnant rats exposed for 6 hr/day on days 0-20 of gestation.
     Manson et al. (unpublished,  1982) conducted a study on postnatal evaluation
of offspring of  the  female rats  exposed to 1000 ± 125 ppm (6780 ± 847 mg/m )
                                    7-5

-------
PCE  before  and/or during  pregnancy.   The four  treatment  groups are those
described above  (Tepe  et al.,  1982).   The  purpose  of  the postnatal  evaluation
was  to  determine if  the reduction in  body weight  or  the excess  skeletal  and
soft  tissue  variants observed in term  fetuses  from the Tepe  et al.  (1982)
study persisted,  and if PCE  possessed  transplacental  carcinogenic activity or
neurobehavioral  toxicity.   Weight gain and  survival  of offspring  up to  18
months  of age,  and frequency of  any gross  lesions  observed at 6- and 18-month
autopsies, were  not  influenced by prenatal exposure to PCE.   Neurobehavioral
tests of general activity in open field tests at 10 and 20 days  of  age, and  in
running wheels  from  40 to  100  days of  age, did  not indicate  treatment-related
effects.  Likewise,  results  from the  visual  discrimination test on  offspring
from  130  to  170 days of age were negative.  Prenatal exposure to PCE did  not
exert a detrimental  effect on any of  the  parameters  of postnatal maturation
examined.

7.1.3   Rabbits
     Beliles  et  al.  (1980)  also  investigated the teratogenic  potential  of  PCE
in  rabbits.   Fifteen to 22  rabbits were used in  this  study.   The rabbits were
divided into  six groups as follows:

                                                 Days of Exposure
Group          Cone  (ppm)             PregestationalGestational
1
2
3
4
5
6
0 (control)
0 (control)
500
500
500
500
none
5 days/week;
none
5 days/week;
none
5 days/week;

3 weeks

3 weeks

3 weeks
0-21
0-21
0-21
0-21
0-21
0-21
After the pregestational exposure period was complete, rabbits from each group
were mated.  A positive identification of spermatozoa in the vaginal canal was
taken as  evidence  of mating and designated  as  day 0 of gestation.  The two
control  groups  (1  and  2) were  exposed  to  filtered  air  and  the  exposure groups
(3, 4, 5, and 6) to 500 ppm (3390 mg/m ) PCE, 7 hours per day.
     Mean body  weights  of  rabbits  during  pregestational  and gestational  expo-
sure indicated  no  significant  difference  between controls  and  treated  groups.
Reduced food consumption during the approximate period of days 10 through 22
of gestation was  observed  in groups 3 and 5 and may have been related to PCE

                                    7-6

-------
             TABLE 7-1.  SUMMARY OF REPRODUCTIVE/TERATOGENIC EFFECTS OF PCE IN LABORATORY ANIMALS
   Study
  TCI Purity
   Species
      Experimental
       Conditions
   Study
      Results
Beliles et al. ,
1980   *
Hanson et al. ,
1982
Technical grade
91.43%
CD rats, New
Zealand rabbits
Elovaara et al.,   99% pure
1979              dissolved in
                  olive oil
Technical grade
(Dow Chemical
 Company)
Nelson et al.,     Technical grade
1980              98.5% pure
                   White Leghorn
                   chick embryos
Long-Evans
hooded rats
                   Sprague-Dawley
                   rats
Schwetz et al.,   Dow-Per
1975              99.99% pure
                   Sprague-Dawley
                   rats, Swiss-
                   Webster mice
Inhalation, 500 ppm
before, during
gestation, and both
before and after
gestation.
Injection into air
space 5 to 10 umole/
egg on days 2,3,6 of
incubation, examined 14
days after incubation.

Inhalation, 1000 ppm
before, during
gestation, and both
before and during
gestation.

Inhalation, 100 or
900 ppm on various
days of gestation.
                  Inhalation, 300 ppm
                  on days 6-15 of
                  gestation.
Teratology
study
                                             Chick egg
                                             teratology
Postnatal
function
evaluations
                                             Postnatal
                                             behavioral
                                             testing
                           Teratology
                           study
Rats maternal
toxicity, changes in
liver weight, embryo
toxicity, reduced
ossification; rabbits:
variations in size
and color of rabbit
placenta.

Malformed embryos at
highest dose, de-
creased embryo length.
No detrimental
effects in off-
spring observed up
until 8 months
after birth.

Maternal weight
loss at highest
dose, slight,
subtle behavioral
changes in off-
spring.

Maternal toxicity,
embryo toxicity
(lowered weight of
mice, increased re-
sorption in rats,
subcutaneous edema
in mice offspring).

-------
I
oo
                                                    TABLE 7-1.   (Continued)
Study
Tepe et al . ,
1982
TCI Purity
Technical grade
(Dow Chemical
Company)
Species
Long-Evans
hooded rats
Experimental
Conditions
Inhalation, 1000 ppm
before, during
gestation, and both
before and during
gestation.
Study
Teratology
study
Results
Maternal toxicity,
embryo toxicity,
depressed fetal
weight, skeletal
and soft tissue
variations.

-------
exposure.  Placenta! abnormalities were reported at all exposure levels (varia-
tion in  size  and  color);  however,  these were  not  statistically  significant as
tested by  rank sum analysis  and  were thought to  reflect  changes  in a few
litters.   In addition, histopathologic evaluation  failed to reveal any signifi-
cant change in the placenta.

7.2  SUMMARY
     The mammalian animal tests performed to date  do not indicate any signifi-
cant teratogenic  potential  of PCE.  On this basis, there  is no evidence that
suggest that the conceptus  is uniquely susceptible to the  effects of PCE.   The
anatomical effects observed primarily reflect delayed development and generally
can be considered  reversible.   The minor  behavioral changes observed probably
reflect maternal  nutritional  deprivation  rather than a direct effect of PCE.
It  is  important to note,  however,  that  the  reversible  nature of an embryonic/
fetal effect in one species might,  in another species, be  manifested in a more
serious and  irreversible  manner.   PCE also has been  implicated in producing
adverse germ  cell  effects with alteration  in  sperm  morphology  (see  Section
8.1).  The  potential  of PCE to cause  adverse  effects  on  reproduction or the
developing conceptus is based on a  limited  number  of studies.   The teratogenic
potential of PCE for humans is unknown.
                                    7-9

-------
7.2  REFERENCES


Bellies, R.  P.; Brusick,  D.  J.;  Mecler, F. J. 1980.  Teratogenic-mutagenic
     risk of workplace contaminants:   trichloroethylene, perchloroethylene,  and
     carbon disulfide.  U.S.  Department of Health Education and Welfare,  Contract
     No. 210-77-0047.

Buelke-Sam, J.; Kimmel,  C.  A.  1979.   Development and standardization  of
     screening methods for behavioral teratology.  Teratology 20:17-30.

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

Lanham, S.   1970.  Studies on placental  transfer:  trichloroethylene.   Ind. Med.
     39:46-49.

Manson, J.  M. ; Tepe, S.  J. ;  Lowrey,  B.; L. Hastings.  1982.  Postnatal evalua-
     tion of offspring exposed prenatally to perchloroethylene.  Unpublished.

Nelson, B.  K. ; Taylor, B. J.;  Setzer, J. V.; Hornung, R. W. 1980.  Behavioral
     teratology of perchloroethylene in rats.  J. Environ. Pathol. Toxicol.
     3:233-250.

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

Tepe, S. J.; Dorfmueller,  M.  K.;  York, R. G. ; J. M. Manson.  1982.   Teratogenic
     evaluation of perchloroethylene in rats.  Unpublished.

U.S. Environmental Protection  Agency.  Proposed guidelines for registering
     pesticides in the United  States.  FR (1978 August 22) 43:37382-37388.

U. S. Environmental  Protection Agency.   Proposed health effects test  standards
     for Toxic Substances Control  Act test rules and proposed good laboratory
     practice standards  for  health effects.   FR (1979 July 26) 44:44089-44092.

U.S. Environmental Protection  Agency.  Determination not to initiate  a rebut-
     table  presumption against registration (RPAR) of pesticide products  con-
     taining carbaryl  availability of decision document.  FR (1980 December
     12) 45:81869-81876.
                                    7-10

-------
                              8.   MUTAGENICITY

     The objective of this mutagenicity evaluation is to determine whether PCE
has the potential  to  cause mutations in  germ  and somatic cells of  humans.
This qualitative  assessment  is  based on data derived from several short-term
tests that  measure different types  of  genetic alterations:  gene mutation,
chromosomal  aberrations,  unscheduled  DMA  synthesis,  and mitotic  recombination
(Table 8-1).  These tests were  conducted using bacteria,  Drosophila, yeast,
cultured mammalian cells,  whole mammal  systems,  and  cytogenetic analyses of
exposed humans.    Consideration  will  also be given to studies concerning the
mutagenicity of known and expected metabolites.

8.1  GENE MUTATION TESTS
8.1.1  Bacteria
     The ability  of  PCE to cause gene mutations  in bacteria has been studied
by several  investigators.  Many of these investigators used the Ames Salmonella/
microsome test  or modifications  of  that test.   Different purities  of  PCE
(stabilized* and  low-stabilized materials)  have been evaluated.  (Bacterial
studies are summarized in Table 8-2.)
     Tetrachloroethylene  is  a volatile  chemical,  and  thus, the standard Ames/
Salmonella plate test, in which precautions are not taken to prevent escape  of
evaporated material,  is  not  entirely suitable for its testing.   Williams and
Shimada  (1983,  sponsored  by  PPG Industries,  Inc.),  however,  modified  the
standard plate  procedure  by exposing the bacteria to the test  agent  in a
sealed chamber.   In their procedure,  a known volume of test chemical was added
to a glass petri plate containing a magnetic bar  to ensure continuous stirring
for even dissipation  of vapors.  The chamber was initially incubated at room
temperature for 20 minutes,  and then exposure  was  continued  at 37°C for 18
hours, after which the bacterial test plates  were removed from  the  chamber,
covered with lids, and incubated another 30-54 hours at 37°C.
     "Stabilization  is  the  intentional  addition of material  to  increase  the
stability of  tetrachloroethylene.   Typically,  the added stabilizers are acid
and  free  radical  scavengers (Dr. A.  Philip  Leber of PPG  Industries,  Inc.,
personal communication, September 1983).
                                    8-1

-------
       TABLE  8-1.   SUMMARY OF MUTAGENICITY TESTING  OF  TETRACHLOROETHYLENE
 A.   Gene  Mutation  Tests

     Ames/Salmonella  assay
                                         Results*
     +** _
     Escherichia  coli  K12/343/113  (  )
     Multi-purpose  test

     Saccharomyces  cerevisiae  D7 reverse
     mutation test  (ivl-1  locus)

     Drosophila sex-linked recessive
     lethal  test

     Host-mediated  tests  in mice:  bacteria  wk
                                 yeast
 B.   Chromosomal  Aberration  Tests
     Rat bone marrow assay
     Mouse bone marrow assay
     Peripheral  lymphocytes  from
     exposed humans

     Drosophila  sex chromosome  loss
     assay

     Rat dominant lethal  assay
      References

Williams and Shimada 1983,
Margard 1978, SRI Inter-
national 1983, NTP 1983,
Bartsch et al. 1979, Cerna
and Kypenova 1977 (abstract)

Henschler 1977, Greim et al.
1975

Call en et al. 1980,
Bronzetti et al.  1983

Beliles et al. 1980
                  Beliles et al.  1980,
                  Cerna and Kypenova 1977
                  (abstract)

                  Bronzetti et al.  1983
                  Rampy et al.  1978,
                  Beliles et al.  1980

                  Cerna and Kypenova
                  1977 (abstract)

                  Ikeda et al.  1980
                  Beliles et al.  1980
                  Beliles et al.  1980
 C.   Other Tests  Indicative  of  DNA  Damaging Activity

     Unscheduled  DNA  synthesis  in WI-38     -"*"

     Hepatocyte primary  culture/DMA repair  +,•
     test
     Mitotic  recombination tests  in
     Saccharomyces  cerevisiae D7
                  Beliles et al.  1980

                Williams and Shimada
                  1983, Williams 1983

                  Call en et al. 1980,
                  Bronzetti et al.  1983
D.  DNA Binding Studies

    Whole mice

E.  Germ Cell Tests

    Altered sperm morphology
mouse
  +
        rat
                  Schumann et al.  1980
                  Beliles et al.  1980
     * + designates positive;   negative; wk weak  response.  Dose-response
relationships were not established for the reported +  results or wk  results.

     **Although increases several fold over background were observed, the positive
results are considered weak because large amounts  of material were needed to
elicit the responses.  Positive results were only  obtained using airtight
chambers (except for the study by Cerna and Kypenova 1977).
      TQuestionable evidence for weak or borderline activity in specific data sets.

     ttPositive results were found with vapor phase exposure and negative results
were obtained using conventional phase exposure.
                                       8-2

-------
            TABLE 8-2.   RESULTS OF BACTERIAL TESTS OF DIFFERENT PURITIES  AND  SOURCES  OF  TETRACHLOROETHYLENE
Test system/strain Purity/source
Ames/Salmonella Perchlor 200-
TA98, TA100, TA1535 low-stabil ized
99.93% purity
PPG Industries,
Inc.
m Ames/Salmonella Perchlor 230-
' TA98, TA100, TA1535 stabilized
TA1537, TA1538 99.80% purity
PPG Industries,
Inc.
Ames/Salmonella High purity
TA100, TA1535 Perchlor
low-stabil ized,
99.98+% purity,
PPG Industries,
Inc.
Concentrations
tested
1% v/v for TA98,
TA1538, and
TA1537; 0.1
1.0, 2.5, 5.0,
7.5, and 10%
for TA100 and
TA1535
1% v/v for TA98,
TA1538, and
TA1537; 0.1
1.0, 2.5, 5.0,
7.5, and 10%
for TA100 and
TA1535
0.1, 1.0, and
2.5% v/v
Metabol ic
activation
Aroclor-
induced rat
S9 mix
Aroclor-
induced rat
S9 mix
Aroclor-
induced rat
S9 mix
Protocol
Gas-phase
exposure in
airtight
chambers
Gas-phase
exposure in
airtight
chambers
Gas-phase
exposure
in airtight
chambers
Reported
result Reference
Positive Williams
at 2.5% (>97% and
toxicity) in Shimada
base-pair 1983
substitution-
sensitive strain
(two to tenfold
i ncreases
+/- M.A.*)
dose-response
not established
Positive Williams
at 2.5% (>97% and
toxicity) Shimada
in base pair 1983
substitution
sensitive
strains (three to
ten-fold increase
+/- M.A.)
dose-response
not established
Negative Williams
and
Shimada
1983
     *+/- M.A. designates response similar in the presence and absence of
metabolic activation
(continued on the following page)

-------
                                                    TABLE 8-2.  (continued)
Test system/strain
Ames/Salmonel la
TA98, TA100, TA1535
TA1537, TA1538
Ames/Salmonel la
TA98, TA100, TA1535
TA1537, TA1538
Purity/source
Nonstabi 1 ized
high purity
Detrex
Industries, Inc.
Stabi 1 ized
99.84% purity
Detrex
Industries, Inc.
Concentrations
tested
0.01, 0.05,
and 0.1 ml/
plate
0.01, 0.05,
and 0.1 ml/
plate
Metabol ic
activation
Arocl or-
induced rat
S9 mix
Aroclor-
induced rat
S9 mix
Protocol
Standard
plate test
in airtight
chambers
Standard
plate test
in airtight
chambers
Reported
result
Negati ve
Positive
(twofol d
increases
in frameshi
Reference
Margard
1978
Margard
1978
ft-
oo
                                                                      sensitive
                                                                      strains and
                                                                      TA100) at 0.1
                                                                      ml/plate (160
                                                                      mg/plate).  >90%
                                                                      toxicity.  S9 mix
                                                                      increased response
                                                                      (10-to 17-fold
                                                                      increases)
   Ames/Salmonella
   TA98, TA100, TA1535
   TA1537
99+% purity
Aldrich
0.025, 0.05,
0.1, 0.5, 1.0
and 1.5 added
to petri plate
at bottom of
desiccator
Aroclor-induced
female and male
rat 1iver S9 mix
and Aroclor-in-
duced female and
male mouse liver
S9 mix
Gas-phase
exposure in
airtight
chambers
Negative
SRI
Inter-
national
1983
   Ames/Salmonella
   TA98, TA100, TA1535
   TA1537
Technical grade    3, 10, 33, 100,
Fisher             333 ug/plate
                  Aroclor-induced
                  rat liver S9 mix
                  Aroclor-induced
                  hamster liver
                  S9 mix
                  Preincu-
                  bation assay
                  10 min. 37°C
               Negative
              NTP 1983
                                                                                    (continued on the following page)

-------
TABLE 8-2.   (continued)
Test system/strain
Ames/Sal monel la
TA100
Ames/Sal monel la
tester strains
not rep'orted
E. coli K12/343/113 (>)
oo
en
Host-mediated assay
ICR mice/Salmonella
TA1950, TA1951,
TA1952
Host-mediated assay
male and female CD
mice/Salmonella
TA98
Purity/source
99.7% purity
Merck-Darmstadt
Not
reported
analytical grade
Merck-Darmstadt
Not
reported
91.43% purity
North Strong
Division
Chemicals
Concentrations
tested
0 to 663
mg/plate
(4x10 M)
Not
reported
0.9 mM
Not
reported
Inhalation at
100 ppm and
500 ppm for
5 consecutive
days
Metabolic
activation
Phenobarbital-
induced mouse
1 iver microsomes
with and without
cofactors

Phenobarbital-
induced mouse
1 iver micro-
somes
Female mice
Male and
female mice
Protocol
Standard
plate test
Spot test
Liquid
suspension
2 hours at
37°C
Not
reported
Bacteria in-
jected intra-
peritoneal ly
after last
exposure and
removed 3
hours
Reported
resul t
Negative
Positive
Negative
Positive
(mortal ity
not
reported)
Positive
(twofold
increases
in revertants
from 100 ppm
in males;
and fourfold
increases
in revertants
from 500 ppm
females).
Reference
Bartsch
et al.
1979
Cerna and
Kypenova
(1977,
abstract)
Greim et
al. 1975
Cerna and
Kypenova
(1977,
abstract)
Beliles
et al.
1980

-------
     In the Williams and Shimada (1983) study, three types of material (provided
by  PPG  Industries,  Inc.)  were tested  in  the  presence  and absence of S9 mix
(derived  from  livers  of Aroclor-induced rats):   Perchlor  200  (low-stabilized,
99.93 percent  purity),  high purity Perchlor  (low-stabilized,  99.98+ percent
purity),  and  Perchlor 230 (stabilized, 99.80 percent purity).*  Perchlor 200
and  Perchlor  230  were evaluated  in  tester strains  TA98, TA1537,  and  TA1538  at
1  percent (v/v) and  in tester strains  TA100 and  TA1535  at 0.1,  1.0,  2.5,  5.0,
7.5, and  10  percent (v/v).   High-purity  Perchlor  was evaluated  in TA100  and
TA1535  at 0.1,  1.0,  and 2.5 percent (v/v). These concentrations  represent the
predicted concentration of  the compound  in the gas phase based upon calcula-
tions that consider the chamber volume and absolute temperature of the chamber,
atmospheric pressure,  and density  of the  test compound.  (For example, 0.06,
0.62,  1.55,  3.1,  4.65,  and 6.2 ml   of PCE was added to the chamber for a 0.1,
1.0, 2.5,  5.0, 7.5, and 5 percent v/v vapor target concentration, respectively.)
Positive  responses  were  obtained for Perchlor 200 and 230 at 2.5 percent v/v
(or 1.55  ml per desiccator) but not for high-purity Perchlor at the concentra-
tions tested.  A  slightly higher response was observed with stabilized material.
Positive  results  were repeated in  a second experiment and were  similar in the
presence  or absence of S9 mix.  For example,  in the absence of S9 mix, Perchlor
200 (at 2.5 percent)  increased the  number of revertant colonies in the base-pair
substitution-sensitive  strains TA1535 (six to tenfold  increases)  and TA100
(two to threefold increases); Perchlor 230 (at  2.5 percent v/v) also caused
increases  in revertant numbers in TA1535  (approximately tenfold increases) and
TA100 (approximately  threefold increases).   (See Figure 8-1 for an illustra-
tion of these responses).   Negative results were found in frameshift-sensitive
strains (TA98  and TA1538) for both Perchlor  200 and  230 in the presence or
absence of S9  mix.   At the next highest  concentration  (5 percent v/v) the
revertant  counts  decreased  to zero as toxicity became an important factor in
the  tests.  The authors also indicated the test materials were toxic (>97
percent killing) at the concentration (i.e.,  2.5 percent v/v) that induced the
mutagenic responses as determined by a simultaneous cytotoxicity test.  Although
     *Perchlor 200 contained  0.012  (percent by weight) of hydroquinone mono-
methyl ether  (HQMME)  which provides  minimal  stabilization;   Perchlor 230
contained 0.011 percent HQMME, 0.07 percent cyclohexene oxide, and 0.05 percent
B-ethoxy propiom'trile;  high-purity Perchlor  contained 0.01 percent  HQMME
(written communication from Dr.  A.  Philip Leber of PPG Industries, Inc., 1983).
                                    8-6

-------
                                   TA 1535
                                                                TA1535
-CD
<
o.
cc
LU
CL


I
               cc
               LU

               UJ
               DC
               m
                    300


                    200


                    100


                     0
                                                 300



                                                 200


                                                 100
                       0.1   1.0
                                2.5
                                         5.0
                                                  7.5
                                                          10.0
                                                                  0.1  1.0
                     TA100
      600


      500


      400


      300


      200


      100
                                                500


                                                400


                                                300



                                                200



                                                100
                                                              TA 100
                       0.1        2.5      5.0       7.5       10.0

                         TETRACHLOROETHYLENE, Stabilized (%V/V)
                                                    0.1 1.0   2.5   '    50        7.5       10.0

                                                     TETRACHLOROETHYLENE, Low Stabilized (%V/V)
Figure  8-1.
Dose-response curves for  Perch!or 200  (low-stabilized tetrachloroethylene) and  Perchlor 230
(stabiliz  tetrachloroethylene) using Salmonella typhimurium  tester strains TA100 and TA1535  in  the
presence of (o	o) and  in the absence  of  (»——•) S9 mix~(150 ul of Aroclor-induced rat  1'ver).
Each data  point represents  the geometric mean  of triplicate  plates from one experiment.   (Williams
and Shimada 1983)

-------
the test materials may  be  quite toxic at 2.5 percent v/v, the investigator's
method of quantifying cytotoxicity may not be accurate because the cell density
that was  used  on each  plate  for  the determination of  toxicity  is several
orders of magnitude  lower  than  that used for determination  of mutagenicity
(i.e., 107-109  cells).   Nevertheless, it should  be noted that because of toxi-
city, the mutagenic  responses observed  for Perchlor 200 and 230 are within a
narrow range of  concentrations.   Because of this narrow  range  of effective
concentrations  and the  concentration increments tested, a clear dose-response
was not demonstrated.
     In the Williams and  Shimada  (1983) study,  high-purity material did not
produce a detectable response under experimental  conditions,  as did the lower-
purity materials  (i.e., Perchlor  200 and 230).   Therefore, the increases  in
revertants observed  may be due to a  contaminant(s).  However, the  high-purity
Perchlor appeared to be more toxic (zero revertants observed at 2.5 percent)
than the other test materials, and it is possible that a weak mutagenic response
may  have  been  masked by its toxicity.  A weak response was observed for high-
purity material  at 1 percent  in TA1535 (with S9 mix), but was not  repeatable.
Also, it  should  be pointed out that  concurrent negative and positive controls
were  not  used  in this  study.   This  weakens the negative conclusions for the
high-purity material  and makes interpretation of the magnitude of the responses
in the presence of S9 mix for the lower-purity materials difficult.
     Margard (1978)  examined PCE  (provided by  Detrex Chemical Industries,
Inc.) in  the Ames/Salmonella  test to determine  whether mutagenic activity of
technical grade  samples is  the  result of added stabilizers.  Precautions  to
prevent escape of material were taken (Margard, personal communication 1981),
but they were not specified in writing.   Tester  strains TA1535, TA1537, TA1538,
TA98, and TA100  were used.   Tests were conducted in the presence and absence
of S9 mix (prepared  from livers of Aroclor-1254 induced rats).  Nonstabilized
and  stabilized materials  were added directly to  the  petri  plates at  0.01,
0.05, and 0.1 ml  per plate rather than vapor exposure.   The nonstabilized test
material was described  as  purified PCE that contained no detectable epoxides
or other stabilizing components, and the stabilized material  was identified as
an industrial  degreasing grade of PCE that  contained 0.07 percent  (by  weight)
epichlorohydrin,  0.007 percent N-methylmorpholine,  0.07 percent beta-hydroxy-
pronitrile,  and  0.01 percent  hydroquinone monoethyl ether (written communica-
tion from L. Schlossberg, Detrex Chemical Industries  Inc., 1981).   Nonstabil-
ized test material was  not detected  as positive in  the  presence or absence of
                                    8-8

-------
                                                                           *
S9 mix.   But  stabilized material  at 0.1 ml per plate (equivalent to 160 mg )
caused a weak response in the absence of S9 mix; twofold increases were observed
for TA1538, TA98,  and TA100.   In the presence  of S9 mix, greater  increases  in
the number  of  revertants for tester strains TA1538 (17-fold  increase), TA98
(10-fold  increase),  and  TA100 (1.7-fold increase) at 0.1 ml  per  plate, were
found.  A  clear  dose-response was  not observed for any  of the tester strains.
The toxicity of  the test material  was  reported as  greater  than  90 percent
killing at  concentrations which caused  increases  in the number  of revertants.
The method used to determine toxicity may be inaccurate, as  discussed previously
for the Shimada and Williams study (1983).   Negative results were reported for
stabilized material in TA1535 and TA1537.  Although the stabilized test material
contained  epichlorohydrin,  which has  been  shown to be  strongly mutagenic in
Salmonella, it is primarily active in base-pair substitution-sensitive strains
(i.e., TA1535  and  TA100; McCann et al.  1975, Anderson et al.  1978), and thus
it does not seem likely that  this  agent can solely account for the activity
observed  for stabilized material  in the frameshift-sensitive strains TA1538
and TA98.   Nevertheless, it appears that the mutagenic  activity is due to the
presence of a contaminant(s).
     Other  investigators have conducted studies  in  which precautions were
taken  to  prevent escape of test material  during testing.  SRI  International
(1983  EPA-sponsored  test)  reported PCE (99+ percent purity,  Aldrich)  to  be
negative  when  tested as a  vapor in sealed desiccators  using  TA1535, TA1537,
TA98,  and TA100 with or  without S9 mix prepared from livers  of female and male
Aroclor-1254 induced rats and mice.  Cells were exposed at 37°C for 8 hours in
desiccators, and then  incubated at 37°C for  an  additional 42  hours.   The
concentrations tested were  0.05, 0.1, 0.5,  1.0, 1.5 ml  added  to a petri plate
at the bottom  of the desiccator chamber in one experiment  and 0.025,  0.05,
0.1,  0.5,  and  1.0 ml added to  a plate  at  the  bottom of the  desiccator in
another experiment.  A toxic  response (reduction or absence of bacterial lawn)
was found at 1.5 ml/per  desiccator.
     The  National  Toxicology Program  (NTP  1983) sponsored Salmonella testing
on PCE (technical  grade, source  Fischer) and obtained negative  results.   Four
standard  tester  strains were used:  TA98,   TA100, TA1535, and TA1537. A pre-
incubation protocol was  followed in which the  cells, S9 activation system, and
chemical are preincubated in  test tubes for 20  minutes  at 37°C

     "Calculation  based  on  the density of tetrachloroethylene  as  1.586 g/ml.

                                    8-9

-------
before addition of the top agar and plating in petri plates.   Two different S9
systems were used;  S9  mix derived from livers  of  Aroclor-1254 induced rat
liver and Aroclor-1254  induced  hamster liver.   Tests were also  conducted  in
the absence of  S9  mix.   Six concentrations (0, 3, 10, 33, 100, 333 ug/plate)
were evaluated. (In TA100,  up  to 10,000 ug/plate was  evaluated.)  Although
incubation was  carried  out in  capped tubes, it  is  possible that evaporation
and some  escape may  have occurred.   However,  it  should  be noted that toxic
levels were tested as indicated by absence or reduction  of  bacterial lawn.
     Bartsch et al. (1979)  investigated the mutagenicity  of PCE  (99.7  percent
purity,  Merck-Darmstadt)  using the  standard  Ames/Salmonella  plate test in
which precautions  to  prevent escape of test material are  not taken.   Negative
results were obtained using tester strain TA100 in the presence of mouse liver
microsomes with and  without cofactors.   The  authors  indicated  that toxic
concentrations  (above 82.9  mg/plate)  were  tested, but did not indicate their
criteria  for determining toxicity.  Interpretation of these negative conclusions
is  limited by  the  amount  of data  presented and because only one  tester strain
was evaluated.
     In  an abstract,  Cerna and Kypenova (1977)  reported  that  in  a spot test
protocol  of the Ames/Salmonella assay,  PCE (purity and source not reported)
induced  both  base-pair  substitution and frameshift mutations  in Salmonella
typhimurium.   The results were  obtained in  the absence of exogenous activation.
These authors  also reported that in a host-mediated  assay using female ICR
mice, PCE induced  significant increases in the number of  revertants in tester
strains TA1950, TA1951,  and TA1952  at dosage levels reported as representing
the LD^g  and one-half of the LDcn-   These  results were reported as not being
dose-dependent.  Because  this report was in abstract form and  did not  provide
details of the  protocol  nor present the data, the  acceptability  of the test
results  is  indeterminate.   Also, the possibility of  mutagenic contaminants
must be considered.
     Beliles  et al. (1980)  used a host-mediated  assay to  evaluate the  effects
of PCE (91.43 percent pure from North Strong Division Chemicals)  in the pres-
ence of  whole  mammal  metabolism.  In  this  study, Salmonella tester  strain
TA98 was  used  as  the  indicator organism, and  male  and female mice (strain
CD-I)  were the  hosts.   The animals were exposed  by  inhalation  7  hours  per  day
for 5 days to  either  100 ppm or  500 ppm.  Bacteria were  injected intraperi-
toneally into the  mice  after the last exposure.   The  bacteria were removed
                                    8-10

-------
from the mice  3 hours later.  At  the  100 ppm dose  level,  increases  in  the
number of  revertants  were  observed for males  (approximately twofold  increase)
but not for  females.  At  500  ppm, a positive response was  reported  for females
(approximately fourfold increase) but  not for males.  Because of the  lack of a
dose-response and the weak responses,  these  findings  should be viewed with
caution.   Also,  the  material used was  of  low-purity,  and the responses may
have been due to contaminants and/or added stabilizers.   In addition, parallel
i_n vitro plate  tests  using Salmonella  were not conducted.   The  parallel  plate
tests are  important  the  determination of the  requirement of whole-mammal
activation.
     Studies on  PCE  have also been conducted  in  Escherichia coli. Henschler
(1977) reported  that  tetrachloroethylene  was  not mutagenic when tested in E^
coli K12 with  metabolic  activation (liver microsomal  fraction  prepared  from
phenobarbital-induced  mice).  This report,  however, is difficult to evaluate
because actual revertant count data (experimental and control number of revert-
ants) and  the  details of  the protocol  are not provided.  It appears that the
conclusions presented in  this report  are actually based on data derived from
the study by Greim et  al.  (1975).
     Greim et al. (1975) reported that negative results were obtained when PCE
(purity reported as  analytical  grade; Merck-Darmstadt) was assayed  in  the
multi-purpose  test  system of Escherichia coli  K12/343/113 ( ). Tests were
conducted  in  the absence  and presence  of  metabolic  activation (phenobarbital-
induced mouse liver microsomal fraction plus NADPH cofactors) at a concentration
of 0.9 mM and a  treatment  time of 2 hours in liquid  suspension at 37°C.  These
treatment conditions  resulted in 99+1 percent survival.   The genetic markers
evaluated  for  mutation induction were  the missense  marker arg   and the frame-
shift marker nad  for  reverse mutation, and gal  and MTR for forward mutation.
Deficiencies  in the  study design  and reporting of  the results reduce the
weight of the negative conclusion.   These deficiencies  are as follows: 1) only
one  concentration  was evaluated,  2)   adequate exposure may not have  been
achieved,  as  indicated by the high survival,  3)  there was no  reporting  of
revertant count  data  (experimental and control), and 4) there was no  reporting
of the number  of replicate  plates used or the number of  the tests conducted.
     The bacterial  tests  discussed above do not clearly  demonstrate  that PCE
itself is  mutagenic.   The  positive responses  found  may be due to contaminants
and/or added stabilizers.  The induced increases in  revertants did not require
                                    8-11

-------
exogenous metabolic activation and were observed in both frameshift  and  base-
pair substitution-sensitive tester strains.   The  positive findings  were not
considered strong in that  large  amounts of material  (estimated vapors at 2.5
percent v/v and at 160  mg/plate)  were needed for the detection of mutagenicity.
Also, there was  a  very narrow range of effective  concentrations because of
toxicity; thus dose-response relationships were not established.   When tested,
highly purifed  samples were not detected as mutagenic under  the  conditions  in
which technical samples caused increases in the number of revertants. Although
some  technical  tetrachloroethylene  samples were positive, there were other
samples  that  were  not  detected as positive.  The  available  results provide
suggestive evidence that certain  technical  samples of PCE are weakly mutagenic
in  Salmonella  and  that the positive responses may be due to impurites and/or
added stabilizers.

8.1.2  Drosophila
     Beliles  et al.  (1980) used  the sex-linked recessive  lethal  assay in
Drosophila melanogaster to test PCE  (91.43 percent purity,  North Strong Division
Chemicals) and reported negative  results.   Adult male flies were exposed for 7
hours  by inhalation  at 100 ppm and  500 ppm.   Treated males  were  mated to
nontreated females at various  times  (2-3-3-2 day mating scheme) to test specific
germ  cell  stages.  No  significant increases (P <  0.05)  over the background
values were observed.   However, only a small sample size was examined. A total
of  3804  chromosomes for the 100 ppm dose and 3956  chromosomes  for the 500 ppm
dose were  evaluated.    This  sample size was only large enough  to exclude the
induction of an approximately  fourfold increase in mutation frequency (Kastenbaum
and  Bowman  1970).   Ideally,  at  least  7000 chromosomes at each  dose level
should be  screened to  preclude a doubling  in  mutation  frequency, which is
generally considered  to be the increase of  biological significance.   Survival
was not  reported in this study, and thus it is uncertain whether a  sufficient
dose was  given.   These  deficiencies  prevent a judgment regarding the mutagenic
activity  of PCE in  Drosophila.
     The  ability of PCE to cause gene mutations in an eucaryotic organism has
not been  adequately  examined.  The  only available study  was  a  sex-linked
recessive lethal test  in  Drosophila  in which PCE was not properly evaluated.
                                    8-12

-------
8.2  CHROMOSOMAL ABERRATION TESTS
8.2.1   Whole-Mammal Bone Marrow Cells
     Rampy et al.  (1978) examined bone marrow cells for chromosome aberrations
from male  and female Sprague-Dawley  rats  after PCE exposure. Animals  were
exposed to  300  ppm  (2.03 mg/1) or  600  ppm (4.07 mg/1) PCE* by inhalation 6
hr/day, 5  days/week,  for one  year.  Three animals  per  dose  were examined.   The
authors reported  "zero"  chromosomal  aberrations per cell  for  both  females  and
males.   The data for females, however, are inadequate for  a clear interpretation
because of  the  very low number of  metaphases  scored  (less than  25  cells per
animal).   In  male  rats,  150  cells  were  scored (50 cells  per animal).  The
negative controls were  also reported as  "zero"  aberrations.   This  observation
is  very  unusual  because, in  general, most laboratories have reported about 1
to  2 percent  total  aberrations for  background  values.   In this  study,  it  is
not  known  whether the highest exposure  level  was near  the maximum  tolerated
dose (MTD)  for  females  because no weight  loss  and  no mortality was observed.
In  males  there  was no weight loss, but significant increases in  mortality
above  control values  were observed  at the highest  dose tested, and therefore
an  MTD may have been approached.   It is  not apparent that the  investigators
determined  the  toxicity of the test  material  to arrive at an MTD  for  this
study,  because  the  dosage levels  used were based on the threshold  limit value
of  100 ppm  for PCE.
     A  rat bone marrow assay was also performed  by Beliles et al.  (1980)  and
reported  as negative.   Ten  males  and ten females  [CRLCOBS  CD(SD)BR]  were
exposed  to an acute dose of  100  ppm and 500 ppm PCE (91.43 percent purity,
North  Strong  Division Chemicals) by  inhalation  for  7 hours.  Bone marrow cells
were harvested  6,  24, and  48 hours later.   No increase in aberrations was
found  for  females,  but for  males, weak clastogenic  effects (breaks, fragments,
deletions,  and  aneuploid cells)  were observed.   At 500  ppm, 3.3  percent cells
with aberrations  versus 0.7  percent  in the  control were found at the 24-hour
kill.  A subchronic study (five exposures, 7  hours  per  day) at 100 ppm  and 500
ppm was  also  conducted.   Animals  were killed 6 hours  after the last exposure.
No  increase in abnormalities  was found in  males,  and  only  a slight  increase at
100 ppm  (1 percent  cells with aberrations) was  found  in females.  This  response
      ^Formulation (liquid volume percent):   trichloroethylene,  3 ppm; hexa-
 chloroethane,   <12 ppm;  carbon  tetrachloride,  2 ppm;  4-methyl  morpholine,  44
 ppm;  nonvolatile residue, 2 ppm; and tetrachloroethylene balance.
                                     8-13

-------
was not  dose-related.   The  female subchronic  control  group had a  very  low
background (0.3 percent cells with aberrations).  Although  isolated  incidences
of  increases  in chromosomal  aberrations  were  observed,  the  lack of dose-
responses precludes an  unequivocal positive  conclusion.   On the other hand, a
negative conclusion cannot  be drawn  because the  authors  did  not discuss the
criteria used  to select the  dosage levels, and thus,  there  is the possibility
that a toxic dose may not have been evaluated.
     Cerna and  Kypenova (1977)  reported in an abstract that  mice  (ICR)  given
an  acute  intraperitoneal  dose one  half  of  the  LD5Q  of PCE or  dosed intraperi-
toneally for  five  applications  in 24-hour intervals  (dose  of one  injection
equalled 1/6  LD™)  did  not  show  cytogenetic  effects in the  bone marrow  cells.
Details  of  the protocol and the cytogenetic data were not available for an
evaluation.  Hence, the negative conclusion of the  authors  cannot  be considered
definitive.

8.2.2  Human Peripheral Lymphocytes
     Ikeda  et  al.   (1980) studied  chromosomal  aberrations,  sister  chromatid
exchanges (SCEs), and variation  in the mitotic  index of peripheral  lymphocytes
cultured  from  ten  workers  (seven males and three  females)  occupationally
exposed  to  technical  grade  PCE  (impurities  not  reported).  The workers  were
divided  into  high  (Group  1) and  low (Group 2) exposure groups.   Group 1
consisted of  six workers  (five males and one female aged  20 to 66 years) from
a  degreasing workshop.  These workers had  a  geometric mean  exposure of  92 ppm
(range 30 to 220 ppm).  The  five males  of  Group 1 had work  histories of  10  to
18  years, whereas the one female had worked  in the  degreasing shop for  only 1
year.  Group 2  included four workers (two  males and two females aged 17  to  31
years) from a  support  department with a shorter work history (3 months  to  3
years) and with an exposure range of 10 to 40 ppm.  The control  group consisted
of  six males  and  five  females.   The authors  did not  indicate  if this was a
matched control, and did  not indicate the medical  histories  of the subjects
(e.g.,  recent  illnesses,  radiation exposures).  There  were no statistically
significant differences (P > 0.05) in the incidences of chromosomal  aberrations
(structural  and numerical)  and  SCEs  between the  exposed  workers and control
group.   The mitotic index  was similar for both exposed and  control groups.
                                    8-14

-------
8.2.3  Drosophila
     In addition to performing the sex-linked recessive lethal assay discussed
earlier, Beliles et  al.  (1980) also conducted a sex chromosome loss assay on
low-purity (91.4 percent, North Strong Division Chemicals) PCE for nondisjunc-
tion in Drosophila  melanogaster.   Males  were exposed  to  100 and 500 ppm of
tetrachloroethylene for 7 hours.  The phenotypic classes used allow for detec-
tion of losses  of  the entire  X  or Y  chromosome  and  the short or  long  arm of
the Y  chromosome.   Although marginal  increases, which were not dose-related,
were observed after PCE treatment, they are not considered sufficient to judge
the data as positive or negative.
     The cytogenetic  tests  discussed  above using mice, rats, Drosophila, and
exposed humans have been reported as negative.  Although these studies are not
considered to be a  thorough evaluation of  the ability  of  PCE  to cause  chromo-
somal  aberrations,  the  data collectively indicates  that  PCE  is not strongly
clastogenic.   However,  there have been no adequate  studies on the ability of
PCE to cause chromosome nondisjunction (aneuploidy).

8.3  OTHER TESTS INDICATIVE OF DNA DAMAGE
8.3.1  DNA Repair
     Unscheduled DNA  synthesis (UDS)  is  measured by repair  of DNA lesions,
which  is indicative of DNA  damage.  Beliles et al.  (1980)  assessed the ability
of PCE (91.43 percent purity,  North Strong Division  Chemicals) to cause UDS  in
human  fibroblast (WI-38) cells.  Because WI-38 cells have  little if any enzyme
activation capability,  the  tests were conducted with  an  exogenous source of
metabolic activation  (i.e., 59 mix).  The  test material was  examined at 0.01,
0.05,  0.1, and  0.5  percent  (v/v) using  conventional liquid  phase exposure.
Scheduled DNA synthesis was blocked by treatment with hydroxyurea, and UDS was
measured by  liquid  scintillation counting  of  incorporated tritiated thymidine
([3H]-TdR) into DNA.   A very  slight  increase was  seen at 0.01 percent (v/v)
PCE  both  in  the presence  (1.5-fold of  control) and absence (1.35-fold of
control) of  Aroclor  rat liver S9 mix.   No  increases occurred at  0.1 and 0.5
percent.  A  toxic  response  was reported at 0.5  percent (i.e., a  decrease in
total  amount of DNA as well as in the incorporation  of [3H]-TdR).  An  increase
in total amount  of  DNA at  0.01 percent was found, suggesting that more cells
are entering the S  phase.   Cells accelerated into S phase would  account for
the increases seen  in [ H]-TdR incorporation at 0.01 percent.  Therefore, it
                                    8-15

-------
is uncertain whether  PCE  induced  a weak UDS response in this  system.   Also,
other problems were  found in this study.   The  positive controls gave  weak
responses.   N-methyl-N1-nitro-N-nitrosoguanidine  elicited  a weak  response
(1.8-fold increase),  but  at a toxic concentration (5 ug/ml) as  indicated by a
decrease in  total  DNA (6.27  ug of DMA versus 23.76 ug of DMA  in  solvent
control).  Benzo[a]pyrene gave a response of 1.16-fold above background.  This
is  an  equivocal  response, particularly because  there  was  also  an  increase
found in the total  amount of DNA.
     Williams and  Shimada (1983,  sponsored  by PPG Industries,  Inc.)  evaluated
two  different  samples of  PCE,  Perchlor  200 (low-stabilized,  99.93  percent
purity)  and Perchlor 230  (stabilized,  purity 99.80 percent), for their  ability
to  cause UDS  in  the hepatocyte primary culture  (HPC)/DNA  repair test.  The
target  cells  in  this  test system  have  a capability to metabolize xenobiotics.
Williams and  Shimada measured UDS by  autoradiographic  determination of the
amount  of  [ H]-TdR (10 uCi/ml) incorporated into nuclear  DNA.   Hepatocytes
were  isolated  from adult male Fischer 344  rats.  Cells were  treated for 18
hours or 3 hours.  For the 3-hour exposure,  cultures  were incubated another 15
hours  in the  absence of  PCE to allow for DNA repair  synthesis. The criterion
that was used  for  a positive result was a  net  nuclear grain  count  of  5 in
triplicate  coverslips.*   Negative  results were reported for both Perchlor  230
(at  concentrations  of 0.001,  0.01, 0.1, and 1.0  percent v/v) and Perchlor  200
(at  concentrations  of 0.0001,  0.001,  0.1,   and 1.0 percent) when tests were
conducted using  conventional  liquid-phase  exposure  in which PCE was added to
the culture medium.
     For both  materials,  positive responses were reported  when testing was
performed using  vapor-phase exposure  in  gas tight  chambers.   Testing was
conducted at 0.1,  1.0,  and 2.5 percent v/v (desired  air concentration) for 3
and 18  hours.  Perchlor 230 at 0.1 percent  caused an  increase  in UDS when the
cells were exposed for 3  hours (6.2 +4.9 net nuclear grain count,  50  percent
cells showing toxic  effects)  and  for  18 hours (15.9 +1.6  net  nuclear  grain
     *Nuclear grain counts were  reported as the mean  + standard deviation.
Cytoplasmic grain counts  in  three nuclear size areas adjacent to the nucleus
were determined.  The highest cytoplasmic grain count was  subtracted  from the
nuclear count.   This value is referred to as "net"  grain count.
                                    8-16

-------
count, 25 percent cells showing toxic effects).  Perchlor 200 caused an increase
in UDS at  0.1 percent v/v for  the  3-hour treatment (10.8 + 6.1 percent net
nuclear grain  count,  75  percent cells showing toxic effects) but not for the
18-hour treatment.  At 1.0 percent  and 2.5  percent  for  both  materials at both
treatment times,  nearly 100  percent of cells  showed toxic effects.   It  should
be pointed out that the background  control  values were  taken from the conven-
tional exposure experiment.   This makes intrepretation of the results difficult,
particularly  because  the  conventional-phase media was different  from the gas-
phase media.   Because these  data  are based  on one test  in which  there were  no
concurrent positive  and  negative controls, the  results  are  considered only
suggestive of  a positive  effect.   To validate these findings it is  necessary
to repeat experiments using appropriate concurrent controls and a concentration
range to demonstrate  a dose-response.  Although the possibility exists that an
impurity(ies) may be  responsible for the observed effects, the authors did not
examine high-purity  Perchlor (99.98+ percent) as they  did in the Salmonella
tests discussed earlier.   If high-purity material  had  tested negative  under
the same experimental conditions under which  the lower-purity materials tested
positive, as  in the  Salmonella test, a  stronger  argument could  be made that
impurities were causing the effects.
     Williams  (1983)  conducted HPC/DNA  repair  tests  using  hepatocytes  from
male  BgC3F,  mice and male Osborne-Mendel  rats  on PCE (99+  percent  purity,
Aldrich Chemical Company).  Conventional  liquid-phase exposure conditions were
used.  Negative results  were reported for  both rat and mice hepatocytes when
PCE was  added to the culture medium at 0.00001, 0.0001, 0.001,  0.01, and 0.1
percent (v/v)  for 18 hours.   The material  was  reported as  "toxic"  at  0.01
percent and  higher.   The  highest cytoplasmic grain count was subtracted from
the nuclear  count.   This  reduces the possibility of false positives, but the
chance of  missing a  weak UDS  inducer would be increased, especially if the
cytoplasmic  grain count  is  high.    This  consideration  also  applies  to the
Williams and  Shimada  conventional-phase  exposure study discussed previously.
The cytoplasmic grain counts were  not reported  in  any  of the HPC/DNA  repair
studies.
     Toxicity  identified  by the absence of S phase cells and general  cellular
morphology.  This  is not  an  accurate method  of determining toxicity.
                                    8-17

-------
8.3.2  MJtotic Recombination
     Callen et al.  (1980)  evaluated the ability of PCE (purity not reported,
stabilized 0.01 percent thymol, Eastman Kodak) to cause mitotic gene conversion
(nonreciprocal recombination) at the trp-5 locus, mitotic crossing-over  (reci-
procal recombination) at the ade-2 locus, and gene mutation  (reversion)  at  the
ilv-1  locus  in  log phase cultures of Saccharomyces cerevisiae D7. Cells were
incubated  for  one  hour in culture medium  containing  0, 4.9, 6.6,  and 8.2 mM
PCE.   No  exogenous source of metabolic activation was used  in these studies.
At  4.9 mM (84 percent survival),  no  significant  increases  in the  frequency of
gene  conversion  (1.9  convertants/10   survivors  versus 1.4  convertants/10
survivors  in background control) and mitotic crossing-over (5.3 mitotic  recom-
binants/10   survivors  versus 3.3 mitotic recombinants/10  survivors in  back-
ground control)  occurred.   As shown in Figure 8-2, when the concentration  of
PCE was  increased  to 6.6 mM  (58 percent survival), increases in mitotic gene
conversion  and  mitotic crossing-over did  occur  (8.3 convertants/10  survivors
and 52.6  mitotic recombinants/10  survivors, respectively).  Mitotic recombina-
tional  activity  was not determined at 8.2 mM  because less  than 0.1 percent
survival  was  found.   No significant  increases  in gene mutations were  observed
at  4.9 mM  (3.8 revertants/10  survivors versus 2.9 revertants/10   survivors in
control).   The  reverse mutation  frequency was  not  determined at 6.6  mM.
Therefore,  the  induced number of revertants is too low to be indicative of a
positive  result, but the high values for the recombinational events do indicate
a  positive  effect  at 4.9 mM  test material. The  possibility  that  the  effects
were  caused  by  a mutagenic  impurity(ies) should be considered.  It should  be
pointed out  that  in the study by Callen  et  al.  the  Ade+  recombinants  were
estimated  from a  total  of 30 plates, ten  of  which contained minimum medium
(plus  adenine  and  isoleucine)  used for estimating  the number of trp-5  conver-
tants.  Because mitotic crossing-over and  gene conversion  are  not  necessarily
distinct events,  in that they probably depend on a common inducible mechanism,
the number of Ade  recombinants may have been overestimated  by including those
Ade  that were already  Trp  .  There  is experimental support  that  these  recom-
binational events  are  inducible  by  a common factor(s)  (see  Fabre Fabre and
Roman  1977, Fabre  1978).   The positive findings of Callen and coworkers for
mitotic recombination  should  be  confirmed  by repeating the assay  using  appro-
priate selection  conditions.   In addition, because the responses were observed
within a narrow "window,"  at least one additional concentration between  the low
                                    8-18

-------
              50
           CO
           cc
           O
          en
          D
          to

          o

          to
          CD

          O
          O
          LU
          CC

          O

          o
              40
                                   4         6

                             TETRACHLOROETHYLENE (mM)
10
Figure 8-2.  Induction of mitotic recombination by tetrachloroethylene  in
             Saccharomyces cerevisiae D7.  The frequency of mitotic  crossing-
             over (•   •) and gene conversion (A—-A) was determined.   Log-
             phase yeast cells were treated with test chemical for one  hour
             without the addition of an exogenous metabolic activation  system.
             (Adapted from Callen et al. 1980)
                                      8'19

-------
least one additional concentration between the low and high doses used  should
be tested to demonstrate  a clear dose-response.
     Bronzetti  et al. (1983) also used the yeast S^ cerevisiae D7 to  evaluate
the effects of PCE  (99.5  percent pure, stabilized with 0.01 percent  thymol,
Carlo Erba  Co.) at the trp-5, ade-2, and ilv-1 loci.  In this study,  however,
negative results were obtained after a 2-hour treatment at 5, 10, 20, 60,  and
85 mM in  the  presence  or absence of S9 mix (Aroclor-1254 induced rat liver).
Differences in the  experimental  protocol  from that used by Call en et al. may
explain these negative  findings.   Bronzetti et al.  used cells in the stationary
phase of growth rather  than in the log phase of growth.   For some chemicals it
has  been  observed that stationary cells are more refractory than  log  cells to
mutagenic treatment  (Mayer and Coin 1980,  Shahin 1975). Bronzetti et  al. were
able to  test  higher  concentrations than Callen and coworkers.  In the absence
of S9 mix,  Callen  et al.  reported less than 0.1 percent survival at  8.2 mM,
while Bronzetti et  al.  did not observe complete killing until 85 mM.  Another
difference  between  these  two studies is the  source  of  tetrachloroethylene.
Bronzetti et  al. purchased  PCE  from Carlo Erba Co.  (Milan, Italy) and Callen
et al.  obtained the test  agent from Eastman Kodak Co.  (Rochester, N.Y.); thus,
it is possible that  the test samples may have contained different impurities.
     Bronzetti et al.  (1983)  also obtained negative results in  an  intrasan-
guineous host-mediated  assay using S. cerevisiae D7 as the indicator  organism
                                                           o
and  CD-I mice  as the host.   Stationary yeast cells (4 x 10  cells) were in-
jected into the retro-orbital sinus of mice.   After injection of the yeast, an
acute oral  dose of 11 g PCE/kg body weight (b.w.) was given.  A  subacute dose
of 2 g/kg b.w. given 5 days a week for a total of 12 administrations  was also
used (the last test  dose  was 4 g/kg b.w.;  therefore,  the total  dose was 26
g/kg b.w.).   In the  subacute study,  yeast cells  were injected after the last
dose of PCE (i.e.,  4 g/kg b.w.).   Four hours after injection of yeast, in both
the  acute  and subchronic study,  the  cells were recovered  from  the liver,
lungs,  and  kidneys of three animals.  No concurrent positive control  chemical
was used in these studies to ensure that the system was functioning properly.
     The positive  responses  discussed above for mitotic recombination in yeast
and DNA repair synthesis  in  mammalian cells provide suggestive evidence that
certain  technical  samples  of PCE may be active in damaging DNA.  However, toxic
concentrations of material were  needed to elicit these responses. The possibi-
lity that impurities and/or  added stabilizers caused the  increased effects
should  be considered.
                                    8-20

-------
8.4  DNA BINDING STUDIES
     Chemical adduct  formation  is a critical  step  in  certain  types  of  muta-
genesis. Schumann  et  al.  (1980)  reported  that  there was  no  detectable bindinq
   14
of   C-labeled PCE (99 percent purity) to DNA when  mice were treated by inhala-
tion at 600 ppm for 6 hours or when they were given an acute dose of 500 mg/kg
orally.  The specific activity of the   C-label (26,593 dpm/umol in the inhala-
tion study  and  133,273  dpm/umol   in the oral  study) is too low, however, to
preclude the possibility  of very low  levels of DNA binding.  For example, at
the specific activities used  under  the experimental conditions,  at an assumed
                    -5
binding  level of  10   alkylations per  nucleotide  (Stott  and  Watanabe 1982)
there  would  be  5-10 dpm per sample.   This  is  at  the limit of practical  detec-
tion.   Therefore, the possibility of binding at slightly  less than 10   alkyla-
tions  per  nucleotide  cannot ruled out.  These  negative findings,  however, are
consistent with  the  negative and weak  results  reported  in the mutagenicity
tests  discussed above.
8.5  STUDIES INDICATIVE OF MUTAGENICITY IN GERM CELLS
     An important aspect of a mutagenicity evaluation is to assess the potential
of the  chemical  to reach the germinal  tissue of humans and cause mutations
that may  contribute to the  genetic disease  burden.   This  assessment  is almost
always  based  on  animal experimentation.  The ability of PCE to cause genetic
damage  in  germinal  tissue has not  been well  studied.   The  only test results
available were from a  dominant lethal study in rats  [CRL:  COBS CD(SD)BR] and a
sperm   morphology   assay   in   both  rats   and  mice   (strain   CD-I).
Tetrachloroethylene  (91.43  percent purity,  North Strong Division  Chemicals)
did  not cause  an increase in dominant  lethals in rats when unexposed females
were mated  during  a 7-week period  to exposed adult  males given an acute dose
of 100  ppm and 500 ppm by inhalation for 7 hours per day for  5 days  (Bellies
et al.   1980).  The  dominant lethal  assay is generally thought  to measure gross
chromosome  damage (Bateman and Epstein  1971).  Also, this test is  not considered
a  sensitive assay  because  of the  high spontaneous  level  of  lethal  events
(Russell  and Matter  1980),  and thus,  negative  results do not necessarily
indicate that the chemical does  not reach and damage the germ  cell DNA.
     Beliles et  al.  (1980) also examined PCE for altered sperm morphology  in
treated rats  and mice.  After dosing at  100 ppm and 500  ppm by inhalation 7
hours per  day  for 5  consecutive days,  groups of four animals were killed at
                                     8-21

-------
the end of  1  week,  4 weeks, and 10 weeks to examine effects on various  germ
cell  stages.  Sperm was collected from the cauda epididymis, and  at  least  500
cells were examined.   Negative results were reported in rats; the mice, however,
showed positive  responses.   At 500  ppm,  19.7 percent  abnormal  sperm were
observed (versus 6.0 percent in the negative controls)  during  the fourth week
after  exposure  (corresponding  to the  spermatocyte stage).   By themselves,
these  positive findings alone  are  not sufficient to conclude that PCE alters
germ cell  DNA because  this  assay is only an indicator of chemical effects on
sperm  and does not provide definitive evidence that a chemical  reached germinal
tissue and  damaged DNA.  Therefore, because limited information was  provided,
it is  not clear whether PCE (or impurities) reaches the germ tissue.   However,
if PCE reached germ cell DNA, there may be no serious risk of mutation because
of the largely  negative or marginal results found in mutagenicity tests dis_
cussed previously.

8.6  MUTAGENICITY OF METABOLITES
     Trichloroacetic acid  (TCA) is  a known human metabolite  of  PCE.   The
formation  of  TCA is  thought to occur through  the  formation of an epoxide,
tetrachloroethylene oxide, and its  subsequent rearrangement to trichloroacetyl
chloride or trichloroacetaldehyde,  which then rapidly hydrolyzes to TCA.   (See
chapter 5  on  metabolism for  a more  detailed discussion.)  These intermediates
are considered relevant in assessing the mutagenicity of PCE.
     Tetrachloroethylene oxide,  which is considered to  be  the biologically
active intermediate  of  the  parent  compound PCE,  was assayed for mutagenicity
in the absence of  exogenous metabolic activation  in several bacterial tests
(Kline et  al.  1982).   Tetrachloroethylene oxide increased  the  number  of re-
vertants in a dose-dependent manner in Salmonella tester strain  TA1535  when
assayed by  a  preincubation  liquid  protocol  (20 minutes at 37°C),  but did not
cause an increase in revertants using Escherichia coll  WP2 uvr A.   In Salmonella
a 14-fold  increase occurred  at 2.5 mM and a 20-fold increase occurred  at 5mM.
Tetrachloroethylene epoxide was toxic at 25mM.
     This  epoxide was also evaluated by Kline et al.  (1982) in the E^ coli pol
A assay.   Positive effects  were observed at 0.04,  0.09,  and 0.44 mM/ml as
measured by differential  growth inhibition  of a  DNA  polymerase-deficient
strain in  comparison with  its polymerase-proficient parent.
                                    8-22

-------
     Waskell (1978)  tested  TCA at 0.45 tug/plate in Salmonella TA98 and TA100
and obtained negative  results.   There are other intermediates which have not
been identified  in  humans but are thought  to  occur (e.g., trichloroethanol
chloral  hydrate).   Waskell  (1978) obtained  negative  results  for trichloro-
ethanol  in  Salmonella  TA98  and TA100  up  to  a dose  of  7.5 mg/plate.  Gu et al.
(1981),  however,  reported  suggestive  evidence  that trichloroethanol  weakly
induced sister chromatid exchange (SCE) formation  in primary cultures of human
lymphocytes.
     Chloral hydrate was reported to be marginally mutagenic (less than twofold
increase over  a  dose range of 0.5 to  10 mg per plate)  in Salmonella TA100
(Waskell 1978).   Gu et  al. (1981) also  provided  suggestive  evidence than
chloral  hydrate  at  54.1  mg/1  caused a  weak  increase in  SCEs in cultured  human
lymphocytes. Chloral hydrate  has also been  shown  to block spindle elongation
in  insect  spermatocytes  (Ris  1949).   Data on metabolites of PCE suggest that
if  the  parent  compound was biotransformed,  its metabolites may be genotoxic;
these data are limited, however, and additional studies are needed on metabolism
and on the mutagenicity of  metabolites to reach a  clear conclusion.

8.7  SUMMARY AND CONCLUSIONS
     Tetrachloroethylene  itself  has  not  been clearly shown to be a mutagen.
Certain commercial and technical preparations have elicited positive responses
in  the Ames bacterial  test,  a yeast  recombinogenic assay, a host-mediated
assay using Salmonella, and DNA  repair assays.   In general, the responses were
weak,  and  eliciting them required rather high toxic concentrations of tetra-
chloroethylene.   No dose-response relationships  were established in these
studies.   The  positive findings  may be explained  by the presence  of mutagenic
contaminants and/or added stabilizers.  Highly purified tetrachloroethylene
has only been  evaluated in the  Ames/Salmonella test, where negative  results
were obtained.
     Several other tests of commercial and technical samples have been reported
to  be negative.  In addition, The National Toxicology Program (NTP) has recently
sponsored  mutagenicity testing (a modified  Ames Salmonella test,  a  sex-linked
recessive  lethal  test  in Drosophila,  sister chromatid  exchange formation and
chromosome  aberrations in Chinese hamster ovary cells j_n vitro  on a technical
sample  of  PCE; the  preliminary  results were negative.   (These studies are  in
the process  of being peer-reviewed,  and  were  not discussed in  this chapter,
                                    8-23

-------
except for the Ames  test.)   The inconsistencies of available results on dif-
ferent samples of PCE may be a function of the toxicity of the test  material,
of exposure conditions used for testing this volatile chemical, or of differ-
ences in sample contaminants and/or added stabilizers.   Information on chemical
composition of the PCE test samples was scarce.
     Although PCE itself  has  not  been shown to  be  mutagenic,  it should be
emphasized that the  negative  results  are not wholly unequivocal.  Appropriate
concurrent controls,  adequate  sample  sizes,  and exposure  conditions were
sometimes not used,  and in some cases  the available data are not sufficient to
determine whether an  adequate  test was conducted.   Also, there have been  no
reliable studies  investigating the ability of PCE to cause chromosome nondis-
junction, which would  result  in aneuploidy,  a significant genotoxic effect.
     Because the epoxide  of PCE was mutagenic in bacterial  studies,  the concern
should be  raised  that  it  may pose a  mutagenic  hazard.   It should be noted,
however, that the parent  compound  was  assayed in the presence of several types
of metabolic  activation systems (i.e., liver homogenates, intact  hepatocytes,
and whole  mammals)  and the  results were  largely negative or weakly positive.
Therefore,  it  is  uncertain  whether these negative or weak findings were the
result of limitations of  the activation systems,  the epoxide not being produced
in sufficient quantities, or  the  epoxide possessing too short a half life to
cause a detectable mutagenic response.
     In  conclusion,  inadequate information exists to warrant  a provisional
classification of PCE either as nonmutagenic  or  mutagenic.   If PCE is a mutagen,
the evidence available thus  far indicates that it is only weakly so.   (Because
of insufficient information, this  conclusion is not made with regard to its
potential  for  causing  chromosome  nondisjunction.)   Certain commercial and
technical preparations of PCE  may  contain mutagenic impurities and/or  added
stabilizers.   Although there may  be  mutagenic agents in certain preparations
of PCE,  usually  large  amounts  of  material (at toxic levels) were required to
elicit weak responses.
                                    8-24

-------
8.8  REFERENCES

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.  41:
     249-277.

Bateman, A.J.  and S.S.  Epstein.   1971.   Dominant lethal  mutations in mammals.
     In: Chemical mutagens:  Principles and methods  for their  detection, Vol.  2
     (A. Hollaender, ed.)  Plenum Press, New York, pp.  541-568.

Beliles, R.P-,  D.J.  Brusisk, and  F.J.  Mecler.   1980.   Teratogenic mutagenic
     risk of workplace contaminants:  trichlorethylene, perchloroethylene,  and
     carbon  disulfide.   Contract  No.  210-77-0047.   Litton bionetics,  Inc.,
     Kensington, Maryland.

Bronzetti,  G.. ,  C.  Bauer, C. Corsi, R.  Del Carratore,  A.  Galli,  R.  Nieri,  and
     M. Paolini.   1983.   Genetic  and  biochemical studies  on perchloroethylene
     i_n vitro and i_n vivo.   Mutat.  Res. 116:323-331.

Callen, D.F., C.R. Wolf, R.M.  Philpot.  1980.   Cytochrome P450  mediated genetic
     activity  and  cytotoxicity of seven halogenated aliphatic  hydrocarbons in
     Saccharomyces cerevisiae. Mutat. Res. 77:55-63.

Cerna,  N.,  and  H.  Kypenova.  1977.  Mutagenic  activity  of chloroethylenes
     analyzed by screening  system  test.  Mutat.  Res. 46:36  (Abst.).

Fabre,  F.   1978.   Induced  intragenic recombination  in yeast can occur during
     the G-L mitotic phase.   Nature 272:795-798.

Fabre,  F.,  and  H. Roman.   1977.  Genetic evidence for  inducibility  of recombina-
     tion  competence  in yeast.    Proc.  Natl.  Acad.  Sci.  USA  74:1667-1671.

Greim,  H. ,  G. Bonse, Z.  Radwan, D.  Reichert,  and D.  Henschler.   1975,  Mutagen-
     icity  in vitro  and potential  carcinogenicity  of chlorinated ethylenes as
     a  function of metabolic  oxirane  formation.   Biochem. Pharmacol. 24:
     2013-2017.

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.   lexicological  European Research
     3:63-67.

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

Ikeda,  M. ,  A. Koizumi, T.  Watanable,  A. Endo, and K. Sato.   1980.   Cytogenetic
     and cytokinetic investigations on  lymphocytes  from workers occupationally
     exposed  to  tetrachloroethylene.  Toxicol.  Letters.   5:251-256.

Kastenbaum,  M.A.  and K.O.   Bowman.  1970.  Tables for  determining statistical
     significance of mutation  frequencies.   Mutat.  Res.  9:527-549.
                                     8-25

-------
Kline, S.A.,  E.G.  McCoy, H.S. Rosenkranz, and B.L. Van Duuren.   1982.   Mutageni
     city  of  chloralkene epoxides  in  bacterial systems.  Mutat.  Res.  101:
     115-125.

Margard,  W.   1978.   In vitro bioassay  of chlorinated hydrocarbon  solvents.
     Battene  Laboratories.   Unpublished proprietary document for  Detrex
     Chemical  Industries.

Mayer,  V.W.,  and C.T.  Coin.  1980.   Induction of mitotic  recombination by
     certain  hair-dye chemicals  in  Saccharomyces cerevisiae.   Mutat.  Res.
     78:243-252.

McCann,  J., E.  Choi,  E. Yamasaki, and B.N. Ames.  1975.  Detection of  carci-
     nogens as  mutagens in  the  Salmonella/microsome  test:   Assay of 300 chem-
     icals.  Proc. Natl. Acad. Sci. USA 72:5125-5139.

NTP.  1983.  Unpublished data on  tetrachloroethylene:  on Salmonella/microsome
     preincubation test provided  by Dr. E. Zeiger.

Rampy,  L.W.,  J.F.  Quast, M.F. Balmer, B.K.J. Leong, and P.J. Gehring.   1978.
     Results  of a long-term inhalation toxicity study.  Perchloroethylene  in
     rats. Toxicology Research Laboratory.  Health and Environmental  Research.
     The  Dow Chemical Company.  Midland,  Michigan.  Unpublished.

Ris,  H.  1949.   The anaphase movement  of  chromosomes  in  the spermatocytes of
     the  grasshopper.   Biol. Bull. 96:90-106.

Russell  L.B.,  and B.E.  Matter.   1980.   Whole  mammal mutagenicity tests.
     Evaluation of five methods.  Mutat.  Res. 75:279-302.

Schumann,  A.M.,  J.F.  Quast,  and  P.G. Watanabe.   1980.   The pharmacokinetics
     and  macromolecular interactions  of perchloroethylene  in mice and rats  as
     related to oncogenicity.  Toxicol. Appl. Pharm.  55:207-219.

Shahin,  M.M.   1975.   Genetic activity of  niridazole  in  yeast.   Mutat.  Res.
     30:191-198.

SRI  International.   1983.    Salmonella  test  results  on tetrachloroethylene.
     Prepared  for U.S  Environmental  Protection Agency,  Dr.  Harry  Milman,
     project officer.   Unpublished.

Stott,  W.T.  and  P.G.  Watanabe.   1982.   Differentiation of genetic  versus
     epigenetic mechanisms of toxicity and its  application to risk  assessment.
     Drug Metabolism Reviews 13:853-873.

Schlossberg,  L.  (Detrex Chemical Industries, Inc. Detroit, Michigan)   January
     5, 1981.   Memorandum to Dr. V. Vaughan-Dellarco of the U.S.  Environmental
     Protection Agency, Reproductive Effects Assessment Group.

Waskell,  L.   1978.   A  study of the mutagenicity  of  anesthetics and their
     metabolites.   Mutat. Res. 57:141-153.

Williams, G.M.   1983.   DMA repair tests of 11 chlorinated hydrocarbon analogs.
     Prepared  for  ICAIR Life systems, Inc. TR-507-18  and  U.S.  Environmental
     Protection Agency.   Dr. Harry  Milman,  project  officer.   Unpublished.

                                    8-26

-------
Williams, G.M.,  and T.  Shimada.  January  1983.  Evaluation of several halo-
     genated ethane and ethylene compounds for genotoxicity.  Final report for
     PPG Industries, Inc., Pittsburgh, Pennsylvania.  Unpublished.
                                     8-27

-------
                              9.  CARCINOGENICITY







     The purpose of this section is to provide an evaluation of the likelihood



that tetrachloroethylene (perchloroethylene, PCE) is a human carcinogen and,



on the assumption that it is a human carcinogen, to provide a basis for esti-



mating its public health impact, including a potency evaluation, in relation



to other carcinogens.  The evaluation of carcinogenicity depends heavily on



animal bioassays and epidemiologic evidence.  However, information  on  mutage-



nicity and metabolism, particularly in relation to interaction  with DNA, as



well as to pharmacokinetic behavior, has an important bearing on both  the



qualitative and quantitative assessment of carcinogenicity.   The available



information on these subjects is reviewed in other sections  of  this document.



This section presents an evaluation of the animal bioassays,  the human  epide-



miologic evidence, the quantitative aspects of assessment,  and  finally,  a



summary and conclusions dealing with all of the relevant aspects of the  carci-



nogenicity of PCE.





9.1  ANIMAL STUDIES



     Two long-term animal bioassays have been performed to  assess the  carcino-



genic potential of PCE.  In one study involving exposure of  rats and mice to



PCE by gavage, the National Cancer Institute (NCI) (1977a)  reported the induc-



tion of hepatocellular carcinomas in male and female mice,  but  determined that



the test with rats was inconclusive because of excessive mortality.  In the



other study, in which Sprague-Dawley rats were exposed to PCE by inhalation,



the Dow Chemical Company (Rampy et al., 1978) reported no evidence  for the



carcinogenicity of the chemical.  However, limitations in this  study make it



difficult to assess the carcinogenic potential of PCE.
                                      9-1

-------
9.1.1  National Cancer Institute Bioassay (1977a)



     The PCE sample used in this bioassay was purchased from the Aldrich Chem-



i-cal Company, Milwaukee, Wisconsin.  Analysis by gas-liquid chromatography and



infrared spectroscopy yielded results indicating a purity of > 99% with at



least one minor impurity not identified in the report.  Identification of the



impurities in the test sample was not made (personal  communications with the



NCI and the Aldrich Chemical Company).



     The carcinogenicity of PCE was tested in Osborne-Mendel  rats and B6C3F1



mice.  The initial age of the weanling animals was 25 days for the mice and 35



days for the rats.  Two treatment groups consisted of 50 males and 50 females,



and matched vehicle (corn oil)  and untreated control  groups comprised 20 ani-



mals of each sex.  Selected dosage levels were those  determined to be maximally



tolerated in an 8-week subchronic study, i.e., a dosage that  was not fatal



and/or did not reduce body weight gain more than approximately 20%, and one-



half maximally tolerated in an  8-week subchronic toxicity test.  Time-weighted



average doses (mg/kg/dose) in the chronic study were  941 and  471 for male rats,



949 and 474 for female rats, 1,072 and 536 for male mice, and 772 and 386 for



female mice.  PCE was administered to the animals by  gastric  intubation in  corn



oil once each day, 5 days/week, for 78 weeks.  During the final 26 weeks of



treatment, doses were administered to rats in a cyclic pattern of 1 week with-



out treatment followed by 4 weeks with treatment.  Body weights and food con-



sumption were obtained weekly for the first 10 weeks  and monthly thereafter.



Mice and rats were permitted to survive an additional  12 and  32 weeks after



treatment, respectively, until  sacrifice.



     Each animal  was submitted  to extensive gross and microscopic examinations.



Specified organs, plus any other tissue containing visible lesions, were fixed



in 10% buffered formalin, embedded in paraplast, and  sectioned for slides.






                                      9-2

-------
Hematoxylin and eosin staining (H and E) was used routinely, but other stains



were employed when needed.  Diagnoses of observed tumors and other lesions were



coded according to a modified Systematized Nomenclature of Pathology (SNOP)



originally developed by the College of American Pathologists in 1965.



     PCE was found to be carcinogenic in mice in this study.  Results  summa-



rized in Table 9-1 indicate that PCE induced highly statistically significant



(p < 0.001) increases in the incidence of hepatocel1ular carcinomas in both



sexes of mice in both treatment groups as compared to untreated controls  or



vehicle-controls.  The microscopic appearance of carcinomas was variable, with



some tumors composed of we!1-differentiated hepatocytes arranged in rather



uniform hepatic cordSj and other lesions consisting of anaplastic cells,  often



with inclusion bodies with vacuolated, pale cytoplasm.   Mitotic figures were



often present.  In male mice, the first  hepatocel1ular carcinomas were detected



at 27 weeks in the low-dose group, 40 weeks in the high-dose group, and 90 and



91 weeks in vehicle-control and untreated control  groups.   In female mice, the



first hepatocellular carcinomas were observed at week 41 in the low-dose  group,



week 50 in the high-dose group, and week 91 in the untreated control group.



Metastases of hepatocellular carcinomas  occurred in  the kidneys of one un-



treated control  male and in the lung of  three low-dose males, one low-dose



female, and one high-dose female.



     Toxic nephropathy in mice was apparent in 40/49  low-dose males, 45/48



high-dose males, 46/48 low-dose females, and 48/48 high-dose females.   Control



animals did not  exhibit this lesion.  Chronic murine  pneumonia was also a fre-



quently observed finding.  A low incidence of bloating  or  abdominal  distension



was noted in treated animals during the  second year  of  the study.  Body weight



gain  was comparable between groups (Figure 9-1).  Median survival  times were



greater than 90  weeks in control  males,  78 weeks in  low-dose males, and 43





                                      9-3

-------
               TABLE 9-1.   INCIDENCE OF HEPATOCELLULAR CARCINOMAS
                              IN B6C3F1 MICE FED PCE
   Dose  (mg/kg/day)a
Hepatocellular carcinomas            P values
         Males

       untreated                       2/17 (12%)

   vehicle-control                    2/20 (10%)

          536                         32/49 (65%)                  P < 0-001

         1072                         27/48 (56%)                  P < 0.001

        Females

       untreated                       2/20 (10%)

    vehicle-control                   0/20 (0%)

          386                         19/48 (40%)                  P < 0.001

          772                         19/48 (40%)                  P < 0.001


aTime-weighted average doses.

bProbability level (p-values) for the Fisher Exact Test comparison of dose
 groups with vehicle-control group.

SOURCE:  National Cancer Institute, 1977a.
                                     9-4

-------
  50 •
cc
CD

I
CD
30 —
>
Q
O
m
  10 —
       MALE MICE
             15
                      |
                      30
                                                        UNTREATED CONTROL

                                                        VEHICLE CONTROL
                                            • ••••ee*> LOW DOSE

                                            .„,„___ HIGH DOSE
                             |
                            45
 I
60
 |
75
 I
90
                                                                          -40
                                                                         30
                                                                        -20
                                                                          - 10
                                                              105
                                  120
                               TIME ON TEST (WEEKS)
ou —
_ 40-
co
cr
0
~ 30-
X
CD
LU
§
> 20-
O
m
z
LU
1 10-


o —
c





/'X'^«.'**- ^•~"~~'
•^ •"•"••"' • "**
• * *^i^-*'^'M>''*""J™M~^ "" ^ r""*"*" ^^^^- ' -^* JJ tj
««» *^^3^^^^^^"^^ \. .^^ *^(j • • •
r4!&
jj£f


•-»•—• — • VEHICLE CONTROL

«•••«•»« LOW DOSE
FEMALE MICE "" ' 	 " HIGH DOSE
i 1 i 1 i I i 1 i 1 i 1 i | i
3 15 30 45 60 75 90 105 12
— bU
— 40

~

— 30




- 20



- 10


0
                              TIME ON TEST (WEEKS)




Figure  9-1.   Growth curves for male  and female mice  in the PCE  chronic study


SOURCE:   National Cancer Institute,  1977a.
                                      9-5

-------
weeks In high-dose males (Figure 9-2).   Median  survival  times  for females were



greater than 90 weeks in control  females,  62 weeks  in  low-dose females,  and 50



weeks in high-dose females (Figure 9-2).



     In rats, toxic nephropathy.  not  found in control  animals, was detected in



43/49 low-dose males, 47/50 high-dose males, 29/50  low-dose  females,  and 39/50



high-dose females.  Figure 9-3 indicates  that treated  rats gained less weight



than controls, though the difference  was  slight,  with  maximum  reduction  being



13% during the first year and 19% during  the second year.  Clinical signs



apparent in treated animals included  a hunched  appearance and  urine stains



on the lower abdomen.  Respiratory abnormalities, characterized by dyspnea,



wheezing, and/or reddish nasal  discharge,  were  noted with increased incidence



in all  groups during aging of the animals, and  chronic murine  pneumonia  was



diagnosed in _> 62% of the animals in  each  group.  As indicated in Figure 9-4,



median survival times were greater than 88 weeks  in the  control  groups,  68



weeks in low-dose females, 66 weeks in high-dose  females, 67 weeks in  low-dose



males,  and 44 weeks in high-dose  males.  The survival  was not  adequate to



support any conclusions about the carcinogenicity of PCE in  rats.



     In an attempt to characterize impurities present  in the PCE  product used



in this study, documented chemical  analyses of  the  test  samples performed at



the carcinogenicity testing laboratory of  the NCI bioassay program were  exam-



ined by the analytical department of  the  Diamond  Shamrock Corporation  (inter-



office  memorandum from E. A.  Rowe to  G. K. Hatfield, Diamond Shamrock  Cor-



poration, October 8, 1979, obtained with  the documented  chemical  analyses



from G. K.  Hatfield, January 28, 1981).   As indicated in the  interoffice



memorandum, PCE samples used  in the National  Cancer Institute  bioassay  (NCI,



1977a)  were not available for analysis  at  Diamond Shamrock;  however,  the



analytical method, with the same  type of  instrument and  column used at the






                                      9-6

-------



: SURVIVAL
0 0
en co
i , i
i_l_
O
t—
5 0.4-
CO
51
0.2 —


.u
c


VLJ^ i
!*". !— U_
j •••• 	 ••;!—.
\ \ ''C\_
\ — i •. l— j_
L, "~'**»7j— .— .
L • •
™i •
— — — UNTREATED CONTROL j_
	 .. VEHICLE CONTROL "I—.—
• «•«««• LOW DOSE
__ _ _ HIGH DOSE
1 | 1 | 4 | 1 | | | T | 1 |
) 15 30 45 60 75 90 105 12

10

-08
-06
-
-04
-
-02

• • n n

0
                              TIME ON TEST (WEEKS)
tr
z>
O

X
CO
o
cc
0-


0.8 —
-
0.6 —
-

0.2-


.u


4 L
i i,. H 	
i \ L
. s
II *
% •
I
1_ •
^"1 A
1 •
— — — — UNTREATED CONTROL • *
1 ••••«•••
	 VEHICLE CONTROL fc^ ***•«
• «a***« LOW DOSE ~ "™ ~" ™T_ «.
— — — HIGH DOSE
i | i | i | i | i ] i | i |


— 0.8
h
— 0.6
-04

- 0.2



0 15 30 45 60 75 90 105 120
                              TIME ON TEST (WEEKS)
Figure 9-2.  Survival  comparisons of male and female mice  in  the PCE
chronic study.


SOURCE:  National  Cancer  Institute, 1977a.
                                     9-7

-------
  750
in
                                                          UNTREATED CONTROL

                                                          VEHICLE CONTROL
                                                • •••••••  LOW OObr

                                                — — — -  HIGH DOSE
                                                                             750
                                                                            — 600
                                                                           i— 450
                                                                           — 300
                                                                           — 150
                                                                          120
                                TIME ON TEST (WEEKS)
  750-r
                                                          UNTREATED CONTROL

                                                          VEHICLE CONTROL
                                                • ••••••  LOW DOSE

                                                 — — —  HIGH DOSE
                       30
                               45
                                         I
                                        60
 I
75
 I
90
                                                                             -750
                                                                            -600
                                                                            - 450
                                                                            - 300
                                                                            - 150
                                                                  105
                                                                          120
                                TIME ON TEST (WEEKS)
    Figure 9-3.   Growth  curves for  male and female rats  in the PCE  chronic  study.

    SOURCE:   National Cancer Institute, 1977a.
                                       9-8

-------


~
0.8-
_i
 -
CT
CL
0.2 —


C


• • • • • «l ' i L . i
~| *J9 l-i — -i

\ '•• L-i
T. • 1
I ••• !
n*«* i» L
***• 1 T_

>-, -.. '^ \
i \ L X

— i^ •.... i L— j
T * * *™"1 i
— *«««««»»'•« T
__-_—— UNTREATED CONTROL ' ~L™ * *] "t

._._. VEHICLE CONTROL "^- *%'
^ Tit*.
• •••••• LOW DOSE ^~- — — j_ »A
„_„»,_ HIGH DOSE ^""""fc,
i | i | 1 | 1 | [ | i | i | i
15 30 45 60 75 90 105 12



~
-0.8


""

-0.6



— 0.4

-

-0.2


0
                              TIME ON TEST (WEEKS)

0.8-
_i
a:
m 0.6 —I
LL
O
=i 0.4 -
CO
m
0
CC
CL
0.2 —




	 •/„— L_ J_ 	 1 rCMALC n ATo
*1« ~i-_ L._._>. — _ L____
... I 1
• j L-.a^,. i
'••%.ST_ -••— , L
* * * **•* H i
* "i
«jo»e* "~J
"""-—!__. ••••t

•— •— • VEHICLE CONTROL
• •••••• LOW DOSE
__ „ _ HIGH DOSE
i | i | i ] i | i | i | i | i
1 n



— -0.6
-
— 0.4
- 0.2



0 15 30 45 60 75 90 105 120
                             TIME ON TEST (WEEKS)
Figure 9-4.  Survival comparisons  of male and female rats in the PCE
chronic study.

SOURCE:  National Cancer  Institute,  1977a.
                                   9-9

-------
carcinogenicity testing laboratory, was  reproduced.   Evaluation  of  the  ana-



lytical method led to the conclusions  presented  in  the  previously mentioned



memorandum that the method could not distinguish epichlorohydrin from tri-



chloroethylene, and that the PCE product  used  in the  NCI  (1977a) bioassay



could have contained one or both of these compounds as  impurities.   The



documented chemical analyses show contaminant  levels  of 0.055%,  0.041%, and



0.010% in PCE samples analyzed at the  beginning  of  the  bioassay, at  1 year



into the bioassay, and at 2 years into the bioassay,  respectively.   The conclu-



sion stated in the memorandum is that  the contaminant was  probably  epichloro-



hydrin, since:  1) epichlorohydrin was a  commonly used  stabilizer for PCE at



that time; 2) a reported analysis by a competitor of  Diamond  Shamrock showed a



maximum of 0.015% trichloroethylene in any PCE product  manufactured  in  the



United States, and 3) the decreased amount of  impurity  found  in  the  PCE samples



as the bioassay progressed suggests some  decomposition  in  that,  according to



the experience of Diamond Shamrock, epichlorohydrin but not trichloroethylene



levels would decrease during storage.  Nonetheless, although  the evidence



provided by Diamond Shamrock indicates that epichlorohydrin was  present in the



PCE product used in the NCI (1977a) bioassay,  the quantity of epichlorohydrin



in the test material  used in this study  remains  uncertain. Furthermore, the



different levels of unknown material shown in  the documented  chemical analysis



may be due to the use of the indicated different lots,  possibly  containing



unequal amounts of impurities.



     An estimate of the likelihood that  the epichlorohydrin impurity in the  PCE



material  used in the NCI experiment could have been  large  enough to account  for



the positive results can be obtained by  considering the results  of  Laskin  et al.



(1980), who exposed rats to epichlorohydrin via  inhalation.   They  found that



1/100 rats exposed to 30 ppm (0.115 mg/L air,  equivalent  to  13.2 mg/kg/day)  of






                                      9-10

-------
epichlorohydrin for 6 hours/day, 5 days/week, for 730 days developed squamous



cell nasal carcinomas, as compared wit.h 0/50 in control animals.



     The epichlorohydrin impurity in the NCI PCE high-dose male mice experiment



was approximately 1,032 mg/kg/day x 0.041% = 0.42 mg/kg/day, from the discus-



sion above.  This is only about 0.42/13.2 = 0.03 times the dose that gave an



incidence of only 1/100 in the Laskin et al. (1980) rat experiment.   Therefore,



it is unlikely that epichlorohydrin impurities at the level  estimated to be



present in the NCI experiment could have contributed appreciably to  the positive



response.



     In a second carcinogenicity bioassay sponsored by the National  Toxicology



Program (NTP, 1983, draft), PCE was given orally to 86C3F1 mice and  to four



strains of rats (Sherman, Fischer 344, Long-Evans, and Wistar).  The PCE sam-



ple used in this ongoing carcinogenicity bioassay did not  contain detectable



amounts of epoxide contaminants.  In the study with female B6C3F1 mice (NTP,



1983, draft), groups of 100 females received 25, 50, 100,  or 200 mg  of more



than 99% pure PCE per kg body weight in corn oil by gavage for 103 weeks, 5



days/week.  Vehicle-control and untreated control groups of  100 mice each were



used.  Survival was not affected significantly.  Doses of  50 mg/kg or more



produced a dose-related cytomegaly of the kidneys.  Other  signs of liver toxi-



city appeared during the study as increases of sorbitol dehydrogenase activity,



and increases in relative liver weight and lipid levels were associated with



increasing dose levels.  Doses of 50 mg/kg or more produced  time- and dose-



related increases in the incidence of hepatocellular adenomas and carcinomas



bearing no causal  relationship to the observed renal damage.  The first adenoma



appeared at week 46; the first carcinoma at week 58.  The  NTP intends to con-



duct an audit of the raw data of this study before the final technical report



is prepared.   The findings of the audit will  determine the validity  of the





                                      9-11

-------
study.  The study remains under audit as of June 1985.





9.1.2  Dow Chemical  Company Inhalation Study (Rampy et al.,  1978)



     Two groups of weanling Sprague-Dawley (Spartan substrain) rats, each com-



posed of 96 males and 96 females, were exposed to 600 ppm (4068 mg/m3 air) or



300 ppm (2034 mg/m3 air) of PCE 6 hours/day, 5 days/week,  for 52 weeks.   An un-



treated group of 192 males and 192 females served as controls.  Controls were



not put in inhalation chambers, but were in the treatment  room during exposure.



At the end of the treatment period of 52 weeks, animals were allowed to  sur-



vive until sacrifice at 31 months.  The composition of the test material  (Lot



A12282D)  by gas chromatographic analysis was as follows:  trichloroethylene, 3



ppm (liquid volume %); hexachloroethane, < 12 ppm;  carbon tetrachloride,  2 ppm;



4-methyl  morpholine, 44 ppm; nonvolatile residue, 2 ppm;  PCE, balance.   Expo-



sure was  done in 3.7 m3 inhalation chambers with a  dynamic airflow system.



Analyses  of PCE levels in the inhalation chambers during  the treatment  period



revealed  analytical  concentrations (mean +_ standard deviation) of 310 +_ 32 ppm



(244 analyses) and 592 +_ 62 ppm (1,245 analyses).  Analytical  concentrations



by infrared analysis within +_ 10% of nominal  levels were  achieved on 89.3%



(low-dose) and 97.1% (high-dose)  of the exposure days.  Animals were evaluated



for clinical  toxic signs, body weight changes, urinalysis at 24 months,  and



hematology at 12 and 24 months.  Survivors and decedents  were given gross and



histopathologic examinations.  Bone marrow samples  were taken from three males



and three females in each group sacrificed at 1 year for  cytogenetic evalua-



tion.



     Clinical  signs  of toxicity were not observed with the nominal  concentra-



tions of  PCE  used in this study.   Mean body weight  gains  were similar among



groups.  Hematology  and urine analyses showed no treatment-related effects
                                      9-12

-------
of PCE.  The mortality patterns exhibited in the study are described in
Table 9-2.  Mortality in high-dose males was slightly greater than in con-
trols during months 5 to 24; the earlier onset of chronic renal  disease in
this treatment group was considered to be a contributing factor  in increased
mortality.
             TABLE 9-2.  CUMULATIVE SURVIVAL OF SPRAGUE-DAWLEY RATS
                          EXPOSED TO PCE FOR 12 MONTHS
Month
of study
Initial
6
12
18
24
31
0
Males
189
187
183
155
44
1
ppm
Females
189
188
185
151
70
12
300
Males
94
94
91
74
26
1
ppm
Females
91
91
91
77
37
6
600
Males
94
88a
84 a
55a
13a
1
ppm
Females
94
94
91
86a
493
5
ap < 0.005 by the Fisher Exact Test.
SOURCE:  Adapted from Rampy et al., 1978.

     No carcinogenic effects of PCE were observed from pathologic examination
of the animals.  Statistical analysis of the data showed numerous nonneoplastic
abnormalities that occurred spontaneously and were within the normal  variation
encountered in lifetime studies with this strain of rat.  With respect to tumor
findings, analysis of the data did not reveal a definite increased tumor inci-
dence in animals exposed to PCE.  Tumors or tumor-like changes in the kidney
were found in 1/189 control, 2/94 low-dose, and 4/94 high-dose males  during
                                     9-13

-------
gross necropsy;  however,  microscopic examination  of  kidney  lesions  did  not  show



a statistically  significant  tumor  incidence  compared to  controls.   Although



many tumor types were found  in  treated  and  control  animals,  there was  no  sta-



tistically significant (p >  0.05)  increase  in  tumor  incidence.



     The results of this  study  do  not indicate a  definite  carcinogenic  effect



of PCE in Sprague-Dawley  rats;  i.e., the  tumor incidence between control  and



treated rats was similar.  However,  this  study has  the following drawbacks:



 1) the period of exposure was only 12 months  rather  than the lifetime  of  the



 animals, which would have been  a more appropriate duration  for  carcinogenicity



 studies; and 2)  the dose  levels in this  study  do  not appear  to  have been  high



 enough to provide maximum sensitivity.






 9.1.3  Intraperitoneal Administration Study  (Theiss  et al.,  1977)



     Theiss et al. (1977) tested PCE for  carcinogenicity in  the strain  A  mouse



pulmonary tumor  induction system.   The  test  sample,  a product of the Aldrich



Chemical Company, was reagent grade  with  a  purity exceeding  95% to  99%.   Strain



A/St male mice,  6 to 8 weeks old,  were  used  in this  assay.   The maximum tolera-



ted dosage, defined as the dosage  which  five mice tolerated  after six  intra-



peritoneal injections over a 2-week  period  followed  by a 4-week observation



period, was determined and used in the  bioassay.  In the main test, 20  mice per



treatment group  received  three  intraperitoneal  injections  of 80, 200,  or  400



mg/kg of PCE weekly until total dosages  of  1,120, 4,800, and 9,600  mg/kg,



respectively, were achieved.  Survivors were sacrificed  at  24 weeks after the



first injection, and the  number of surface  adenomas  was  counted.  Results were



compared with findings in vehicle  (tricaprylin) and  untreated controls  by the



Student t test.   PCE did  not statistically  increase  (p > 0.05)  the  incidence of



pulmonary tumors in this  study  (Table 9-3).  This strain was sensitive  to the
                                      9-14

-------
positive control chemical urethane, as shown in Table 9-3.



     A negative result in this assay is not considered conclusive,  since



several chemicals known to be carcinogenic in chronic rodent  bioassays  induce



no response in this assay.
                  TABLE 9-3.  PULMONARY TUMOR RESPONSE TO PCE
Compound
Tri capryl i n
PCE
Urethan
Dosage (mg/kg)
—
80
200
400
1,000 mg/kg
(1 injection)
No. survivors/
No. animals
46/50
15/20
17/20
18/20
20/20
No. lung
tumors/mouse
0.39 +; 0.06a
0.27 + 0.07
0.41 + 0.10
0.50 _+ 0.12
19.6 _+ 2.4
aMean +_ S.E.



SOURCE:  Adapted from Theiss et al., 1977.







     The strain A mouse pulmonary tumor assay is  relatively  insensitive  to



mouse carcinogens for which the effect is confined to the liver  (Theiss  et  al.,



1977).  For example, chloroform, 2-chloroethyl  ether, and hexachlorocyclohexane



induce tumors of the liver (not other sites)  in mice (NCI,  1976,  1977b;  Innes



et al., 1969) but were not carcinogenic in  the  assay by  Theiss et  al.  (1977).



The reasons for the negative lung response  are  not understood, but  it  may be



due to a smaller concentration of activating  enzymes in  the  lung  than  in the



liver.
                                      9-15

-------
9.1.4  Skin Painting Study (Van Duuren et al.,  1979)



     A carcinogenicity study of purified PCE in ICR/Ha Swiss mice was  described



by Van Duuren et al. (1979).  Maximum tolerated dosages were determined in



range-finding studies 6 to 8 weeks in duration, and were selected as  dosages



that did not affect body weight gain or produce clinical  signs  of toxicity.



This study included the following experiments:   1)  30 females were treated



topically on the dorsal skin with a single application of 163 mg  PCE  followed



14 days later by the applications of 5.0 yg phorbol myristate acetate  (PMA)



to the skin three times weekly until termination of the study at  428 to 576



days; median survival time was 428 to 576 days.  2) 30 females  were given



thrice weekly topical applications of 54 mg PCE for the duration  of the test



(440 to 594 days), with a median survival time  of 317 to 589 days.  A  vehicle



(acetone) control group of 30 mice and an untreated control  group of 100 mice



were included in these experiments.



     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, and were housed six to a cage.  Test sites on  the skin were  shaved as



necessary and were not covered; however, it was the authors'  impression that



PCE was immediately absorbed and that evaporation from test  sites was  minimal



(personal communication, B. L. Van Duuren, New  York University).   The  animals



were weighed monthly, and each animal was examined  by necropsy.   Tumors and



lesions as well  as skin, liver, stomach, and  kidneys  were examined  histologic-



ally.



     PCE did not show initiating activity in  the initiation-promotion  experi-



ment; the number of mice with skin papillomas (squamous  cell  carcinomas) was:



4 (0) initiated with PCE, 15 (3) treated with PMA alone,  and 0  (0)  in  the



control  groups.   The study involving repeated application to the  skin  produced





                                      9-16

-------
lung and stomach tumors in 16 and 0 PCE-treated mice, respectively; 11 and 2



vehicle-treated controls, respectively; and 30 and 5 untreated controls,



respectively.



     The negative results for PCE in mouse skin as an initiator and as a com-



plete carcinogen can be reconciled with the positive mouse liver response in



the NCI study if it is hypothesized that the skin does not have the necessary



enzymes to convert PCE to an active metabolite, whereas the liver does have



this capability.



     The lack of sensitivity of skin application tests, as compared with tests



using other routes of exposure, is apparent from the results of Van Duuren et



al. (1979) with 1-chloroprene, cis-1,3-dichloropropene, and 2-chloroproponal.



They found that none of the three compounds induced a response as initiators in



initiation-promotion experiments or with repeated topical  application on the



skin.  However, they did observe a statistically significant increase in the



incidence of forestomach tumors in female Ha:ICR Swiss mice dosed by gavage



with 1-chloroprene (p < 0.0005) and 2-chloroproponal  (p <  0.05) and in the



incidence of local sarcomas in female Ha:ICR Swiss mice treated with cis-1,3-



dichloropropene (p < 0.0005) by subcutaneous injection.





9.2  EPIDEMIOLOGIC STUDIES



     There are seven epidemiologic studies either completed or currently in



progress that relate to PCE exposure.   Only two of these studies, however, have



actually identified persons exposed to PCE.  Because PCE has been used in the



dry-cleaning industry, however, the present discussion also includes three pro-



portionate mortality studies of decedents who had worked in the dry-cleaning



industry,  as well  as two case-control  studies in which the cases and controls



were asked about their employment histories, including employment in the dry-
                                      9-17

-------
cleaning industry.





9.2.1  Kaplan (1980)



     Kaplan (1980) did a retrospective  cohort  mortality  study  of dry-cleaning



workers exposed to PCE for at  least  one year prior  to  1960.   The study was



performed under contract to the Biometry Section  of the  National Institute for



Occupational  Safety and Health (NIOSH)  Industry-Wide Studies  Branch.



     In a preface to the discussion  of  the  study, Kaplan reported that levels



of PCE exposure were "much higher"  for  cleaners  (machine operators) than  for



other employees of dry-cleaning establishments.   A  geometric  mean time-weighted



average for PCE was 22 ppm for the  machine  operators.  For all  other  jobs, the



highest corresponding value was reported to be 3.3  ppm.   These  data were  provid-



ed through a NIOSH industrial  hygiene survey of dry-cleaning  facilities.



     The study cohort, selected from records maintained  by several  labor



unions, consisted of 1,597 dry-cleaners exposed for more than  1 year  prior to



1960.  The primary solvent used was  PCE.  Efforts were made by  the author to



exclude all persons with previous  occupational exposure  to carbon tetrachloride



or trichloroethylene.  By September 30, 1977,  the end  of the  study period,



1,058 individuals were found to be  alive, 285  were  deceased,  and 254  were of



unknown vital status.  The extent  of follow-up varied  by sex; 8% of the males



and 20.4% of the females remained  lost  to follow-up.   Race was  known  only for



deceased workers, and was obtained  from death  certificate data.   Because  of the



lack of information regarding  race,  observed deaths by cause were compared to



expected deaths by means of a  standardized  mortality ratio (SMR) for  whites,  an



SMR for blacks, and a composite point estimate SMR  for both.  Based on the



assumption that every member of the cohort  was white,  expected  deaths  for



whites were derived by multiplying  the  person-years accumulated for the entire
                                      9-18

-------
cohort by the white death rates within 5-year age groups (separately for males

and females and then with the two combined).  Expected deaths for blacks were

similarly generated by assuming that all  cohort members were black.   Composite

expected deaths were calculated by weighting the total accumulated person-years

for the cohort in each 5-year age group by the proportion of person-years

attributable to the deceased blacks and the proportion attributable  to deceased

whites, and then multiplying each total separately by the corresponding death

rates for whites and blacks, and finally  adding across age groups to get the

"composite" expected deaths.

     Because death certificates could not be located for all of the  deceased,

it was assumed that those deaths for which no death certificates could be found

had the same distribution by cause as those for which death certificates were

available.  Thus the SMRs for each cause  of death were corrected to  reflect

the missing death certificates.  Using the SMR for deaths from malignant neo-

plasms of the colon in whites as an example, this correction was made in the

following manner:


                     (11/247 x 38) + 11  x  11  x 100 = 182
                             11            6.98


where:  247 is the number of deaths in the cohort with death certificates;

        38 is the number of deaths in the cohort without death certificates;

        11 is the observed number of colon cancer deaths identified  by death
           certi fi cates;

        6.98 is the expected number of white colon cancer deaths;

        100 is a constant used in calculating SMRs (by convention, SMRs are
            expressed as a factor of 100); and

        182 is the corrected white colon  cancer SMR.


No tests of significance of any of the SMRs were done by the author.  However, the


                                      9-19

-------
author called attention to the elevated SMR from malignant neoplasms  of the colon



as possibly related to occupational  exposure (6.98 white expected deaths,  6.77



black expected deaths, and composite SMR = 182).  Using an observed number of



colon cancer cases corrected for the loss of death certificates,  the  Carcinogen



Assessment Group (CAG) found that the SMR for either whites or blacks would be



statistically significant (p < 0.05).



     The author points out that because the expected numbers of deaths were



calculated using U.S. rates, which include a higher socioeconomic class than



the dry-cleaners in this study, and because higher socioeconomic  class is



associated with a risk of colon cancer, the risk of colon cancer  from exposure



to PCE found in this study is probably underestimated.   Furthermore,  the expec-



ted number of colon cancer cases in this study was calculated using mortality



rates for neoplasms of the intestine, except rectum.   Although most of the



deaths expected using mortality rates for neoplasms of  the intestine  would be



deaths from malignant neoplasms of the colon, some deaths would be from neo-



plasms of the small intestine.  Since all of the observed deaths  were from



malignant neoplasms of the colon, the comparison of the observed  to expected



colon cancer deaths -in this study would also tend to underestimate the colon



cancer risk from exposure to PCE.



     It should also be noted that although it could be  argued that the number



of observed colon cancer deaths in each of the four union locals  in this study



was small, an elevated colon cancer SMR did exist in each of the  four locals.



Finally, it should be noted that the colon cancer SMR appeared to demonstrate



a positive correlation with the length of the follow-up period.  This last



finding must be viewed with a great deal of caution,  however, because of the



author's difficulty in defining length of exposure and  hence length of follow-



up.





                                      9-20

-------
     In addition to colon cancer, the SMRs for cancer of various other sites



were also elevated.  These included rectum, pancreas, respiratory system,



urinary organs, and "other and unspecified sites (major)."   None of these  were



significant at the p < 0.05 level when tested using an observed number that  was



corrected for lost death certificates; however,  the SMRs for cancer of three of



these sites [respiratory system, urinary organs, and "other and unspecified



sites (major)"] could be considered borderline significant  (0.10 < p < 0.05).



     Perhaps the major weakness of this study with  regard to evaluating PCE  as



a carcinogen is that the history of solvent exposure prior  to 1960 was unknown



for nearly half of the union member shops.  Because the majority of dry-clean-



ing establishments in the United States used petroleum distillates as  the



primary cleaning agent prior to 1960, it is quite possible  that most of the



shops in this study used petroleum distillates as the cleaning solvent prior to



changing to PCE.



     Other important confounding variables were  also not controlled.  For



example, smoking is a major confounding variable to be considered when evalu-



ating a potential risk for respiratory or bladder cancer, both of which were



found in excess in this study.  Socioeconomic status, as has been discussed,



is a confounding variable for colon cancer.



     Another weakness of this study is that 16%  of  the study cohort was



lost to follow-up.  Currently, NIOSH is attempting  to improve the percentage



of follow-up as well as to add to the length of  follow-up.   In addition, NIOSH



has identified other individuals who were exposed to PCE for at least  one



year prior to 1960, so that the size of the cohort  has also been increased.



The results should be available by 1986.



     In summary, this study appears suggestive that dry-cleaning workers ex-



posed to PCE are at an elevated risk of colon cancer mortality.  Potential





                                      9-21

-------
exposure to petroleum distillates for approximately half of the cohort, how-



ever, limits any conclusions with regard to the carcinogenicity of PCE in



humans.






9.2.2  Blair et al.  (1979)



     Blair et al.  (1979) reported, as the preliminary results of a cohort



study of 10,000 laundry and dry-cleaning workers,  a proportionate mortality



analysis of 330 of the workers who had died during the period 1957-1977.



Deaths were identified from the mortality records  of two union locals  in  St.



Louis, Missouri.  The distribution by cause of death among  the 330 was compared



to that expected based on the proportionate mortality experience of the United



States.  Only 279  of the 330 had worked exclusively in dry-cleaning establish-



ments; however, the authors did not report  the proportionate  mortality analyses



of these 279.  Furthermore, the dry-cleaning agent used by  the dry-cleaners in



this study is unknown.  Among the 330 deaths,  deaths from cancer of the lung,



cervix, and skin contributed to the finding of a significant  (p < 0.05) excess



proportion of deaths from cancer at all  sites.  For lung cancer, there were 17



deaths observed versus 10 expected (p < 0.05); for cervical  cancer, 10 observed



deaths versus 4.8  expected  (p < 0.05); and  for skin cancer,  3 observed deaths



versus 0.7 expected (p < 0.05).  On the other  hand, a significant deficit of



deaths occurred in the cause category identified as "all  circulating diseases,"



with 100 observed  versus 125.9 expected (p  < 0.005)--a finding that could



account  for the excess proportion of cancers observed.



     Only limited  conclusions can be drawn  from this study  as to the potential



carcinogenicity of PCE.  This is true for several  reasons.   First, the study



did not  report the distribution of deaths for  dry-cleaners  alone.  Second,  the



dry-cleaning agent  these workers used is not known.  Lastly,  certain possibly
                                     9-22

-------
confounding variables were not considered in the analysis of the results.



These variables include smoking (with.regard to the lung cancer excess),



socioeconomic status (with regard to the cervical  cancer excess),  and  sunlight



exposure (with regard to the skin cancer excess).





9.2.3  Katz and Jowett (1981)



     Katz and Jowett (1981) analyzed the death certificate records  of  671  white



female laundry and dry-cleaning workers who had died during the period 1963-



1977.  The records of dry-cleaning workers were not studied separately from



those of laundry workers, since both groups of workers shared the  same occupa-



tional code.   Furthermore, it is not known whether those decedents  who were



dry-cleaners  used PCE as a dry-cleaning agent.  During the 1940s,  petroleum



derivatives were the predominant dry-cleaning agents used in the United States.



A shift to the use of PCE began in the late 1940s  and gained momentum  in the



1950s and 1960s.  By 1977, approximately 75% of the dry-cleaning plants in the



United States were using PCE.  However, in the period before 1960,  petroleum



distillates were still the dominant solvents in use (Kaplan, 1980).



     In this  study, cause-specific proportionate mortality for the  671 de-



ceased laundry and dry-cleaning workers was compared to that for the deaths of



all other working females in Wisconsin during the  same period, and  to  deaths



of females in "lower-wage occupations"  in Wisconsin during the same period.



Significantly elevated proportionate mortality ratios (PMRs) were  found for



deaths from cancer of the genitals (unspecified) (p < 0.01) and for deaths



from cancer of the kidney (p < 0.05) when deaths among women of all occupations



and deaths among women of "lower-wage occupations" were used as the comparison



groups.  In summary, this study, although suggesting an association of employ-



ment in the dry-cleaning industry with  excess risks of certain types of cancer,
                                      9-23

-------
cannot be said to demonstrate that  PCE  exposure  presents  a  carcinogenic  risk



because of the possible confounding influence  of other  dry-cleaning  agents,  and



because of the fact that no distinction was  made between  laundry  and dry-clean-




i ng workers.






9.2.4  Lin and Kessler (1981)



     Lin and Kessler (1981) did a case-control study  of 109 pancreatic cancer



cases diagnosed during the period 1972-1975  from 115  hospitals  in  five metro-



politan areas.  The control group was  composed of subjects  who  were  free  from



cancer but who were similar to the  study group in age (+_  3  years), sex,  race,



and marital status, and who had been selected  at random from among contempora-



neous admissions to the same hospital.   Cases  and controls  were asked about



demographic characteristics, residential  history, toxic exposures, animal  con-



tacts, smoking habits, diet, medical  history,  medications,  and  family history.



Males were asked about sexual practices and  urogenital  conditions.   Females



were questioned on these topics and also on  their marital,  obstetric, and



gynecologic histories.  Among other statistically significant associations,  an



association was found between pancreatic cancer  and employment  either as  a dry-



cleaner or in a job involving exposure  to gasoline.   It is  not  known, however,



how many individuals with pancreatic cancer  were employed as dry-cleaners and



how many were employed in occupations  involving  exposure  to gasoline.  Further-



more, the dry-cleaners may have used a  variety of dry-cleaning  agents other



than PCE.  Thus this study, although suggestive  of an association  between



employment as a dry-cleaner and an  excess risk of pancreatic cancer, cannot  be



said to demonstrate an association  between pancreatic cancer and  PCE exposure.





9.2.5  Duh and Asal (1984)



     Duh and Asal (1984) analyzed the  death  certificates  of 440 laundry  and





                                      9-24

-------
dry-cleaning workers who had died in Oklahoma during the period 1975-1981.



Occupations were identified from the death certificates.  Laundry workers could



not be separated from dry-cleaning workers because they shared similar occupa-



tional codes.  Age, sex, race, and cause distribution of deaths in the United



States in 1978 were used as the standards of comparison.  An analysis was done



using the Standardized Mortality Odds Ratio (SMOR) (Miettinen and Wang, 1981).



The SMOR differs from the more traditional Proportionate Mortality Ratio (PMR)



in that it compares the number of deaths from the cause of interest to the



number of deaths from auxiliary causes in the exposed population (the odds)  to



the expected odds derived from a comparison population, taking into account



covariates such as age, sex, and race.  The authors used "other diseases" and



"all circulatory diseases" as auxiliary causes for the analyses, and indicated



that the results were almost the same as that of the PMR method when "all other



diseases" were used as auxiliary causes, but they differed on two points when



"all circulatory diseases" were used as auxiliary causes.  (The authors did  not



indicate what these two points were.)  The SMOR for homicide was the only SMOR



in the study group that was significantly elevated for non-cancer causes of



death (SMOR = 3.8, 95% confidence interval (CI) = 1.4-10.6;  observed deaths  =



7, expected deaths = 2.9).  Deaths from ischemic heart disease were signifi-



cantly decreased (SMOR = 0.8, 95% CI = 0.7-1.0; observed = 134, expected =



153.7).  Among deaths from cancer, the SMORs for cancer of the respiratory



system (SMOR = 1.8, 95% CI = 1.3-2.5; observed - 39, expected = 23.8); cancer



of the lung (SMOR = 1.7, CI = 1.2-2.5; observed = 37, expected = 22.6); and



cancer of the kidney (SMOR = 3.8, 95% CI = 1.9-7.6; observed = 7, expected =



1.9) were significantly elevated.  Deaths from cancer of the breast were



significantly decreased (SMOR = 0.1, 95% CI = 0.0-0.4; observed = 1, expected



= 10.5).   Since, as the authors noted, wage levels in the laundry and dry-






                                      9-25

-------
cleaning industry are not high,  the excess  of homicide and the deficit of
breast cancer and ischemic heart disease may  have resulted from the inverse
correlation with socioeconomic status.
     Smoking histories were not  available for any of the death certificates
analyzed.  Because both lung and kidney cancer have  been associated with
smoking, it is difficult to evaluate the study's  elevated SMORs for those
diseases with respect to dry-cleaning exposures.   The authors  did  note, how-
ever, that mortality from other  smoking-related diseases such  as emphysema,
ischemic heart disease, cancer of the buccal  cavity, and pharyngeal  cancer was
slightly less than expected.
     As the authors noted, there may have been a  bias in the analysis  because
the occupational code that identified the study group included both laundry and
dry-cleaning workers.  Such a bias would tend to  weaken any association that
might have been present by adding to the group at potential  risk a number of
individuals that would not otherwise have been considered exposed.
     Of interest is the authors' report that  petroleum solvents accounted for
greater than 50% of the dry-cleaning solvents used in Oklahoma in  1983.
As with several of the other epidemiologic  studies reviewed here,  this study
suggests that dry-cleaners and/or laundry workers may be at an excess  risk of
cancer.  However, the possible confounding  effects of both petroleum solvents
and smoking, and the fact that dry-cleaners could not be separated from laundry
workers in the analysis, preclude the conclusion  that PCE exposure is  associ-
ated with an excess risk of cancer.

9.2.6  Asal (personal communication, 1985)
     Dr. Nabih Asal of the University of Oklahoma School of Public Health is
currently analyzing the data from a kidney  cancer case-control study using both
                                      9-26

-------
hospital and population controls.  A variety of potential  risk factors are



being evaluated including occupation.  While information will  not be available



from this study on cases and controls who worked in the dry-cleaning industry



with regard to exposure to specific dry-cleaning solvents, Dr. Asal  is also



conducting a cohort study of dry-cleaning establishment owner-operators.   Data



on solvent use is extensive for the different dry-cleaning establishments, and



thus it is hoped that this study will provide useful  data with which to evalu-



ate PCE with regard to carcinogenicity.





9.3  RISK ESTIMATES FROM ANIMAL DATA



     The evidence for the carcinogenicity of PCE serving as the basis for the



quantitation of risks consists of a single positive mouse study (NCI, 1977a).



In this study, there was a statistically significant  increase  in the incidence



of malignant tumors of the liver in both sexes of B6C3F1 mice.



     Because of a question regarding the possible presence of  contaminant



epoxides in PCE (< 0.05% the maximum possible), the assay was  repeated (NTP,



1983, draft) with PCE without detectable amounts of epoxides,  although at lower



dose levels.  Doses of 50 mg/kg to 200 mg/kg produced time- and dose-related



increased incidences of hepatocarcinomas.  At the present  time, the  NTP final



report is not available.



     The cancer data from epidemiologic  studies are inconclusive.  The evidence



for mutagenicity of PCE is also inconclusive.  However, the more conservative



public health view would regard PCE as a possible human carcinogen.   The  calcu-



lations in this document for estimating, in quantitative terms, the  impact of



PCE as a carcinogen are made independently of the overall  weight of  evidence



for the carcinogenicity of PCE.  The calculations are made as  if PCE were a



human carcinogen.
                                      9-27

-------
9.3.1  Selection of Animal  Data



     For some chemicals, several studies in different animal  species,  strains,



and sexes, each run at several  doses and different  routes  of  exposure,  are



available.  A choice must be made as to which  of these data sets  to use in



the mathematical extrapolation  model.  It may  also  be appropriate to correct



for metabolic differences between species and  for different routes of  admini-



stration.



     For PCE, only the data from the NCI (1977a) oral  gavage  lifetime  study



in male and female mice, showing a significant excess  of tumors  (hepatocellu-



lar carcinomas) have been used.  Table 9-4 gives the  time-weighted average



dosage and tumor incidence (at  "low" and "high" doses) in  male and female



B6C3F1 mice for the NCI study.   PCE in corn oil was given,  once a day,  by



gastric intubation, 5 days/week, starting at 25 days  of mouse age and  con-



tinuing for 78 weeks.  During the final  26 weeks, doses were  given in  a cyclic



pattern of one week without treatment followed by four weeks  with treatment.



After the 78-week treatment period, the mice were given an  additional  12 weeks



before sacrifice.



     Taking into account differing dosage levels, male and  female mice  appear



to be equally susceptible to a  carcinogenic response  to PCE.  Furthermore,



since the "low" dose and "high" dose response  did not  differ  statistically,



a plateauing of the response is indicated.  Hepatic tumors  (0-10%) were not



detected in untreated or vehicle-treated groups until  90-91 weeks.  For treated



male mice, tumors were detected at 27 weeks in the  "low" dose group and 40



weeks in the "high" dose group, and for treated female mice at 41 weeks in the



"low" dose group and 50 weeks in the "high" dose group.
                                      9-28

-------
          TABLE  9-4.   INCIDENCE  OF  HEPATOCELLULAR  CARCINOMAS  IN MALE AND  FEMALE B6C3F1 MICE FED  PCE  BY  GAVAGE
I
ro
Sex and no.
mice
Males
20
20
50
50
Femal es
20
20
50
50
aSum (dose in
Median Average
survival terminal
(weeks) weight (g)
>90
>90
30
78
43
>90
>90
25
62
50
mg/kg x no. of days at that dose)
Time-weighted average
gavage dose9
mg/kg/day
untreated
0; vehicleb
536
1,072
untreated
0; vehicle
386
772

mg/animal /day
0
0
16.1
32.2
0
0
9.7
19.3

Incidence
2/17 (12%)
2/20 (10%)
32/49 (65%)
27/48 (56%)
2/20 (10%)
0/20 ( 0%)
19/48 (40%)
19/48 (40%)

    Sum  (no.  of  days  receiving  any  dose)
   bCorn  oil.

   SOURCE:   NCI,  1977a.

-------
9.3.2  Interspecies Dose Conversion
9.3.2.1  General  Consi derations--Sea ling of  toxicologic  effects,  including car-
cinogenicity, among species has been elaborated  in  several  published  papers
(Freireich et a!., 1966; Gillette,  1976; Krasovski,  1976;  Dedrick,  1973;  Rail,
1977; Dedrick and Bischoff, 1980;  and Gehring  et al.,  1980).   The major  compo-
nents requiring consideration in determining an  appropriate extrapolation  base
for scaling carcinogenicity data from laboratory animals to man are:   1)  toxico-
logic data, 2) metabolism and kinetics,  and  3) covalent  binding.  The  published
literature concerning the biological basis for extrapolation  of the dose-carci-
nogenic response relationship of laboratory  animals  to man  has been reviewed
in a paper prepared by Davidson (1984) for the Carcinogen  Assessment  Group
and presented by Parker and Davidson (1984).
9.3.2.1.1  Toxicologic Data.  Across species (mice,  rabbit,  cat,  dog,  and  man)
the acute 1059 of PCE is similar (Tables 6-1 and 6-2;  Chapter 6).  This  simi-
larity is most probably due to an  equally effective  concentration, leading to
central nervous system depression  and resulting  in  respiratory depression  and
death.  While other toxicologic end  points (modalities of  liver and kidney
damage) have been measured (Klassen  and Plaa,  1967), these  measures are  dif-
ficult to evaluate in the numerous  and separate  studies, and  comprehensive com-
parative studies  have not been conducted across  species  using the same toxico-
logic end point.
9.3.2.1.2  Metabolism and Kinetics.   It  is generally accepted that the liver
and kidney toxicity. and carcinogenicity potential  of  the  halogenated ethy-
lenes relate to the metabolic conversion of  these compounds to biologically
reactive intermediates  (reviewed by  Bolt et  al., 1982).  All  of the chlorinated
ethylenes are initially biotransformed by microsomal monooxygenases to epoxide
intermediates which bind to cellular macromolecules, are molecularly  rearranged

                                      9-30

-------
to less reactive species, or are modified by conjugation through the "detoxi-



fication mechanisms."  Bolt et al. (1982) postulated that the reactivities and



hence toxicities of the individual epoxides formed from chlorinated ethenes



depend on the type of chlorine substitution.  Symmetric substitution such as



is seen with PCE renders the epoxide relatively more stable and therefore less



reactive, while asymmetric substitution results in highly unstable reactive



epoxides (e.g., vinyl chloride).  However, tumors may arise in animals  by



genetic or epigenetic mechanisms in which recurrent cytotoxicity may lead



to the development of cancer.  For example, the accumulative effect of  even



minimally reactive metabolites may produce cancer.  Therefore, the extent of



metabolic activation- of these metabolites, across species, provides a compara-



tive index of cellular toxicity and a basis for determining a "mouse to man"



extrapolation scaling factor.  Moreover, such a comparison provides one index



in the evaluation of the "susceptibility" of a species (e.g., mouse versus



rat).



     The metabolism and kinetics of PCE have been studied in three species:



mouse, rat, and man.  The principal end metabolites of PCE that have been



reported as common to these species are C02 and the urinary metabolites



trichloroacetic acid (TCA) and oxalic acid.  Pegg et al. (1979) specifically



identified the major urinary metabolite in rats as oxalic acid.  There  is



strong evidence that PCE oxide is formed by mixed-function oxidase (MFO) systems



as a precursor of these metabolites,  and the additional reactive intermediate



trichloroacetyl chloride as a precursor of TCA (Bolt et al., 1982).  In



general, all of the end metabolites have been poorly identified and quantified



in mouse, rat, and man, and there is  no comparative experimental evidence that



the metabolic pathway(s) for PCE qualitatively differs for these species.
                                      9-31

-------
9.3.2.1.2.1  Oral  Studies.  Comparative balance studies  for the disposition of



PCE (extent of metabolism) have been carried out after both oral  and inhalation



exposure in both rats and mice.  The oral  exposure studies  are  the most perti-



nent to the conditions of the NCI-PCE bioassay.



     Pegg et al. (1979) and Schumann et al.  (1980) administered 14C-PCE in  corn



oil vehicle as single intragastric doses of  1 and 500 mg/kg to  Sprague-Dawley



rats and B6C3F1 mice.  14C-radioactivity was determined  for exhaled breath,



urine, feces, and carcass for 72 hours following dosing.   In addition,  pulmonary



excretion of unchanged PCE was measured.  The data from  these two studies are



collated and re-expressed in milligram equivalent units  in  Table  9-5, thus



allowing comparison of the metabolism of PCE in the rat  and mouse.



     As indicated in Table 9-5, the recovery (95-97%)  of  ^C-radioactivity  in



these experiments indicates that virtually  complete absorption  of PCE from  the



gastrointestinal tract occurs in the rat and mouse.  In  the rat,  peak blood



concentration of PCE occurred 1 hour after  oral  dosing;  disappearance followed



apparent first-order kinetics with a half-life of 6 hours  (a reflection of



first-order kinetics for pulmonary excretion and not for  metabolism).   This



relatively long half-life suggests that PCE  doses are not  eliminated completely



from the body in less than 24 to 30 hours.



     For rats, oral doses of 1 and 500 mg/kg were metabolized 29% (0.07 mg)



and 10% (12.52 mg)  respectively, while 71%  and 90% of the  doses were excreted



unchanged through  the lungs.  Thus, even at  low doses (1  mg/kg) the metabo-



lism of PCE is limited and saturable.  For  mice, the metabolism of PCE  can  be



compared to rats at the 500 mg/kg dose.  Mice metabolized  17% (1.75 mg) of  this



dose and rats 10% (12.52 mg).  In the rats,  the greater  portion of the  dose



(83%)  was eliminated unchanged through the  lungs, indicating limited metabolism



in this species also.  These observations  of very limited  metabolism by rats





                                      9-32

-------
                   TABLE  9-5.   DISPOSITION OF  14C_PCE  RADIOACTIVITY  FOR  72 HOURS  AFTER  SINGLE  ORAL  DOSES TO
                                              SPRAGUE-DAWLEY  RATS  AND  B6C3F1  MICE
UD
I
GJ
CO

Expi red
unchanged
Metabolized
14co2
Urine
Feces
Carcass
Total
Rats
1 mg/kg
(0.25 mg/animal )
Mg-eq
per animal
0.174 (71%)

0.007
0.040
0.015
0.008
0.070 (29%)
0.244
(av. of 3)a
500 mg/kg
(125 mg/animal )
Mg-eq
per animal
110.67 (90%)

0.57
5.72
4.82
1.41
12.52 (10%)
123.19
Mice (av. of 3)
500 mg/kg
(12.25 mg/animal )
Mg-eq
per animal
8.90 (83%)

0.14
1.53
0.13
0.05
1.85 (17%)
10.75
      aBased  on  average  experimental  animal  weight  (grams):  250,  rat;  24.5,  mouse,

      SOURCES:   Adapted  from Pegg et  al.,  1979 and  Schumann  et  al.,  1980.

-------
 and mice are in accord with those of Filser and Bolt (1979) who, for the rat,

 estimated Vmax at  less than 1.2 mg/kg/hour.  Using this estimate for a 250-g

 rat, 7 mg of an absolute dose of 125 mg (500 mg/kg) would be expected to be

 metabolized in 24  hours.  This compares favorably with the 12.52 mg experimen-

 tal value given in Table 9-5.

     The disposition of the oral dose of 500 mg/kg to rats and mice is compa-

 rable to the oral  doses given to mice in the NCI bioassay of PCE (386 to 1,072

 mg/kg; Table 9-4).  The ratio of the dose metabolized per animal for the rat

 versus mouse from  Table 9-5 is as follows:


                                           Amount metabolized
             Dose                           ratio:  rat/mouse

             500 mg/kg                      12.52/1.85 = 6.77


 The body weight ratios for the allometric extrapolation bases wO«67 (surface

 area) and t-jl-O (mg/kg) are:


               W0.67 = (250/24.5)°-67 = 4.74

               Wl-0  = (250/24.5) = 10.20


 Hence, the comparative metabolism of PCE for the rat and mouse is more con-

 sistent with metabolism proportional to surface area (wO-67) than to body

wei ght.

     The long half-life of PCE (6-7 hours) in the rat or man (55-65 hours)

suggests that for daily dosing (24-hour interval) metabolism continues through-

out the entire interval between successive exposures.  Therefore, 1) the daily

exposure to PCE represents chronic administration rather than repetitive dosing

in which the dose is completely cleared from the body before the next dose, as

is the  case  for other chloroethylenes with short half-lives (e.g.,  trichloro-
                                      9-34

-------
ethylene, vinylidene chloride); and 2) the "repair" mechanisms are constantly



overshadowed by metabolism.



     Buben and O'Flaherty (1985) examined the extent of PCE metabolism in male



mice during subchronic administration by gavage in corn oil vehicle for 6 weeks



(5 days/week; conditions similar to the NCI bioassay of PCE).   While the strain



of mice was Swiss-Cox and not the B6C3F1 used in the NCI bioassay, the results



in this study are probably relevant and generally applicable.   The investiga-



tors found that PCE metabolism in mice was capacity-limited and dose-dependent.



Metabolism was measured by urinary metabolite(s) excreted per  day.  TCA was



found to be the only urinary metabolite.  Oxalic acid was not  considered to



be a significant PCE metabolite, since the oxalic acid excreted in treated and



control animals was comparable.  However, labeled PCE was not  used, and hence



there was no assurance that a portion of the oxalic acid excreted was not de-



rived from PCE, as was found for the rat by Schumann et al. (1980) and for mice



by Yllner (1961).  Because urinary TCA was the sole measure of metabolism in



the Buben and O'Flaherty (1985) study, it is judged that this  study represents



somewhat of an underestimate of total metabolism, since other  metabolites,



including CO;? in exhaled air, tissue-bound metabolites (Schumann et al., 1980;



Pegg et al., 1979), and metabolite(s) in feces (< 5%; oxalic acid, etc.),  were



not taken into account.  Nonetheless, urinary TCA can be expected to represent



70% to 80% of the amount of PCE metabolized.  Daily urinary TCA tended to in-



crease in amount during each week of treatment, with leveling  (plateauing)



towards the end of the week, presumably because of the relatively long half-



life of PCE.  Urine collections (24-hour) were taken at the plateau or steady-



state metabolism.



     Figure 9-5 shows the dose-dependent relationship between  administered dose



(mg/kg/day) and excreted urinary metabolite (mg TCA/kg/day).  This hyperbolic





                                      9-35

-------
U>

cn
             O>
             D)
            O
            CD

            i<
            Ld
                            I I I I 11 I I I I[I 11 I I I I I I[I I I I I I I I I [ I I I I I I I I I [ I I 11 1111 I I I II I I 11 I I I
200
400    600
800    1000   1200

 DOSE (mg/kg)
                                                                     1400   1600
11 I I I I I I I I I I I I I I 11 I I I I I


       1800   2000
           Figure 9-5.  Relationship  between the PCE dose and the amount of total urinary metabolite

           excreted per day by mice in each group.



           SOURCE:  Buben and O'Flaherty, 1985.

-------
curve fitted well to the Michaelis-Menten equation with a Vmax of 136 mg

PCE/kg/day and a Km of 660 mg/kg.  The curve deviates from linearity above 200

mg/kg doses, indicating that even at "low" doses the metabolism of PCE is

capacity-limited in mice.  Schumann et al. (1980) and Pegg et al. (1979, Table

2) found that for a single 500 mg/kg dose to B6C3F1 mice, 75.5 mg PCE/kg was

metabolized.  From Figure 9-5, for a comparable dose 60 mg TCA/kg was observed.

On a molar equivalent basis, these two metabolized doses agree surprisingly

well (455 ymoles PCE versus 367 ymoles TCA, where 1 ymole PCE is  metabolized

to 1 pinole TCA).  This indicates that about 80% of the metabolism is repre-

sented by urinary TCA, i.e., measuring only the TCA excretion of  the mice

probably underestimates total  metabolism by about 20%.  (Milligrams  of urinary

TCA is essentially equivalent to milligrams of PCE metabolized, since the

molecular weight of TCA is 163.40 and that of PCE is 165.85.)  Finally, an

important observation can be made from the data of Buben and O'Flaherty (Figure

9-5) relevant to the NCI bioassay.  The NCI gavage averaging doses were for

male mice, 536 and 1,072 mg PCE/kg/day, and for female mice, 386  and 772 mg

PCE/kg/day.  The amount of metabolites contributing to a carcinogenic response

can be estimated as follows from Figure 9-5 (assuming that Swiss-Cox mice and

B6C3F1 mice metabolize PCE similarly):


                                       Uri nary
                    NCI              metabolite           % increase
                gavage dose         from Fig. 9-5          with dose

                 mg/kg/day          mg TCA/kg/day

Males              536                  60.95
                                                              38
                 1,072                  84.18

Females            386                  50.19
                                                              46
                   772                  73.32
                                      9-37

-------
Thus, for a 100% increase in PCE administered doses,  metabolism increases  only



38% to 46%.  This observation may account in part for the failure to observe a



dose-related tumor increase in mice in the NCI study  (Table 9-4).  The NTP



bioassay used lower doses of 50 to 200 mg/kg, and dose-related increases  of



hepatocarcinomas in mice were found.   This is consistent  with  the dose-depen-



dent findings given in Figure 9-55 since metabolism contributing to tumorigenic



response increases proportionally greater for low gavage  doses than for higher



doses.  In fact, Figure 9-5 shows nearly a direct linear  relationship  between



amount of dose metabolized and dose administered below 200 mg/kg.



     Finally, it is important to note that Buben and  O'Flaherty (1985) demon-



strated that indices of PCE hepatotoxicity (increased liver weight,  liver



triglyceride accumulation, glucose-6-phosphatase activity,  and serum SGPT



activity) were highly correlative with amount of PCE  metabolized by  mice.   The



degree of liver response, as measured by each toxicity parameter when  plotted



against total urinary metabolites, was linear in all  cases,  suggesting that



the hepatotoxicity of PCE in mice is  directly related to  the metabolism of PCE.



     9.3.2.1.2.2  Inhalation Studies.  Pegg et al.  (1979)  and  Schumann et  al.



(1980) have also determined body burdens and metabolism of Sprague-Dawley  rats



and B6C3F1 mice after inhalation exposure to 10 and 600 ppm PCE for 6  hours.



The animals were exposed to ^C-PCE and the radioactivity determined in urine,



feces,  expired air,  etc., and unchanged l^C-PCE in  expired air for 72  hours



postexposure.  The data from the Pegg et al. and Schumann et al. studies  are



collated and re-expressed for comparison of rat and mouse metabolism in Table



9-6.




     For rats exposed to 10 or 600 ppm, only 32% (0.47 mg) and 12% (9.11  mg)



respectively, of the estimated PCE uptake was metabolized.   The remaining  PCE
                                      9-38

-------
           TABLE 9-6.  DISPOSITION OF l^C-PCE RADIOACTIVITY FOR 72 HOURS AFTER INHALATION EXPOSURE FOR 6 HOURS
                                         TO SPRAGUE-DAWLEY RATS AND B6C3F1 MICE
CO
UD

Expired
unchanged
Metabol i zed
14co2
Urine
Feces
Carcass
Total
Rats (av. of 3)a
10 ppm
1.008 (68%)

0.053
0.275
0.076
0.063
0.467 (32%)
1.475

600 ppm
68.39 (88%)

0.54
4.54
2.36
1.67
9.11 (12%)
77.50
Mice (av.
of 3)
10 ppm
0.048

0.032
0.285
0.027
0.012
0.356
0.404
(12%)




(88%)

   aAverage experimental animal weights (grams): 250, rat; 24.5, mouse.

   SOURCES:  Adapted from Pegg et al., 1979 and Schumann et al., 1980.

-------
uptake, 68% and 88%, was excreted via the lungs  unchanged.   Thus,  limited  and



saturable metabolism of PCE is evident.   For mice,  the  metabolism  of PCE  can



be compared to rats only at the low exposure of  10  ppm  exposure  concentration.



The percentage of uptake metabolized was  considerably more  extensive (88%  of



the dose), and only 12% was excreted unchanged.   However, the  absolute  amount



of dose metabolized is comparable at 10  ppm between  mice  and rats.   These



results suggest that PCE metabolism is limited and  saturable in  these species



and confirm similar observations after oral  exposure.



     The data of Table 9-6 from inhalation  studies,  when  compared with  that of



Table 9-5 from oral studies, indicated that, for the rat, inhalation of PCE at



600 ppm for 6 hours resulted in 9.1 mg of compound  metabolized.  This amount  is



comparable to that of an oral  125 mg/kg  dose (12.5 mg).   Thus, inhalation  expo-



sure to 600 ppm for 6 hours is approximately equivalent to  an  oral  dose of 125



mg/kg.



     For mice, the metabolism  of PCE was  determined  only  at  10 ppm  exposure for



6 hours.  The amount metabolized, 0.36 mg,  is comparable  to  the  amount  metabo-



lized in rats, 0.47 mg.



     9.3.2.1.2.3  Inhalation Exposure in  Humans. PCE is  readily absorbed



through the lungs into the blood by first-order  passive diffusion processes.



Two major processes account for the known elimination of  PCE from the body:



1) pulmonary excretion of unchanged PCE  and 2) metabolism of PCE to urinary



metabolites.  The metabolism of PCE appears to be very  limited in humans,



as it is in experimental animals.  Only  a small  percentage  of  the estimated



amounts absorbed by inhalation are metabolized to TCA and other  chlorinated



metabolites found in the urine (see Chapter 4).   Saturation  of metabolism  has



been estimated to occur in humans at relatively  low  exposure concentrations



(between 100 and 400 ppm PCE in inspired  air--see Chapter 4).  Estimates  of the






                                      9-40

-------
extent of metabolism in humans have been made from balance studies  by measuring



urinary metabolites after accounting for a retained inhalation exposure dose



(see Chapter 4).  However, there are several  difficulties associated with



balance studies in humans.  For example, problems are encountered in obtaining



an accurate measurement of the retained dose of PCE from inhalation exposure.



Balance studies in humans are obviously incomplete.  One shortcoming in some



of the older PCE studies is the imprecision of the analytical  methodologies;



for example, the Fujiwara reaction for measuring metabolites.   Also, urinary



sampling may not have been sufficient considering the long half-life of PCE



metabolites.  In addition, individual variations among subjects can be rather



high.  Only urinary TCA was used in these studies as a measure of metabolism,



and the possibility exists that significant amounts of metabolites  other than



TCA are excreted in the urine or by other routes.  It has been proposed that a



much higher percentage of the dose is actually metabolized, and that a so-far



unrecognized pathway may exist for PCE and for other "metabolites which have  not



yet been identified.  However, measuring the amount of urinary TCA  is one



approach that may be used to compare the metabolized dose of PCE among species.



The metabolized dose may be considered to be the effective dose in  organ toxi-



city and carcinogenicity.



9.3.2.1.3  Covalent Binding.  Reactive metabolites of PCE have been shown  to



bind irreversibly to cellular macromolecules in vitro (Costa and Ivanetich,



1980), and in vivo (Pegg et al., 1979; Schumann et al., 1980).  Pegg et al.



(1979) exposed  rats by inhalation to 10 and 600 ppm l^C-PCE for 5 hours and  to



oral doses of 1 and 500 mg/kg.  For rats exposed by inhalation, hepatic irre-



versible binding of l^C-PCE occurred in a ratio of bound/total metabolized of



0.72 - 0.87; the turnover of bound radioactivity was slow as measured over 72
                                      9-41

-------
hours postexposure.  There was no significant  difference  in the  ratio between
inhalation and oral exposure,  or between  low  and  high  doses,  showing that
binding occurs proportionally  to the amount metabolized  (Tables  9-5 and  9-6).
     Schumann et al. (1980) compared the  irreversible  binding in vivo for  rats
and mice exposed to 14C-PCE by gavage (500 mg/kg) as  in the NCI  bioassay.
Their results are shown in Table 9-7.  Peak binding to liver  protein occurred
at 6 hours for mice and 6 to 12 hours for rats, with  binding  in  mice four- to
eightfold greater than in rats.  These results are in  accord  with  PCE metabo-
lism in these species  (Tables  9-5 and 9-6).   Schumann  et  al.  found no detect-
able binding of high specific activity 14C-PCE to liver DNA when mice were
exposed by inhalation  (600 ppm) or gavage (500 mg/kg). However, the limit of
detection was calculated to be 10 to 14.5 alkylations  per 106 nucleotides,  and
therefore very  low  levels of DNA binding  may  have occurred.   These investiga-
tors also chronically  administered PCE (11 daily  doses) by gavage  at increas-
ing doses to 1,000 mg/kg.   Mice, but not rats,  showed a  20%  to  25% increase
of liver/body weight ratio.  However, the DNA/g of liver  protein decreased,
indicating hypertrophy rather than hyperplasia.   Nonetheless, DNA  synthesis
(l^C-thymidine  incorporation)  increased in both mice  and  rats in a dose-related
manner, indicating activation of "repair" mechanisms.   These  investigators
interpreted their data as demonstrating cytotoxicity  (mouse more susceptible
than rat)  enhancing spontaneous incidence of  liver tumors,  rather than  to
genotoxicity.  However, the present evidence  is insufficient  to  support  one
mechanism  over another.
9.3.2.2  Calculation of Dose Metabolized  from Animal  Data
9.3.2.2.1   Method 1.   In the NCI study, the time-weighted average  gavage dose
in corn oil  was 536 and 1,072 mg/kg/day for male  mice, and 386 and 772 mg/kg/
day for female mice; or, based on the average  terminal weights,  16.1 and 32.2
                                      9-42

-------
        TABLE 9-7.   TOTAL HEPATIC MACROMOLECULAR BINDING FOLLOWING ORAL
                 EXPOSURE OF RATS AND MICE TO 500 MG/KG
Time
(hours)
1
6
12
24
48
72


A
0.9643
2.7168
1.6868
1.3862
1.0812
0.4723
ymole of
Mean
- Mi ce
+_ 0.1154
+ 0.5333
_+ 0.1119
_+ 0.2622
+_ 0.2895
+ 0.1310
PCE-bound/g
+ S.D.
B
0.1699
0.3310
0.4017
0.6120
0.3474
0.3210
hepatic protein

- Rats
_+ 0.0251
+_ 0.0271
+ 0.0860
+_ 0.0269
+ 0.0595
+_ 0.0601

A/B
5.7
8.2
4.2
2.3
3.1
1.5
SOURCE:   Schumann et al., 1980.
                                      9-43

-------
mg/male mouse/day, and 9.7 and 19.3 mg/female mouse/day (Table 9-4).  For a
comparable dose of 500 mg/kg  (12.25 mg/animal) in corn oil, virtually complete
absorption was observed  (Schumann et al., 1980; Table 9-5).  Therefore, it can
be  assumed that the PCE  doses given by gavage in the NCI study were completely
absorbed.  However, a dose-dependent portion of the oral dose is excreted via
the lungs unchanged (83% for  a 500 mg/kg dose) and only a fraction is metabo-
lized  (17%).   Since the  percent of the dose exhaled unchanged increases with
increasing dose  (Table 9-5),  and percent metabolized decreases, these values
 represent a  conservative estimate (overestimate) for any higher dose of the NCI
bioassay.  Hence,  the daily body burdens (amount of dose metabolized contribu-
ting to  carcinogenic response) for the assay mice are given in Table 9-8.
 9.3.2.2.2  Method  2.  The  estimations of PCE body burden metabolized and con-
tributing to a tumorigenic response may be further refined using the data of
Buben  and O'Flaherty (Figure  9-5) describing the relationship between gavage
dose and amount metabolized.  Two reservations should be noted:  1) an assump-
tion is made that  the relationship determined with Swiss-Cox mice is also
appropriate  for B6C3F1 mice of the NCI bioassay, and 2) the relationship under-
estimates metabolism (by about 20%) since only urinary TCA excretion was used
as  an  index  of total metabolism (see discussion in Section 9.3.2.1.2.1).
However, at  the doses of the  NCI bioassay (male mice, 536 and 1,072 mg/kg and
female mice,   386 and 772 mg/kg) urinary excretion is by far the major route of
metabolized  compound, and  therefore total urinary metabolites should be a
reasonable approximation of the amount of metabolism.  Furthermore, the exper-
imental conditions of chronic daily administration of PCE by gavage closely
mimic the conditions of the NCI bioassay.  The daily body metabolic load (amount
of dose metabolized contributed to carcinogenic response, as calculated from
the  relationships of Figure 9-5) for the NCI bioassay are given in Table 9-9.

                                      9-44

-------
  TABLE 9-8.  LIFETIME AVERAGE DAILY EXPOSURE (LAE)  FOR  B6C3F1  MICE  IN  THE  NCI
                 BIOASSAY, ASSUMING 17% OF DOSE IS  METABOLIZED
Gavage
dose
(mg/kg/day)
Males 536
1072
Females 386
772
Dose
metabol i zed
(mg/kg/day)
91.12
182.24
65.62
131.24
LAEa
(mg/kg/day)
56.41
112.82
40.62
81.24
aThe LAE of the mice is given by:

       78 weeks x 5 days x dose metabolized (mg/kg/day)
       90 weeks   7 days

SOURCE:  NCI, 1977a.
  TABLE 9-9.  LIFETIME AVERAGE DAILY EXPOSURE (LAE)  FOR B6C3F1  MICE  IN  THE  NCI
     BIOASSAY, USING A DOSE-METABOLISM (IN URINARY  EXCRETION) RELATIONSHIP
Gavage
dose
(mg/kg/day)
Male mice 536
1072
Female mice 386
772
Dose
metabol ized
(mg/kg/day)
60.95
84.18
50.19
73.32
LAEa
(mg/kg/day)
37.73
52.11
31.07
45.39
aThe LAE of the mice is given by:

       78 weeks x 5 days x dose metabolized (mg/kg/day)
       90 weeks   7 days

SOURCE:  Buben and O'Flaherty, 1985.


                                     9-45

-------
9.3.3  Choice of Risk Model



9.3.3.1  General Considerations—The unit risk estimate for PCE represents an



extrapolation below the dose range of experimental  data.  There is currently no



solid scientific basis for any mathematical  extrapolation model that relates



exposure to cancer risk at the extremely low concentrations, including the unit



concentration given above, that must be dealt with  in evaluating environmental



hazards.  For practical reasons the correspondingly low levels of risk cannot



be measured directly either by animal experiments or by epidemiologic studies.



Low-dose extrapolation must, therefore, be based on current understanding of



the  mechanisms  of carcinogenesis.  At the present time the dominant view of the



carcinogenic process involves the concept that most cancer-causing agents also



cause  irreversible damage to DNA.  This position is based in part on the fact



that a  very large proportion of agents that  cause cancer are also mutagenic.



There  is reason to expect that the quanta! response that is characteristic of



mutagenesis is  associated with a linear (at  low doses) 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.  This is particularly



true at the lower end of the dose-response curve; at high doses, there can be



an upward curvature, probably reflecting the effects of multistage processes on



the mutagenic response.  The linear 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 evalu-



ation possible  (e.g., radiation-induced leukemia, breast and thyroid cancer,



skin  cancer induced by arsenic in drinking water, liver cancer induced by



aflatoxins  in the diet).  Some supporting evidence  also exists from animal



experiments (e.g., the initiation stage of the two-stage carcinogenesis model





                                      9-46

-------
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 such a 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.



     For several reasons, the unit risk estimate based on animal bioassays



is only an approximate indication of the absolute risk in populations exposed



to known carcinogen concentrations.  First, there are important species dif-



ferences in uptake, metabolism, and organ distribution of carcinogens, as well



as species differences in target site susceptibility, immunological responses,



hormone function, dietary factors, and disease.  Second, the concept of equi-



valent doses for humans compared to animals on a mg/surface area basis is



virtually without experimental verification as regards carcinogenic response.



Finally, 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 po-



tency of a given agent as compared with other carcinogens.  Such estimates



are, of course, more reliable when the comparisons are based on studies in



which the test species, strain, sex, and routes of exposure are similar.



     The quantitative aspect of carcinogen risk assessment is addressed here



because of its possible value in the regulatory decision-making process, e.g.,



in setting regulatory priorities, evaluating the adequacy of technology-based



controls, etc.  However, the imprecision of presently available technology for



estimating cancer risks to humans at low levels of exposure should be recog-





                                      9-47

-------
 nized.   At  best,  the  linear extrapolation model used here provides a  rough  but



 plausible estimate  of the upper  limit of risk—that is, with this model  it  is



 not  likely  that the true risk would be much more than the estimated risk, but



 it could be considerably lower.  The risk estimates presented in subsequent



 sections should not be  regarded, therefore, as accurate representations  of



 the  true cancer risks even when  the exposures involved are accurately defined.



 The  estimates  presented may, however, be factored into regulatory decisions



 to the  extent  that  the  concept of upper-risk limits is found to be useful.



 9.3.3.2 Mathematical Description of the Low-Dose Extrapolation Model—The



 mathematical formulation chosen  to describe the linear nonthreshold dose-



 response relationship at low doses is the linearized multistage model.   This



 model employs  enough arbitrary constants to be able to fit almost any mono-



 tonically increasing dose-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.




     Let P(d)  represent the lifetime risk (probability)  of cancer at dose d.



 The  multistage model-has the form






                 P(d)  = 1 - exp  [-(q0 + qjd + qxd2 + .... + qkdk)J





where






                          qi  >_ 0, i  = 0,  1,  2,  ...,  k





Equi valently,






                  Pt(d)  = 1  -  exp [-(qLd  +  q2d2  +  ...  +  qkdk)J
                                      9-48

-------
where



                              pt(d)  - fW  - R(o)
is the extra risk over background rate at dose d or the effect  of  treatment.



     The point estimate of the coefficients qi , i  = 0,  1,  2,  . . . ,  k,  and  con-



sequently, 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 P-t(d) has approximately the form  P-t(d)  =  q^  x d.



Therefore, q  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



Lg be the maximum value of the log-likelihood function.   The  upper-limit,  q-^,



is calculated by increasing q-,  to a value q-i  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 Lj satisfies the  equation





                             2 (L0 - 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 P^(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
                                     9-49

-------
 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-response model, the number of terms in the polynomial

 is  chosen equal to (h-1), where h is the number of dose groups in the experi-

 ment, 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 refit 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
                               2
                              X  =
                                        -jP-j  (1-Pi)
 is  calculated where N^ is the number of animals in the ith dose group, X^  is

 the  number of animals in the i*-n dose group with a tumor response, PJ  is the

 probability of a response in the i^n 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.

 9.3.3.3  Adjustments for. Less Than Lifespan Duration of Experiment—If the

duration of experiment Le is less than the natural lifespan of the test animal

 L, the slope qls or more generally the exponent g(d), is increased by multi-

 plying a factor (L/l_e)3.   We assume that if the average dose d is continued,

the age-specific rate of cancer will  continue to increase as a constant function
                                      9-50

-------
of the background  rate.  The age-specific rates for humans increase at least by

the third power of the age and often by a considerably higher power, as demon-

strated by Doll (1971).  Thus, it is expected that the cumulative tumor rate

would increase by  at least the third power of age.  Using this fact, it is

assumed that the slope q^, or more generally the exponent g(d), would also

increase by at least the third power of age.  As a result, if the slope q^

[or g(d)] is calculated at age Le, it is expected that if the experiment  had

been continued for the full lifespan L at the given average exposure,  the slope
 -ff                                                       Q
q-^ [or g(d)] would have been increased by at least (L/L ) .

     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)J.



9.3.4  Unit Risk for a Compound

9.3.4.1  Definition of Unit Risk—This section deals with the unit risk for

PCE in air and water and the potency of PCE relative to other carcinogens that

the CAG has evaluated.  The unit risk estimate for an air or water pollutant is

defined as the incremental  lifetime cancer risk occurring in a hypothetical

population in which all individuals are exposed continuously from birth

throughout their lifetimes to a concentration of 1 pg/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.



                                      9-51

-------
      For most cases of interest to risk assessment, the 95% upper-limit risk



 associated with dose units/day as obtained by means of the GLOBAL83 computer



 program  (Howe, 1983), can be adequately approximated by P(d) = 1 exp [-(q^)].



 9.3.4.2  Unit Risk Estimates and Interpretation—The unit risk estimate based



 on  animal bioassays is only an approximate indication of the absolute risk in



 populations  exposed to known carcinogen concentrations.  There may be impor-



 tant  differences  in target site susceptibility, immunological responses, hor-



 mone  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 potency



 of  a  given agent  compared with other carcinogens.  The comparative potency of



 different agents  is more reliable when the comparison is based on studies in



 the same test species, strain, and sex.



      The quantitative aspect of carcinogen risk assessment is included here



 because  it may be  of use in the regulatory decision-making process, e.g., set-



 ting  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 extrapola-



 tion  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 esti-



mates presented in subsequent sections 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 9-10 gives the upper-limit slope



estimates,  q-^, from the linearized multistage model for PCE.  The slope





                                      9-52

-------
estimate can be used to compare the relative carcinogenic potency of PCE to



that of other potential human carcinogens.  The slope is also used to calculate



upper-bound incremental risks at low levels of exposure.



9.3.5  Alternative Low-Dose Extrapolation Models



     In addition to the multistage model, two more models, referred to as the



probit and the Weibull, are often employed by the CAG for low-dose extrapolation



for purposes of comparison.  However, calculations made with these two alter-



native models are not presented in this document, since no meaningful  compari-



sons can be made on the basis of these calculations.   For example, the 95%



upper-bound estimate of cancer risk at the metabolized dose 0.1  mg/kg/day is



predicted to be 1 by either the probit or the Weibull model, while the corre-



sponding risk estimated by the multistage model is only 0.02.





9.4  CALCULATION OF UNIT RISK



9.4.1  Potency (Slope) Estimate on the Basis of Animal  Data



     The dose metabolized, which is defined as the total uptake  minus  the amount



expired unchanged, is used as the dosimetry in the risk assessment.  The cancer



potency (slope) calculated on this basis can be converted back to the  administered



dose by using the dose-metabolism relationship estimated from experimental  data.



     Table 9-10 presents the carcinogenic potency for humans along with the



metabolized doses calculated by two approaches and tumor incidence rates in the



NCI (1977a) bioassay.  The detailed derivations of the metabolized doses are



presented in Tables 9-8 and 9-9.  To account for the high mortality rates in



the dosed groups, the animals that died before the occurrence of the first



hepatocellular carcinomas are excluded from the denominators of  tumor incidence



rates.  Another approach would be to use the time-to-death data  for risk calcu-



lation.  This approach yields a slightly higher risk value than  that calculated
                                      9-53

-------
         TABLE 9-10.   DOSE-RESPONSE  DATA  AND  POTENCY  (SLOPE) ESTIMATES
                                                          Potency  estimate0
Data base for
calculating
body burden
Method 1: Use
total metabolized
dose (Schumann
et al., 1980)




Method 2: Use
dose metabolized
in urinary
excretion (Buben
and O'Flaherty,
1985)



(animal )
(mg/kg/day)
Male mice:
0
56.41
112.82
Female mice:
0
40.62
81.24
Male mice:
0
37.73
52.11

Female mice:
0
31.07
45.39
Tumor
incidence13

2/20
32/48
27/45

0/20
19/48
19/45

2/20
32/48
27/45


0/20
19/48
19/45
(mg/kg/day )-•>-
qi(M) q*(A)

1.8x10-1 4.4x10-2



1.6x10-1 4.0x10-2



3.4x10-1 6.8x10-2




2.5x10-1 5.1x10-2

aSee Tables 9-8 and 9-9 for the animal  lifetime  average  exposures  (LAE).
bThe denominators are the number of animals  that  survived  at the time the  first
 hepatocellular carcinoma occurred in  each  study.   The first tumor occurred at
 2£ weeks for the male mice and at 41  weeks  for  the female mice.
cqi(M) is the potency estimate expressed in  terms  of the metabolized dose.
 q*(A) is the potency estimate expressed in  terms  of the administered dose.
                                      9-54

-------
on the basis of incidence data with a modified denominator.  However,  the risk



calculated from the time-to-death data is difficult to interpret because the



majority of animals with tumors were discovered at the terminal  sacrifice.



Therefore, only the potency estimates calculated on the basis of incidence  data



are presented.



     Two sets of potency estimates are presented in the last two columns of



Table 9-10, using either the total dose metabolized (Method 1) or the  dose  meta-



bolized and eliminated in urine (Method 2).  These estimates are calculated by



the linearized multistage model, using the dose-response data in Table 9-10.



The values q^(M) and q^(A) are potency estimates expressed, respectively,



in terms of the metabolized and the administered doses.



     The carcinogenic potency q^(M) for humans is calculated by  multiplying



the animal potency q-^(M) by a factor (W^/W^)1/ , where WH and W^ are,  respec-



tively, the body weights for humans and animals.  The underlying assumption for



this adjustment is that the metabolized dose per surface area is equally effec-



tive in inducing cancer among species.  To convert from q^(M) to q-^(A),  the



relationship between administered dose (A) and the amount metabolized  (M) is



utilized.  For the estimates calculated according to Method 2, the relationship



M = 0.20A is used.  This relationship can be obtained from the dose-metabolism



relationship derived by Buben and O'Flaherty (1985) when the administered dose



A is small.  For the estimates calculated by Method 1, the relationship  M =



0.25A is used, as in Buben and O'Flaherty (1985), and considering the  fact  that



the urinary metabolites account for about 80% of the total metabolites.   It is



worthwhile to note that these relationships are consistent with  the empirical



observations of Pegg et al. (1979), in which approximately 29% of the  dose  was



metabolized after rats were given a small dose (1 mg/kg or 0.25 mg per animal).



     Because the body burdens calculated by Methods 1 and 2 result in  compa-





                                      9-55

-------
rable potency  estimates,  and  because  the  available  human body burden via inha-



lation is given in  terms  of urinary metabolites,  it  is  recommended that the



potency estimates  on the  basis  of  Method  2  be used  to  represent the carcino-



genic potency  of PCE.  Since  more  reliable  dose-response data exist for female



mice than for  male  mice,  the  carcinogenic potency calculated from these data is



used.  The slope,  q^, estimate  for PCE  is:





                       qj(A) - 5.1 x ID'2  mg/kg/day





expressed in terms  of the administered  dose, and





                         q£(M)  - 2.5  x  10"1 mg/kg/day





expressed in terms  of the metabolized dose  in urinary  excretion.





9.4.2  Risk Associated with 1 ug/L of PCE in Drinking  Water



     The daily PCE  intake (mg/kg/day) for a 70 kg person consuming 2 L of



water contaminated  with 1 Mg/L  of  PCE is:





                   d = (1 ug/L) x  (2  L/day) x (1Q-3  mg/pg)/70 kg



                             =  2.9 x  1CT5 mg/kg/day





Therefore, the upper-bound estimate of  the  incremental  lifetime risk due to con



suming water contaminated with  1 pg/L of  PCE is:





                    P = 5.1 x ID'2 x  2.9  x  ID"5 = 1.5  x 1CT6





9.4.3  Risk Associated with 1 U9/m3 of  PCE  in Air



     In order  to estimate the risk due  to 1 yg/m3 of PCE in air, it is neces-



sary to estimate the amount metabolized that accompanies exposure to this con-



centration in  air.   For PCE,  the body burden can  be  estimated from both human
                                      9-56

-------
and animal data.  Since only the metabolized dose in urine is  available  for



humans, the potency q^(M) = 2.5 x 10'1 mg/kg/day  in terms  of  urinary  metabo-



lites will be used in the risk calculation.



9.4.3.1  Unit Risk Calculated on the Basis of Body Burden  Derived  from Human



Data — In an experimental human exposure to PCE vapor,  Fernandez  et  al.  (1976)



observed that in 72 hours, the quantity of TCA metabolized and eliminated  in



urine was about 25 mg following 150 ppm exposure  for 8 hours.  Assuming  that



the amount metabolized in urine is linearly proportional  both  to the  air



concentration and to the duration of exposure, the observations  of  Fernandez et



al. (1976) imply that a continuous (24 hours) exposure to  1 ppm  of  PCE produce"



25 x 3/150 = 0.5 mg/day of urinary metabolites.  Since 1  ppm  = 6,908  yg/m3, a



continuous exposure to 1 yg/m3 of PCE yields a metabolized dose  of  7.3 x 10~5



mg/day.  For chronic and continuous exposures, the metabolized dose would  be



7.3 x 1Q-5/70 = 1.0 x 10~6 mg/kg/day.  Therefore, the  incremental  lifetime risk



due to 1 yg/m3 of PCE in air is:





                    P = 2.5 x 10-1 x i.o x 10-6 = 2.5  x 10~7





     A study by Bolanowska and Golacka (1972) also provides some information



on the relationship between the air concentration and  the  amount metabolized



in urinary excretion.  In this study, five subjects were  exposed to 390,000



yg/m3 of PCE (approximately 50 ppm) for 6 hours.   The  concentration of meta-



bolites in urine (mg/hr) is presented graphically over the 20-hour  period



during and after exposure.  The total amount of metabolites in urine  (area



under curve) is estimated to be about 13 mg.  The area under  the curve from  20



hours to infinity is calculated by C x ti/2/0.693, where  C is the  concentration



at the last sampling time (i.e., 20 hours) and ty?, the  half-life, is  assumed



to be 100 hours.  Under the assumption that the amount metabolized is linearly





                                      9-57

-------
proportional  to the air concentration  and  the  duration  of  exposure,  the  amount

metabolized,  associated with  1  pg/m3 of  PCE  in air,  is


           (13 mg/390,000)  x  (24 hours/6 hours)  -  1.33  x  1Q-4  mg/day


Equivalently, the amount metabolized due to  continuous  exposure  to  1 yg/m3 of

PCE in air would be 1.33 x  10'4 mg/day for a 70-kg man, which  would  be 1.9 x

ID'6 mg/kg/day.  Thus, the  cancer risk due to  1  yg/m3 of  PCE in  air  is esti-

mated to be


                   P - 2.5  x  10-1 x 1.9  x  ID'6 = 4.8 x  10"7


     This risk estimate is  about two times higher  than  that  calculated on  the

basis of Fernandez et al. (1976).  Since the study by Bolanowska and Golacka

(1972) was conducted at lower concentrations,  the  risk  calculation  on the

basis of this study is preferred.  It  is therefore recommended that  4.8 x  10~?

be used as the unit risk estimate for PCE  in air.

9.4.3.2  Unit Risk Calculated on the Basis of  Body Burden  Derived from Animal

Data—Table  9-6 contains some animal data  that can be used to  infer  human

metabolized  dose when exposed to 1 yg/m3 of  PCE in air.  The dispositions  of

14C-PCE radioactivity in rats and mice 72  hours after exposure to 10 ppm of PCE

in air for a period of 6 hours  are as follows:


                                  Rats                      Mice
                               (mg/animal)                (mg/animal)


Expired unchanged                 1.008                      0.048

Total metabolized                 0.467 (32%)                 0.356  (88%)

Metabolites  in urinary
  excretion                       0.275 (19%)                 0.285  (71%)

Total uptake                      1.475                      0.404


                                      9-58

-------
     These data indicate that the percentage of PCE that  is  metabolized  de-



creases with an increase of body weight and that the total  uptake,  not the



amount metabolized, is approximately proportional  to the  body  surface area



between the two species.  To estimate the amount metabolized in  human urinary



excretion after PCE exposure, it is assumed that the total  uptake  is propor-



tional to the body surface area among species,  including  humans, and the



relationship between the percentage metabolized (M) and the  body weight  (W)



follows the equation M = AWK.  The parameters A and K are estimated to be



A = 0.146 and K = 0.420, using the animal data  presented  above.  Thus, for a



70-kg person, the percentage metabolized is predicted to  be  2.5%.   This  is



consistent with the empirical observation in humans (Monster,  1979).  On this



basis, if humans were exposed to 10 ppm of PCE  in  air, as were the  animals,



the amount metabolized would be





                     1.48 x  (70/0.25)2/3 x 0.025 = 1.58 mg





where 1.48 x (70/0.25)2/3 js the estimated total uptake using  the  surface area



correction factor (70/0.25)2/3.



     Under the assumption that the amount metabolized (in urine) is linearly



proportional to both the air concentration and  the duration  of exposure, the



amount metabolized due to a continuous (24-hour) exposure to 1 ppm  of PCE in



air is 1.58 x 4/10 = 0.63 mg/day.  Equivalently. the body metabolic load is 0.63/



6,908 = 9.2 x 10~5 mg for a 24-hour exposure when  the air concentration  is



1 yg/m3, using the fact that 1 ppm = 6,908 ^g/m3.   For chronic and  continuous



exposure, the amount metabolized per day due to 1  yg/m3 of  PCE in  air is 9.2



x ID'5 mg/70 kg/day = 1.3 x 10'6 mg/kg/day.  Therefore, the  incremental  lifetime



risk due to 1 pg/m3 of PCE is:
                                      9-59

-------
                      P = 2.5x10-1 x 1.3x10-6 = 3.3x10-7


This estimate is about the same as that calculated on the basis of direct ob-

servations regarding human metabolized dose.  Since data based on direct obser-

vation of humans is preferred to the animal  data,  it is recommended that the

value 4.8 x 10~7, calculated on the basis of actual human metabolized dose, be

used as the unit risk estimate for PCE in air.

9.4.3.3  Unit Risk Calculated on the Basis of a Negative Inhalation Study--

For comparison, a negative inhalation study  on  rats, sponsored by the Dow

Chemical Company (Rampy et al., 1978) is used to calculate an upper-bound

estimate of risk for PCE.  Ninety-four animals  of  each sex were exposed to

either 300 ppm or 600 ppm of PCE, 6 hours/day,  5 days/week for 52 weeks.

Animals were allowed to survive until the terminal sacrifice at 31 months.

A shortcoming of this study was that the animals were exposed to PCE for only

52 weeks.  The use of data from such a study may seriously underestimate the

risk.  Among the four PCE-exposed groups, the data from the female high-dose

group yield the smallest upper-bound estimate of risk, and thus will be used

to calculate the unit risk for PCE in air.  Forty-nine female rats in the high-

dose (600 ppm) group survived beyond 2 years.  The 95% upper-bound estimate of

the probability of cancer among these 49 animals with no cancer response is Pu

= 0.059.   The corresponding lifetime average exposure is
           d = 600 ppm x 5 days  x  6 h°urs  x   52 weeks   = 53.57  ppm
                         7 days    24 hours    104 weeks
or
                      d - 370,071 yg/m3
                                      9-60

-------
Thus, the carcinogenic potency of PCE, defined as the slope in  the  one-hit



model, is calculated to be








                       b = [-ln(l-Pu)]/d = 1.6 x 10-7/yg/m3



for animals.



     To calculate the risk to humans associated with 1 yg/m3 of PCE  in  air,



it is necessary to determine the human exposure concentration C(yg/m3)



equivalent to 1 yg/m3 of PCE in air for rats.   Assuming that dose in  rela-



tion to surface area is equally effective among species,  the human  exposure



concentration C(yg/m3) satisfies the equation





                       C(yg/m3) x (20 m3/day)/(70 kg)2/3





                     = 1 yg/m3 x (0.224 m3/day)/(0.35 kg)2/3





where 20 m3/day and 0.224 m3/day are the daily air intakes,  respectively, for



a 70-kg person and an 0.35-kg rat.   Thus, C(yg/m3) = 0.39 yg/m3.  Therefore,



the human risk due to 1 yg/m3 of PCE in air is





                        P = 1.6 x 10-7/0.39 -  4.1 x 10~7





This risk estimate is again comparable to those calculated previously.



9.4.3.4  Discussion



     To the extent possible, the available metabolism and pharmacokinetic



information for PCE has been used in the risk  calculations.   Nevertheless, the



use of pharmacokinetic information  may reduce  some of the ambiguity  associated



with three major areas of risk assessment uncertainty listed below:



     1.  Extrapolation from high dose-response in animals to low dose-response
                                      9-61

-------
         in animals (i.e., low-dose extrapolation),



     2.  Extrapolation from low dose-response in animals to low dose-response



         in humans (i.e., species conversion), and



     3.  Extrapolation from gavage to inhalation (i.e., route-to-route



         extrapolation).



     In calculating the dose-response relationship for PCE, 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 receptor 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 in the



 high-dose region.



     To extrapolate from animals to humans, the metabolized dose per body sur-



 face area is assumed to be equivalent (i.e., equally potent) among species.



 This assumption is by no means supported by the empirical data.  An alternative



 approach for animal-to-human extrapolation would be to assume that mg metabolized



 dose/kg/day is equivalent among species.  If this assumption is made, the



 slope factor ql5 as calculated for PCE and expressed in terms of (mg/kg/day)-1,



would be reduced by approximately 10 times.



     The carcinogenic potency calculated from the gavage study and on the basis



of metabolized dose does provide a better basis for extrapolating the risk



estimate from gavage  to inhalation.  This is so because the relationship between



the ambient  air concentration and the amount metabolized for PCE is available
                                      9-62

-------
for both humans and animals.  Therefore, the uncertainty due to route-to-route

extrapolation should be greatly reduced by the use of metabolized dose.


9.5  RISK PREDICTION AGAINST HUMAN EXPERIENCE

     Although there is no sound epidemiologic data that could be used to esti-

mate the carcinogenic potential of PCE, efforts are made herein to determine

whether or not the risk estimate of 4.8 x 10~? per yg/m3 is  reasonable in

view of the fact that none of the epidemiologic studies on  dry-cleaners  showed

significant elevation of cancer risk,,  The major human experience in  PCE expo-

sure has been on the part of dry-cleaning workers.  Among them,  the machine

operators are the workers most heavily exposed, with a time-weighted  average

concentration of about 22 ppm.  For other dry-cleaners, the  mean exposure

concentration is approximately 3 ppm (Ludwig et al., 1983).

     According to Campbell et al . (1980), the number of persons  employed as

machine operators in commercial, industrial, and coin-operated dry-cleaning

facilities in the United States is estimated to be 27,700.   A question to  be

asked is whether a significant increase of cancer deaths should  be expected

among dry-cleaning workers, and in particular, among the machine operators.

As demonstrated below, a cohort study on dry-cleaning workers is unlikely  to

show an elevation of cancer risk.

     At an ambient air concentration of 22 ppm (for machine  operators) and 3

ppm (for other dry-cleaners), workers who work 8 hours/day,  250  days/year for

one-third of their lifetime would have an average daily exposure of
              d = 22 ppm x 6908 (Mg/m3)/ppm x  8 hours  x 25°  days  x I
                                              24 hours    365  days    3

                = 11,565.90 ug/m3
                                      9-63

-------
for machine operators  and,  similarly,





                                d  =  1577.17  Mg/m3






for other dry-cleaners.   Therefore,  the  incremental  lifetime  risk  is





                     P = 4.8 x lO"7  x  11,565.9  =  5.6 x  10'3






for machine operators, and





                     P = 4.8 x 10-7  x  1577.17  = 7.6  x 1(H






for other dry-cleaners.



     The incremental lifetime risks  for  machine operators  (5.6 x 10"3)  and



other workers (7.6 x ICT4) are not likely to be detected by  a cohort  study



because of the large sample sizes  required.   The  sample sizes required  to



show a significant elevation of cancer risk  at  a  level  of  0.05 (one-sided test)



and a power of 0.80 are as follows,  using the formula presented in Schlesselman



(1974).








           Background              	Sample  size  required
1 ifetime
cancer rates
0.01
0.05
0.10
Machine operators
5,000
25,000
36,263
Other workers
218,000
509,000
1,705,000
     The sample size required depends on the background cancer rate.  Since



no specific cancers are known to be associated with PCE exposure, three hypo-



thetical background rates are used in the calculations, which should cover most



of the  rates for specific cancers in the U.S. population.





                                      9-64

-------
9.6  COMPARISON OF POTENCY WITH OTHER COMPOUNDS


     One of the uses of the concept of unit risk is to compare the relative


potencies of carcinogens.  Figure 9-6 is a histogram representing the frequency


distribution of potency indices for 54 suspect carcinogens  evaluated  by  the


CAG.  The actual data summarized by the histogram are presented in Table 9-11.


The potency index is derived from q-^, the 95% upper bound of  the linear  compo-


nent in the multistage model, and is expressed in terms  of  (mmol/kg/day)~1.


Where human data are available for an agent,  they have been used to calculate


the index.  Where no human data are available, animal oral  studies are used in


preference to animal inhalation studies, since oral studies have constituted


the majority of animal studies.


     Based on available data concerning hepatocellular carcinomas in  female


mice, the potency index for PCE has been calculated as 8 x  10^.  This figure is

                                  *            o
derived by multiplying the slope q-j_ = 5.1 x 10   (mg/kg/day)  and the  molecular


weight of PCE, 165.8.  This places the potency index for PCE  in the fourth


quartile of the 54 suspect carcinogens evaluated by the  CAG.


     The ranking of relative potency indices  is subject  to  the uncertainties


involved in comparing a number of potency estimates for  different chemicals


based on varying routes of exposure in different species, by  means of data


from studies whose quality varies widely.  All of the indices presented  are


based on estimates of low-dose risk, using a  low-dose linear  extrapolation


model.  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 may be more appropriate.
                                      9-65

-------
                          4th
                       QUARTILE
                             3rd
                           QUARTILE
                                    2nd
                                 QUARTILE
   1st
QUARTILE
                        10'
                                                         2 x 1CT3
  -1
0
123456
LOG  OF POTENCY  INDEX
Figure 9-6.  Histogram representing the frequency distribution  of
the potency indices of 54 suspect carcinogens evaluated  by  the
Carcinogen Assessment Group.
                              9-66

-------
         TABLE 9-11.  RELATIVE CARCINOGENIC POTENCIES AMONG 54 CHEMICALS EVALUATED BY THE CARCINOGEN ASSESSMENT GROUP

                                                 AS SUSPECT HUMAN CARCINOGENS
UJ>
i
Level
of evidence3
Compounds
Aery! onitrile
Aflatoxin B1
Al drin
Ally"! chloride
Arsenic
B[a]P
Benzene
Benzidene
Beryl! ium
1,3-Butadiene
Cadmi urn
Carbon tetrachloride
Chlordane
CAS Number
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
Humans
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
Groupi ng
based on
IARC
criteria
2A
2A
2B

1
2B
1
1
2A
2B
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
Molecul ar
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
! __.__!• _ — — i 	 -
Potency
index0
lxlO+1
9xlO+5
4x1 0+3
9x10-!
2xlO+3
3xlO+3
2x10°
4xlO+4
2X10+1
5x10°
7xlO+2
2xlO+1
7x10+2
~ 	 TT 	 f _ 1 ^ . 	 ~
Order of
magnitude
(log10
index)
+1
+6
+4
0
+3
+3
0
+5
+1
+1
+3
+1
+3

-------
TABLE 9-11.   (continued)
Compounds
Chlorinated ethanes
1,2-Dichloroethane
hexachloroethane
Level
of evidence3
CAS Number Humans

107-06-2 I
67-72-1 I
1,1,2,2-Tetrachloroethane 79-34-5 I
1,1,2-Trichloroethane
Chloroform
Chromium VI
f DDT
01
00
Dichlorobenzidine
1,1-Dichloroethylene
(Vinylidene chloride)
Dichloromethane
(Methylene chloride)
Dieldrin
2,4-Dinitrotoluene
Diphenylhydrazine
Epichlorohydrin
Bis(2-chloroethyl)ether
79-00-5 I
67-66-3 I
7440-47-3 S
50-29-3 I

91-94-1 I
75-35-4 I

75-09-2 I

60-57-1 I
121-14-2 I
122-66-7 I
106-89-8 I
111-44-4 I
Animal s

S
L
L
L
S
S
S

s
L

L

S
S
S
S
S
Grouping
based on
IARC
criteri a

2B
3
3
3
2B
1
2B

2B
3

3

2B
2B
2B
2B
2B
SI opeb
(mg/kg/day)-1

9.1xlO-2
1.42xlO-2
0.20
5.73x10-2
7x10-2
41(W)
0.34

1.69
1.16(1)

6.3xlO-4(I)

30.4
0.31
0.77
9.9xlO-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°
3xlO+1
8x10°
8x10°
4xlO+3
1x10+2

4x10+2
1x10+2

5x10-2

1x10+4
6xlO+1
lxlO+2
9x10-!
2x10+2
Order of
magnitude
(Iog10 '
index)

+1
0
+ 1
+ 1
+ 1
+4
+2

+3
+2

-1

+4
+2
+2
0
+2
                                    (continued  on  the  following page)

-------
TABLE 9-11.  (continued)
Compounds
Bis(chloromethyl )ether
Ethylene dibromide (EDB)
Ethylene oxide
Heptachlor
Hexachlorobenzene
Hexachlorobutadiene
Hexachlorocyclohexane
technical grade
alpha isomer
beta isomer
gamma isomer
Hexachlorodibenzodioxi n
Nickel
Nitros amines
Dimethyl nitrosamine
Diethylnitrosamine
Dibutylnitrosamine
N-nitrosopyrrolidine
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
of
Level
evidence3
Humans Animals
S
I
L
I
I
I


I
I
I
I
L

I
I
I
I
I
S
S
S
S
S
L


S
L
L
S
S

S
S
S
S
S
Groupi ng
based on
IARC
criteria
1
2B
2A
2B
2B
3


2B
3
2B
2B
2A

28
2B
2B
2B
2B
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
6.2x10+3
1.15(W)

25.9(not by
43.5(not by
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

qf) 74.1
qf) 102.1
158.2
100.2
117.1
Potency
indexc
1x10+6
8x10+3
2xlO+1
1x10+3
5x10+2
2xlO+1

1x10+3
3x10+3
5x10+2
4x10+2
2xlO+6
7xlO+1

2x10+3
4xlO+3
9x10+2
2x10+2
4x10+3
i: u^. £~^i i _. .
Order of
magnitude
(Iog10
index)
+6
+4
+ 1
+3
+3
+1

+3
+ 3
+ 3
+ 3
+6
+2

+3
+4
+ 3
+ 2
+4

-------
                                                TABLE 9-11.   (continued)
Compounds CAS Number
N-nitroso-N-methyl urea
N-nitroso-diphenyl amine
PCBs
Phenol s
2,4,6-Trichl orophenol
Tetrachl orodibenzo-
^ p-dioxin (TCDD)
o
Tetrachl oroethyl ene
Toxaphene
Trichloroethylene
Vinyl chloride
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
Level
of evidencea
Humans
I
I
I
I
I
I
I
I
S
Animal s
S
S
S
S
S
L
S
L/S
S
Grouping
based on
IARC
criteria
2B
2B
2B
2B
2B
3
2B
3/2B
1
SI opeb
(mg/kg/day)"1
302.6
4.92xlO-3
4.34
1.99x10-2
1.56xlO+5
5.1x10-2
1.13
1.1x10-2
1.75x10-2(1)
Molecul ar
weight
103.1
198
324
197.4
322
165.8
414
131.4
62.5
Potency
indexc
3x1 0+4
1x10°
lxlO+3
4x10°
5xlO+7
8x10°
5x10+2
1x10°
1x10°
Order of
magnitude
(Iog10
index)
+4
0
+3
+1
+8
+1
+3
0
0
dS = 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
 (nig/kg/day)-! by the molecular weight of the compound.

-------
9.7  SUMMARY



9.7.1  Qualitative



     Tetrachloroethylene (PCE) induced a statistically significant  increase  in



the incidence of hepatocellular carcinomas in both high- and low-dose  male and



female B6C3F1 mice when administered by gavage for a period of 78 weeks.  The



PCE used was over 99% pure, but was estimated to contain epichlorohydrin  con-



centrations of less than 500 ppm.  It is unlikely that this response could be



attributed to the low concentration of epichlorohydrin.



     No carcinogenic effects were observed in lifetime studies of Osborne-



Mendel rats given PCE by gavage or in Sprague-Dawley rats exposed via  inhala-



tion for 12 months followed by 12 months of observation.  However,  because of



excessive dose-related mortality in the gavage experiment, and because of the



low dose level in the inhalation study, no conclusions can be made  about  the



carcinogenicity of PCE in rats.



     Intraperitoneal injection of PCE in strain A mice induced no statistically



significant incidence of pulmonary adenomas.  In mouse skin initiation-promo-



tion experiments, PCE did not initiate skin tumors, nor did it induce  skin



tumors when applied alone three times per week for the lifetime of  the animals.



However, because of inherent limitations in these assays, these negative



results do not detract from the positive findings of the National Cancer



Institute (NCI) mouse experiment.



      In a cohort study of dry-cleaning workers exposed to PCE, it was  found



that workers were at an elevated risk of colon cancer mortality; however,



as many as one half of these workers may have been exposed to petroleum dis-



tillates in their working history.  Other epidemiologic studies that found



an association between cancer risk and employment in the dry-cleaning industry,



did  not  identify which dry-cleaning solvents the employees had been exposed to.





                                      9-71

-------
9.7.2  Quantitative



     Data on hepatocellular carcinomas  in  male  and  female  mice  from an  NCI



(1977a) gavage study on PCE have been  used to calculate  the  unit  risk for



water and air.  The unit risk is defined as the increased  lifetime  cancer  risk



due to exposure to one unit of dose in  drinking water  (1 yg/L)  or in air



(1 yg/m3).  Evaluation of the metabolism data in mice,  rats,  and  humans shows



that common principal end metabolites  have been identified and  that there  is



no comparative experimental evidence that  the metabolic  pathways  for PCE



qualitatively differ for these species.   The use of relevant  metabolism and



pharmacokinetic data in the development  of unit risk calculations helps to



minimize some of the uncertainty associated with route to  route extrapolation.



     The upper-bound estimate of the incremental  cancer  risk  due  to 1 yg/L



of PCE in drinking water is 1.5 x 10~6.   The upper-bound estimate of the



incremental cancer risk due to 1 yg/m3  of  PCE in air is  4.8  x 10~7. The



fact that the weight-of-evidence for the carcinogenicity of  PCE is  limited



should be taken into consideration in the  use of these estimates.



     None of the epidemiologic studies  reviewed in  this  report  are  adequate



for use in estimatin-g unit risks.  If PCE  does  cause cancer  in  humans,  it  is



unlikely that a cohort study on dry cleaners would  reveal  this  fact, since the



incremental lifetime risk to this limited  number of workers  is  too  small to be



detected.



     The potency index for PCE is 8 x  10°,  ranking  it  in the  lowest quartile



of the 54 chemicals that the CAG has evaluated  as suspect  carcinogens.  This



potency index is calculated on the basis of oral  exposures.
                                      9-72

-------
9.8  CONCLUSIONS



     PCE has induced malignant tumors of the liver in both  male  and  female



B6C3F1 mice.  This constitutes a signal  that PCE might be a carcinogen  for



humans.  The technical adequacy and the  strong nature of the positive  response



in the 1977 NCI bioassay makes it likely that a repeat bioassay  would  also  be



positive.  In fact, the recent NTP study, currently  under audit,  showed  similar



positive results.  This study, if validated, would influence the  evidence for



carcinogenicity of PCE.



     According to the Agency's Proposed  Guidelines for Carcinogen Risk Assess-



ment (U.S. EPA, 1984), the evidence for  PCE carcinogenicity in animals is



limited, and the epidemiologic data is inconclusive.   The overall weight-of-



evidence for the human carcinogenic potential  of PCE  is Group C,  i.e., a



possible human carcinogen.



     According to a literal interpretation of the criteria  of the International



Agency for Research on Cancer (IARC), the animal  data supporting  the carcino-



genicity of PCE would be classified as limited.  Also, since existing epidemi-



ologic data for PCE are inconclusive, its overall IARC ranking would be



classified as Group 3, meaning that the  data are inadequate, and  PCE cannot be



classified as to its human carcinogenicity.



     While the weight of evidence is limited for PCE  carcinogenicity (a



possible human carcinogen by proposed EPA criteria and inadequate data  for



assessing carcinogenicity by IARC criteria), an upper-bound incremental  unit



risk has been estimated for both ingestion and inhalation exposure.  This has



been done to help answer the "what-if" question—if PCE is  carcinogenic  in



humans what is the possible impact upon  public health.  The upper-bound



incremental unit risk for inhalation of  1 pg PCE/m^ is 4.8  x 10~? and  the
                                      9-73

-------
unit risk for ingestion  of  1 ^g  of  PCE/L in drinking water is 1.5 x 10"^.



The potency index for PCE,  based on  NCI gavage data, places it in the lowest



quartile of the 54 chemicals that the CAG has evaluated as suspect or known



carci nogens.
                                     9-74

-------
 9.9  REFERENCES
 Blair,  A.;  Decoufle,  P=; Grauman, D.   (1979)  Causes of death  among  laundry
      and  dry-cleaning workers.  Am. J. Pub. Health 69:508-511.

 Bolanowska, W.; Golacka, J.  (1972)  Absorption and elimination of tetra-
      chloroethylene in humans under experimental conditions.   Medycyna  Pracy
      23(2):109-119.  English translation.

 Bolt, H.M.; Laib, R.J.; Filser, J.G.   (1982)  Reactive metabolites and  carci-
      nogenicity of halogenated ethylenes.  Pharmacology 31(1):l-4.

 Buben, J.R.; O'Flaherty, E.J   (1985)  Delineation of the role of metabolism
      in the hepatotoxicity of trichloroethylene and perchloroethylene:  a dose-
      effect study.  Toxicol. Appl. Pharmacol. 78:105-122.

 Campbell, G.L.; Cohan, D.; North, D.W.  (1980)  The application of decision
      analysis to toxic substances: proposed methodology and two case studies.
      Contract No. 69-01-6054.  Prepared for the U.S.  Environmental Protection
      Agency, Office of Toxic Substances, Washington,  D.C.

 Costa, A.K.; Ivanetich, K.M.  (1980)  Tetrachloroethylene metabolism by  the
      hepatic microsomal cytochrome P-450 system.  Biochem. Pharmacol. 29:2863-
      2869.

 Cox,  C.B.  (1972)  Regression model  and life tables.   J.  Roy.  Stat.  Soc. B
      34:187-220.

 Crump, K.S.; Guess, H.A.; Deal, L.L.  (1977)  Confidence  intervals and test
      of hypotheses concerning dose-response relations  inferred  from  animal
      carcinogenicity data.  Biometrics 33:437-451.

 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)  The biological  basis for  extrapolation of  dose-
      carcinogenic response relationships of laboratory  animals  to  man:
      a review paper.  Prepared for the Carcinogen  Assessment  Group,  U.S.
      Environmental Protection Agency.

 Dedrick, R.L.   (1973)   Animal scale-up.  J. Pharmacokinet. Biopharm.  1:435-461.

 Dedrick, R.L.;  Bischoff, K.B.  (1980)  Species similarities  in  pharmacokine-
     tics.  Fed. Proc. 39:54-59.

 Doll, R.  (1971)  Weibull  distribution of cancer:  implications  for models of
     carcinogenesis.  J. Roy. Stat.  Soc. A 13:133-166.

Duh,  R.W.; Asal, N.R.   (1984)  Mortality among laundry  and dry-cleaning
     workers in Oklahoma.   Am.  J.  Public Health  74(11):1278-1280.
                                     9-75

-------
Fernandez, T.; Guberman,  E.;  Caperos,  J.   (1976)   Experimental  human exposures
     to tetrachloroethylene vapor and  elimination  in  breath  after inhalation.
     Am. Ind. Hyg. Assn.  J. (March):143-150.

Filser, J.G.; Bolt, H.M.   (1979)   Pharmacokinetics  of halogenated ethylenes
     in rats.  Arch. Toxicol.  42:123-136.

Freireich, E.J.; Gehan, E.A.;  Rail,  D.P.;  Schmidt,  L.H.;  Skipper, H.E.   (1966)
     Quantitative comparison  of toxicity  of  anticancer agents  in  mouse,  rat,
     hamster, dog, monkey,  and man.   Cancer  Chemother. Rep.  50(4):219-244.

Gehring, P.J.; Watanabe,  P.G.; Blau,  G.E.   (1980)   Risk assessment of environ-
     mental carcinogens utilizing pharmacokinetic  parameters.   Ann.  NY  Acad.
     Sci. 329:137-152.

Gillette, J.R.  (1976)  Application  of pharmacokinetic principles in extrapola-
     tion of animal data  to humans.   Clin. Toxicol.  9:709-722.

Howe, R.B.   (1983)  GLOBAL83:  an  experimental  program, developed for  the  U.S.
     Environmental Protection  Agency  as an update  to  GLOBAL82:  a  computer
     program to extrapolate quantal  animal toxicity  data  to  low doses (May,
     1982).  K.S. Crump and Co.,  Inc., Ruston,  LA.   Unpublished.

Innes, J.R.M.; Ulland, B.M.;  Valerio,  M.G.;  Petrocelli, L.;  Fishbein, L.;
     Hart, E.R.; Pollota, A.J.; Bates, R.R.;  Falk,  H.L.;  Gart,  J.J.; Klein, M.;
     Mitchell, I.; Peters,  J.   (1969)   Bioassay of  pesticides  and industrial
     chemicals for tumorigenicity in  mice: a  preliminary  note.  J. Natl.
     Cancer Inst. 42:1101-1114.

Kaplan, S.D.  (1980)  Dry-cleaners workers exposed  to perchloroethylene:  a
     retrospective cohort mortality  study.   Contract  No.  210-77-0094.
     National Institute for Occupational  Safety and  Health,  Washington,  D.C.

Katz, R.M.; Jowett, D.  (1981)  Female laundry  and  dry-cleaning workers  in
     Wisconsin:  a mo-rtlity  analysis.   Am.  J.  Pub.  Health  71:305-307.

Klassen, C.D.; Plaa, G.L.  (1967) The relative effects of various chlori-
     nated hydrocarbons on  liver  and  kidney  function  in mice.   Toxicol.  Appl.
     Pharmacol.  9:129-131.

Krasovski, G.N.   (1976)  Extrapolation of  experimental  data  from  animal  to  man.
     Environ. Health Perspect. 13:51-58.

Laskin, S.; Sellakumar, A.R.;  Kuschner, M.;  Nelson,  N.; La Mendola,  S.;
     Rusch, G.M.; Katz, G.V.;  Dulak,  N.C.; Albert,  R.E.  (1980)  Inhalation
     carcinogenicity of epichlorohydrin in non-inbred Sprague-Dawley rats.
     J. Natl. Cancer Inst.  65(4):751-755.

Lin, R.S.; Kessler, I.I.   (1981)   A multifactorial  model  for pancreatic
     cancer in man.  J. Am. Med.  Assoc. 245:147-152.

Ludwig, H.; Meister,  M.;  Roberts, D.;  Cox, C.   (1983)   Workers' exposure  to
     perchloroethylene in the  commercial  dry-cleaning industry.  Am. Ind. Hyq
     Assoc. J. 44(8):600-605.

                                      9-76

-------
Miettinen, O.S.; Wang, J.D.  (1981)  An alternative to the proportionale moral-
     ity ration.  Am. J. Epidenriol.  114(1) :144-148.

Monster, A.  (1979)  Differences in uptake, elimination,  and metabolism in
     exposure to trichloroethylene, 1,1,1-trichlorethane, and tetrachloro-
     ethylene.   Int. Arch. Occup. Environ. Health 42:311-317.

National Cancer Institute  (NCI).  (1976)  Report on carcinogenesis  bioassay  on
     chloroform.  National Technical Information Service, Springfield,  VA.
     PB-264018.

National Cancer Institute  (NCI).  (1977a)  Bioassay of tetrachloroethylene
     for possible carcinogenicity.  DHEW Pub. No (NIH) 77-813.  U.S.  Depart-
     ment of Health, Education, and Welfare,  Public Health Service,  National
     Institutes of Health, Bethesda, MD.

National Cancer Institute  (NCI).  (19775)  Bioassay of lindane for  possible
    'carcinogenicity.  DHEW Pub. No (NIH) 77-814.  U.S. Department  of Health,
     Education, and Welfare, Public Health Service, National Institutes of
     Health, Bethesda, MD.

National Toxicology Program (NTP).  (1983)  Bioassay of tetrachloroethylene  in
     female B6C3F1 mice.  Draft (not peer-reviewed).

Parker, J.C.; Davidson, I.W.F.  (1984)   Extrapolation of  the carcinogenic
     response (abstract).  In:  Uncertainty in risk assessment, risk  manage-
     ment, and decision making: proceedings of symposium  held by the  Society
     of Risk Analysis,  p. 38.

Pegg, D.G.; Zempel, J.A.; Braun, W.H.;  Watanabe, P.G.  (1979)  Disposition of
     tetrachloro(l^C)ethylene following oral  and inhalation exposure  in rats.
     Toxicol. Appl. Pharmacol. 51:465-474.

Rail, D.P.   (1977)  Species differences in carcinogenesis testing.   In:
     Origin of human cancers, Book C.  Hiatt, H.H.; Watson, J.D.; Winsten, J.A.
     (eds).  New York: Cold Spring Harbor Laboratory, pp. 1382-1390.

Rampy, L.W.; Quast, J.F.; Balmer, M.F.; Leong, B.K.J.; Gehring, P.J.
     (1978)  Results of a  long-term inhalation toxicology study on  rats of a
     perchloroethylene (tetrachloroethylene)  formulation.  Toxicology Research
     Laboratory, Dow Chemical Co., Midland, MI.

Schumann, A.M.; Quast, J.F.; Watanabe,  P.G.  (1980)  The  pharmacokinetics and
     macromolecular interactions of perchloroethylene in  mice and rats  as
     related to oncogenicity.  Toxicol. Appl. Pharmacol.  55:207-219.

Schlesselman, J.T.  (1974)  Sample size requirements in cohort and  case-control
     studies of disease.  Am. J. Epidenriol. 99(6).

Theiss, J.C.; Stoner, G.D.; Shimkin, M.D.; Weisburger, E.K.   (1977)   Tests for
     carcinogenicity of organic contaminants  of United States drinking waters
     by pulmonary tumor response in strain A mice.  Cancer Res. 37:2717-2720.
                                      9-77

-------
U.S. Environmental Protection  Agency.   (1984)  Proposed guidelines  for carci
     nogen risk assessment.  Federal  Register 49(227):46294-46301.

Van Duuren, B.L.; Goldschmidt,  M.;  Loewengart, 6.; Smith, C.; Melchione,  S.;
     Seidman, I.; Roth, D.   (1979)   Carcinogenicity of halogenated  olefinic
     and aliphatic hydrocarbons in  mice.   J.  Natl. Cancer Inst. 63(6):1433-
     1439.

Yllner, S.  (1961)  Urinary  metabolites  of 14C-tetrachloroethylene  in  mice.
     Nature 191:820-821.
                             4U.S. GOVERNMENT PRINTING OFFICE: 1986-646-116-40626
                                      9-78

-------